Biochar and Activated Carbon for Enhanced Trace Organic

Apr 24, 2015 - *E-mail: [email protected]; Phone: (720)-984-2116; Fax: (303)-273-3413. ... Multiple Pathways to Bacterial Load Reduction by Stormwate...
1 downloads 0 Views 2MB Size
Subscriber access provided by NEW YORK UNIV

Article

Biochar and Activated Carbon for Enhanced Trace Organic Contaminant Retention in Stormwater Infiltration Systems Bridget Anne Ulrich, Eugenia Im, David Werner, and Christopher P. Higgins Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.5b00376 • Publication Date (Web): 24 Apr 2015 Downloaded from http://pubs.acs.org on May 3, 2015

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 29

Environmental Science & Technology

1

Manuscript

2 3

BIOCHAR AND ACTIVATED CARBON FOR ENHANCED TRACE ORGANIC

4

CONTAMINANT RETENTION IN STORMWATER INFILTRATION SYSTEMS

5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26

Bridget A. Ulrich,† Eugenia A. Im,† David Werner,§ and Christopher P. Higgins†,* †

Department of Civil and Environmental Engineering Colorado School of Mines, Golden, CO §

School of Civil Engineering and Geosciences Newcastle University, Newcastle upon Tyne, United Kingdom Submitted to Environmental Science & Technology *To whom correspondence should be addressed. E-mail: [email protected] Phone: (720)-984-2116 Fax: (303)-273-3413 Number of pages: Number of figures: Number of tables: Number of words:

28 6 1 7019

27 28 29 30 31 32

ACS Paragon Plus Environment

1

Environmental Science & Technology

Page 2 of 29

33

ABSTRACT

34

To assess the effectiveness of biochar and activated carbon (AC) for enhanced trace

35

organic contaminant (TOrC) retention in stormwater infiltration systems, an approach

36

combining forward-prediction modelling and laboratory verification experiments was

37

employed. Batch and column tests were conducted using representative TOrCs and

38

synthetic stormwater. Based on batch screening tests, two commercially available

39

biochars (BN-biochar and MCG-biochar) and an AC were investigated. The AC

40

exhibited the strongest sorption, followed by MCG-biochar and BN-biochar. Langmuir

41

isotherms provided better fits to equilibrium data than Freundlich isotherms. Due to

42

superior sorption kinetics, 0.2 wt% MCG-biochar in saturated sand columns retained

43

TOrCs more effectively than 1.0 wt% BN-biochar. A forward-prediction intraparticle

44

diffusion model based on the Langmuir isotherm adequately predicted column results

45

when calibrated using only batch parameters, as indicated by a Monte Carlo uncertainty

46

analysis. Case study simulations estimated that an infiltration basin amended with F300-

47

AC or MCG-biochar could obtain sorption-retarded breakthrough times for atrazine of 54

48

years or 5.8 years, respectively, at a 1 in/hr infiltration rate. These results indicate that

49

biochars or ACs with superior sorption capacity and kinetics can enhance TOrC retention

50

in infiltration systems, and performance under various conditions can be predicted using

51

results from batch tests.

52

INTRODUCTION

53

Urbanization has degraded the quality and quantity of urban water resources, and

54

ineffective stormwater management has played a major role.1 Flooding and erosion have

55

worsened as urban landscapes have become less permeable.2 Further, urban stormwater

ACS Paragon Plus Environment

2

Page 3 of 29

Environmental Science & Technology

56

runoff has contaminated receiving waters.3 While occurrence and treatability of metals

57

and nutrients has been widely assessed,4,5 less information is available for trace organic

58

contaminants (TOrCs). It is clear that TOrCs are widely present in urban runoff at

59

harmful levels.6,7 For example, multiple studies have detected the herbicides diuron and

60

atrazine at μg/L levels in roadside and roof runoff, respectively.7–9

61

Distributed stormwater treatment presents an opportunity to mitigate

62

contamination to receiving waters while potentially augmenting water supplies.1,10 For

63

example, stormwater infiltration basins, where runoff is infiltrated rapidly through sand-

64

based media, have become a popular urban stormwater best management practice.11

65

However, some contaminants, polar TOrCs in particular,12 may pass through these

66

systems. For example, less than 50% removal of atrazine was observed during in situ

67

challenge tests for a field-scale vegetated bioinfiltration basin.13 This suggests that

68

sorption by conventional organic amendments (i.e. mulch and compost) and naturally

69

developed biological material does not achieve adequate retention of polar, recalcitrant

70

TOrCs.

71

Engineered sorbents are a promising means of enhancing TOrC retention. Black

72

carbons (BCs), such as activated carbon (AC) and biochar, are especially attractive

73

because their high sorptive affinities enable even small amounts to achieve significant

74

retention. The BC particle size will be an important performance parameter, as larger

75

particles allow faster infiltration while smaller particles allow faster sorption kinetics.

76

However, BC particle size will be difficult to control during construction. The amount of

77

BC added to base infiltration media (i.e., the BC dose in sand) presents a more plausible

78

design variable, as well as a performance trade-off: an increased dose of BC will increase

ACS Paragon Plus Environment

3

Environmental Science & Technology

Page 4 of 29

79

TOrC retention but also may decrease hydraulic conductivity.14 Potential advantages of

80

both biochar and AC make it unclear which is best-suited for this application. Biochar

81

more effectively retains heavy metals,15 acts as a sink for atmospheric carbon dioxide,16

82

and may be less expensive than AC if produced from waste material.17 However,

83

adequate TOrC retention may be maintained at lower AC doses due to its higher sorption

84

capacity, potentially allowing faster infiltration.

85

BC sorption of commonly detected stormwater TOrCs, such as diuron,18

86

benzotriazole,19 and atrazine,18,20 has been widely studied in batch experiments. Sorption

87

capacity and isotherm nonlinearity increase with production temperature, pore volume,

88

and surface area (SA),21–23 and dissolved organic carbon (DOC) suppresses sorption by

89

pore blockage.24–27 In contrast, the effects of BC on TOrC transport in porous media have

90

been less widely studied, and previous studies have focused on hydrophobic TOrCs.28–30

91

Strong kinetic sorption effects have been observed in sediments containing BC, which

92

have been attributed to intraparticle diffusion.31,32 Considering that kinetic effects

93

increase for shorter distances and faster flows,28 kinetic limitations may be especially

94

relevant for infiltration systems.

95

The objective of this study was to verify a forward-prediction model for the

96

sorption-controlled retention of TOrCs in BC-amended sand, such that TOrC

97

breakthrough times for BC-amended infiltration basins could be predicted. It was

98

hypothesized that significant retention of polar TOrCs could be achieved at high

99

infiltration rates and low BC doses, despite fouling by DOC, and further, that kinetic

100

effects could be described by sorption-retarded intraparticle diffusion. The following

101

three questions were addressed: (1) Can the TOrC sorption performance of commercially

ACS Paragon Plus Environment

4

Page 5 of 29

Environmental Science & Technology

102

available BCs be assessed using simple batch experiments? (2) Which forward-prediction

103

transport models, when calibrated using parameters obtained only from batch

104

experiments, best predict TOrC retention in BC-amended sand? (3) Can BC dose be

105

adjusted in infiltration basins to simultaneously achieve effective TOrC retention and fast

106

infiltration? To address these questions, a series of batch and column tests were

107

conducted using commercially available BCs, synthetic stormwater containing DOC, and

108

representative polar stormwater TOrCs.

109

EXPERIMENTAL SECTION

110

Materials. Eighteen BCs (characterization and production details in Tables S1 and S2)

111

were initially screened (described below), and two commercially available biochars and

112

one AC were selected for further evaluation in model verification experiments. The two

113

wood-based biochars were a pyrolysis biochar produced by Biochar Now in Berthoud,

114

CO (BN-biochar; 108 m2/g SA) and a gasification biochar produced by Mountain Crest

115

Gardens in Etna, CA (MCG-biochar; 318 m2/g SA). The AC was Calgon Filtersorb® 300

116

(F300-AC; 883 m2/g SA). BC SA and pore volume were determined by N2 adsorption,

117

and solid densities (dBC) were estimated by helium adsorption, both at 77 K using an

118

automated surface area analyzer. Micropore volume was determined from N2 adsorption

119

data using the t method.33 BC intraparticle porosity (pBC) was estimated gravimetrically

120

by water uptake. Small BC particle diameters (dpart) were used to achieve homogeneous

121

surface properties (less than 246 μm for screening experiments, between 53 and 246 μm

122

for model verification experiments), and BCs were rinsed with deionized water and air

123

dried to remove aggregated small particles. The suite of representative stormwater TOrCs

124

included 2,4-diphenoxyacetic acid (2,4-D), benzotriazole, atrazine, diuron, oryzalin,

ACS Paragon Plus Environment

5

Environmental Science & Technology

Page 6 of 29

125

tris(3-chloro-2-propyl)phosphate (TCPP), prometon, and fipronil (properties and standard

126

sources in Table S3), and was selected according to occurrence in urban stormwater,

127

recalcitrance, and mobility.6

128

TOrC Analysis. Aqueous TOrC concentrations were quantified by liquid

129

chromatography – tandem mass spectrometry (LC-MS/MS), using a 1200 Series binary

130

pump (Agilent), a CTC Analytics HTS PAL autosampler (LEAP Technologies) with a 1

131

mL sample loop, a 150 x 46 mm Luna C18 column with a 5 µm particle size

132

(Phenomenex), and a 3200 QTrap MS/MS (ABSciex). TOrCs were analyzed by

133

electrospray ionization (ESI) in positive or negative mode (details in Table S4). Eluents

134

prepared with HPLC-grade water and OptimaTM methanol (Fisher Scientific) were

135

pumped at an overall flow rate of 0.8 mL/min throughout the gradient method. The ESI

136

positive eluent contained 2 mM ammonium formate (Sigma Aldrich) and 0.1% formic

137

acid (Fluka), and the ESI negative eluent contained 4 mM ammonium acetate

138

(Mallinckrodt). Isotope dilution was performed as described elsewhere.34 Briefly,

139

aqueous aliquots were diluted by 10% with a solution of methanol containing 4 μg/L

140

isotope surrogate, such that the final surrogate concentration was 0.4 μg/L.

141

General Methodology for Batch Experiments. Batch experiments were carried out in

142

triplicate in amber bottles containing BC and synthetic stormwater. Synthetic stormwater

143

was prepared from deionized water and salts to obtain total carbonate, nitrogen, and

144

phosphorus levels representative of event mean concentrations in urban runoff (Table

145

S5).3 Its pH was initially adjusted to 7, and moderate buffering capacity from 1 mM

146

bicarbonate allowed observation of effects from BC-induced pH changes. The synthetic

147

stormwater was either “clean” or contained added DOC (source varied, described below).

ACS Paragon Plus Environment

6

Page 7 of 29

Environmental Science & Technology

148

The volume of synthetic stormwater (200 mL or 400 mL) and mass of BC (1 mg to 30

149

mg) were adjusted to achieve a fractional TOrC uptake between 0.2 and 0.8 and limit

150

uncertainty in calculated partition coefficients (Kd, Calculation S1).35 Sodium azide was

151

also added to the synthetic stormwater to prevent biodegradation and biological fouling.

152

BCs were pre-equilibrated in synthetic stormwater overnight to simulate abiotic fouling,

153

then TOrCs were added to batches with a methanol carrier, such that methanol

154

concentrations were less than 0.1 vol%. During sample collection, 1530 μL aliquots were

155

diluted with isotope surrogate solution, centrifuged, and transferred to autosampler vials.

156

Additional methodology details (i.e., pH effects, QA/QC) are provided in the supporting

157

information (Method S1).

158

Batch Screening. Batch screening experiments were conducted to identify commercially

159

available biochars to represent a realistic range of performance in model verification

160

experiments, using a reproducible and broadly applicable method for BC selection. Only

161

wood-based biochars were evaluated due to their commercial availability and

162

demonstrated sorption capacity,17 and F300-AC was evaluated in tandem to set a

163

performance benchmark. A short equilibration time of 5 days was used to simultaneously

164

assess effects of kinetics and capacity, and the initial aqueous TOrC concentration (Cw,0)

165

was 10 μg/L. Initial tests were conducted in clean synthetic stormwater for 17 biochars

166

and F300-AC. Eight biochars that spanned the observed performance range were

167

subsequently tested using a reproducible ‘standard’ synthetic stormwater, which

168

contained 20 mg/L Suwanee River Natural Organic Matter (SRNOM; International

169

Humic Substances Society, St. Paul, MN) and 500 mg/L sodium azide. BN-biochar and

ACS Paragon Plus Environment

7

Environmental Science & Technology

Page 8 of 29

170

MCG-biochar were selected for further evaluation because they spanned the full range of

171

observed performance.

172

Model Calibration. Batch kinetic and isotherm experiments were conducted in a

173

representative matrix to obtain calibration parameters. Experiments were carried out over

174

67 days, and bottles contained 3 mg MCG-biochar, 25 mg BN-biochar, or 1 mg F300-AC

175

in 200 mL ‘representative’ synthetic stormwater (elemental composition in Table S6)

176

containing 200 mg/L sodium azide. This matrix was prepared by adding DOC to clean

177

synthetic stormwater with a concentrated stock to obtain DOC levels representative of a

178

first flush event (10 – 30 mg/L).36 Five gallons of runoff were collected during an August

179

storm from a residential curb in Golden, CO (DOC 2.5 mg/L; Sievers TOC Analyzer),

180

equilibrated for one month with 53 g aspen leaves and 100 g grass from the bank of Clear

181

Creek (Golden, CO), and 450 g plant-based EcoGro compost (A1 Organics; Eaton, CO),

182

then filtered to 0.45 μm (118 mg/L DOC). For kinetic experiments, sampling occurred

183

every 1 – 14 days, and the Cw,0 and DOC were 10 µg/L and 10 mg/L, respectively. For

184

isotherm experiments, Cw,0 ranged from 3 – 100 µg/L, and the DOC was 30 mg/L. To

185

further evaluate DOC effects, sorption kinetics were also assessed for MCG-biochar at 30

186

mg/L DOC, and 67-day Kd values were also determined at 10 mg/L DOC and 10 µg/L

187

Cw,0 for the three BCs. Only ESI positive TOrCs were analyzed in model calibration and

188

verification experiments due to constraints with respect to sample quantity, and models

189

were not calibrated for diuron because its sorbed fraction was greater than 0.8 in all

190

batches.

191 192

Parameters describing equilibrium sorption as either linear or nonlinear were obtained from the equilibrium data. Linear Kd values (Calculation S1) were averaged

ACS Paragon Plus Environment

8

Page 9 of 29

Environmental Science & Technology

193

from the linear isotherm regions, and best-fits to the Langmuir (Equation S1) and

194

Freundlich isotherms (Equation S2) were obtained using Igor Pro.37 As the residual mean

195

square error (RMSE) demonstrated better fits for the Langmuir isotherm (Table S7), the

196

nonlinear sorption parameters were the Langmuir coefficient (Kl) and maximum TOrC

197

concentration on the BC (Cs,max). Further, two kinetic parameters were obtained from the

198

kinetic data. Tortuosity was used to describe intraparticle diffusion kinetics, as obtained

199

from the best-fit of a sorption-retarded diffusive uptake model (Equation S3).38 A first-

200

order rate for TOrC mass transfer to the BC was also estimated by Equation S4. In

201

addition to these sorption parameters, transport models were also calibrated for BC

202

particle properties (pBC, dBC, dpart) and column properties (porosity, n; dispersion

203

coefficient, Edisp; described below).

204

Transport Models. Four forward-prediction models describing one-dimensional,

205

sorption-retarded solute transport in porous media were evaluated: (I) a model assuming

206

local sorption equilibrium (Equation S5); (II) a model assuming first-order sorption

207

kinetics (Equation S6); and two models describing sorption-retarded intraparticle

208

diffusion kinetics, assuming (III) linear sorption, and (IV) Langmuir sorption (Equation

209

S7). Models I – III have been previously verified for sediments containing BC.28 Model

210

IV uses the same partial differential equations as Model III, with modifications to account

211

for Langmuir sorption (Equation S8).

212

Model Verification. Column experiments were conducted with 15 cm length x 2.5 cm ID

213

Kimble-Chase Kontes Chromaflex columns (Fischer Scientific) and a standard sand

214

(Sigma Aldrich; 2.54 cm3/g solid density) with 210 – 297 μm diameter particles, such

215

that the ratio of column diameter to particle size exceeded the minimum value of 50

ACS Paragon Plus Environment

9

Environmental Science & Technology

Page 10 of 29

216

recommended to limit boundary effects.39 A 12 cm portion of each column was dry-

217

packed with a sand-BC mixture, which was fixed between 2 cm sand plugs at the inlet

218

and outlet. There was no visible BC particle migration in the outlet sand plugs. Different

219

doses were used for each BC (0.4 wt% F300-AC, 0.2 wt% MCG-biochar, 1.0 wt% BN-

220

biochar) due to wide variability in TOrC retention and constraints with respect to

221

experimental duration. A homogeneous F300-AC distribution could not be achieved at a

222

lower dose, thus effluent concentrations (Cw,effluent) were much lower and results are

223

reported as breakthrough pore volumes (Cw,effluent above quantification limit, 0.025 µg/L).

224

The porosity of the BC-amended column portion (nBC 0.35 ± 0.1, Table 1) was estimated

225

gravimetrically after saturation, and corrected to account for the porosity in the sand

226

plugs (nsand 0.32).

227

Columns were flushed with CO2 and then deionized water in upflow to ensure

228

saturation. Tracer tests were carried out to confirm that early breakthrough did not occur

229

using KBr as a conservative tracer and a Hach® conductivity meter. Edisp values were

230

obtained from best-fits to tracer data using Hydrus 1D (Table 1). Control columns

231

without BC confirmed less than 5% loss of TOrCs to column components. A

232

multichannel peristaltic pump delivered influent at 0.42 mL/min (2.1 ± 0.2 in/hr or 5.3 ±

233

0.5 cm/hr linear velocity) to represent infiltration in a basin sized to drain in 24 hr

234

(recommended rates for sizing calculations are 4.9 and 1.3 in/hr for sand and loamy sand,

235

respectively).40 A pulsed influent profile was carried out to observe kinetic effects (i.e.,

236

fronting and tailing during breakthrough and desorption, respectively). Columns were

237

equilibrated overnight with representative synthetic stormwater (10 mg/L DOC, 200

238

mg/L sodium azide) to allow pre-equilibration with DOC, pulsed with TOrCs (Cw,influent

ACS Paragon Plus Environment

10

Page 11 of 29

Environmental Science & Technology

239

20 µg/L) for 4 days, then flushed with TOrC-free representative synthetic stormwater for

240

8 days. Effluent samples were collected at 1 – 30 pore volume intervals throughout the

241

pulse and desorption phases, and isotope dilution was carried out directly in autosampler

242

vials. The effluent pH ranged from 7.2 – 8.5, and varied by less than 0.1 log units

243

throughout the experiments.

244

Transport models were verified by comparison of column data and model

245

predictions, based on a Monte Carlo analysis evaluating the uncertainty of outputs from

246

Models III and IV, as described elsewhere.28 Briefly, the standard deviation of the model

247

output was determined after 50 simulations, using parameter inputs which were normally

248

distributed around their means with estimated (30% of the mean parameter) or

249

experimental standard deviations (Table 1). Uncertainty bounds were generated by

250

adding and subtracting the model-output standard deviation from the mean-parameter

251

output.

252

Case Study. Case study simulations were carried out to predict sorption-controlled

253

atrazine breakthrough times for an infiltration basin amended with F300-AC or MCG-

254

biochar, and to assess the potential performance trade-off associated with increased BC

255

dose (increased TOrC retention and decreased hydraulic conductivity). Ranges of

256

possible BC doses were identified by setting minimum specifications for two cases: (Case

257

1) a maximum breakthrough time (when 95% removal can no longer be maintained) at a

258

minimum infiltration rate of 1 in/hr, and (Case 2) a maximum infiltration rate for a

259

minimum breakthrough time of 1 year. Simulations with Model IV were carried out for

260

an infiltration basin with an underdrain, such that the infiltration rate was the unsaturated

261

hydraulic conductivity. The dependence of saturated hydraulic conductivity on BC dose

ACS Paragon Plus Environment

11

Environmental Science & Technology

Page 12 of 29

262

was estimated (Calculation S2) according to results from falling head tests (Method S2)

263

and previously observed trends,14 using sand dosed with F300-AC and MCG-biochar.

264

The unsaturated hydraulic conductivity was estimated assuming 90% saturation

265

(Calculation S3).41,42 The simulated basin was sized to 103 m2 to allow 12 hr drainage of

266

50% runoff volume from a 3 acre catchment during an 80th percentile storm (Calculation

267

S4).40 Annual treatment rates for a catchment receiving 16 inches of precipitation per

268

year were estimated as 65 pore volumes and 78 pore volumes for basins with depths of 3

269

feet and 1 foot, respectively (Calculation S5). To account for long-term fouling,

270

Langmuir parameters were approximated as half of experimental values, and tortuosity

271

was doubled (Table S12).24,43 An average BC dpart of 1 mm was simulated to represent

272

increased kinetic effects for larger particles. To account for channeling, the estimated

273

pore volumes to breakthrough was halved to approximate a 50% bed contact volume

274

(30% reduction previously observed for a conservative tracer).44

275

RESULTS AND DISCUSSION

276

Batch Screening. Batch screening tests revealed trends in the 5-day LogKd dependent on

277

TOrC LogKow and BC surface properties, and the presence of DOC weakened trends.

278

Representative results are shown in Figure 1 (additional results in Table S8). LogKd

279

values increased with TOrC LogKow, and LogKd was suppressed in the presence of 20

280

mg/L SRNOM (Figure 1A). Diuron had the highest LogKd, though its LogKow is similar

281

to the other TOrCs, while 2,4-D had the lowest LogKd. Benzotriazole also had a

282

relatively high LogKd despite a low LogKow, particularly for BN-biochar. Among the

283

different BCs, significant correlations (Pearson p-values < 0.05, Table S9) with LogKd

284

values were observed for N2 pore volume (Figure 1B) for most TOrCs, and for SA in

ACS Paragon Plus Environment

12

Page 13 of 29

Environmental Science & Technology

285

comparatively fewer cases. Further, peak LogKd values were observed for mesopore

286

volumes between 0.1 and 0.2 cm3/g (Figure 1C), potentially due to counteracting effects

287

of faster diffusion and fewer sorption sites in mesopores. MCG-biochar performed

288

especially well, achieving LogKd values similar to F300-AC.

289

These results show that previously identified TOrC sorption trends can be

290

extended to commercially available biochars in the presence of DOC.22 In light of the

291

difficulties predicting performance from measureable BC properties, BC selection will

292

likely be most successful if based on batch tests, and the method described herein may

293

provide a reproducible and broadly applicable BC selection method. Nevertheless, if

294

batch tests are not possible, N2 pore volume and mesopore volume may indicate sorption

295

performance.

296

Model Calibration. Best-fits of the sorption-retarded kinetic model to the apparent

297

LogKd (LogKd,app) for benzotriazole are shown in Figure 2A (additional results in Figure

298

S1). The kinetic model fit the batch data well; the SSR for the fit to LogKd,app was below

299

0.1 for all cases (Table 1). TOrC sorption to MCG-biochar appeared to reach equilibrium

300

at 30 days (no detectable change Kd,app), while sorption of some TOrCs to BN-biochar

301

and F300-AC did not reach equilibrium within 67 days (atrazine and benzotriazole for

302

BN-biochar; atrazine, prometon, and TCPP for F300-AC). For these cases, both

303

equilibrium LogKd (LogKd,eq) and tortuosity were adjusted, resulting in LogKd,eq values

304

that were 3% - 12% greater than the 67-day LogKd,app values (Figure S1). Kinetics were

305

fastest (i.e. lowest tortuosity) for MCG-biochar, followed by F300AC and BN-biochar

306

(Table 1). Overall, higher DOC and lower mesopore volume resulted in slower kinetics

ACS Paragon Plus Environment

13

Environmental Science & Technology

307

(Table S10), suggesting that pre-equilibration with DOC might increase intraparticle

308

mass transfer resistances by pore blockage.45

309

Page 14 of 29

Best-fits of the Langmuir isotherm to batch data for benzotriazole are shown in

310

Figure 2B (additional results in Figure S2) and Langmuir parameters are shown in Table

311

1. Though the Freundlich isotherm has been widely applied for TOrC sorption to BC,46

312

previous studies have primarily used clean matrices. It is likely that pre-equilibration with

313

DOC affected the sorption mechanism; pre-equilibration with humic substances has

314

previously been observed to increase isotherm linearity for sorption of non-polar aromatic

315

compounds to biochar.27 The highest sorption capacity was achieved by F300-AC,

316

followed by MCG-biochar and BN-biochar. LogKd values were lower in the presence of

317

30 mg/L DOC than 10 mg/L DOC, indicating that higher DOC levels suppressed sorption

318

(Table S11). LogKd suppression increased with molecular size (TCPP >prometon

319

>atrazine >benzotriazole), as previously observed.27 Benzotriazole and atrazine exhibited

320

the greatest nonlinearity in the relevant Cw range, as indicated by high Kl and low Cs,max.

321

While the isotherm parameters may be affected by the isotherm Cw range (only points

322

with fsorbed >0.2 are included), all isotherms included the maximum Cw in column tests

323

(20 μg/L), thus sufficiently accounting for the expected nonlinearity.

324

Model Verification. Figure 3 shows representative pulsed breakthrough curves for

325

columns containing (A) 1.0 wt% BN-biochar and (B) 0.2 wt% MCG-biochar. MCG-

326

biochar retained TOrCs more effectively than BN-biochar. Atrazine, prometon, and

327

TCPP, with LogKow values between 2.5 and 3.0, behaved similarly. Though its LogKow is

328

also in this range, diuron showed especially enhanced retention. Further, BN-biochar

329

showed relatively strong retention of benzotriazole, despite its lower LogKow of 1.4. For

ACS Paragon Plus Environment

14

Page 15 of 29

Environmental Science & Technology

330

columns with 0.4 wt% AC, the breakthrough order was TCPP, prometon (both at 363

331

pore volumes), atrazine (370 pore volumes), and benzotriazole (428 pore volumes);

332

diuron breakthrough was not observed within 428 pore volumes. These results indicate

333

that BCs with high sorption capacities and fast kinetics effectively retain TOrCs, and that

334

LogKow does not indicate TOrC breakthrough order.

335

Strong kinetic effects were observed for all column experiments, indicated by

336

fronting and tailing in breakthrough curves. Figure 4 shows predictions from all four

337

models for prometon using the mean parameter values shown in Table 1 (additional

338

results in Figure S3). Models I (first-order kinetic) and II (local equilibrium)

339

overestimated and underestimated retention, respectively, while Models III (diffusion-

340

limited, linear) and IV (diffusion-limited, Langmuir) were in better agreement with

341

column results. Predictions were generally improved when models were calibrated using

342

tortuosity values from batches containing 30 mg/L DOC (example for atrazine in Figures

343

S4A and S4B), possibly because elevated DOC levels better represented increased

344

fouling in column systems. TCPP was the exception to this trend (Figures S4C and S4D).

345

As TCPP sorption was most suppressed by DOC (Table S11), a threefold higher DOC

346

level may have overestimated fouling effects. Because kinetic parameters were not

347

additionally determined at 30 mg/L DOC for BN-biochar, the models could not calibrated

348

to account for increased fouling, and TOrC retention is over-predicted for BN-biochar.

349

To characterize the significance of discrepancies between model predictions and

350

observations, a Monte Carlo uncertainty analysis was carried out for Models III and IV.

351

Figure 5 shows that for MCG-biochar columns, most deviations from mean model

352

predictions fell within the uncertainty bounds (additional results in Figures S5 – S9). This

ACS Paragon Plus Environment

15

Environmental Science & Technology

Page 16 of 29

353

agreement is notable in light of inherent differences between batch and column sorption

354

processes, and the complexity of the representative synthetic stormwater. The upper

355

uncertainty bounds were in better agreement with BN-biochar breakthrough curves

356

(Figures S6 – S9), but still over-predicted retention, suggesting that calibration at higher

357

DOC levels is necessary.

358

Trends in nonlinearity observed during isotherm experiments were reflected in

359

model predictions. Models III and IV gave similar predictions for TCPP (Figures 5A and

360

5B) and prometon (Figure S5), as expected due to their observed sorption linearity for Cw

361

up to 20 μg/L. There were more pronounced differences between predictions from

362

Models III and IV for atrazine (Figures 5C and 5D) and benzotriazole (Figures 5E and

363

5F), which exhibited stronger nonlinearity in the relevant Cw range (i.e. lower Cs,max and

364

higher Kl for sorption to MCG-biochar). Model IV predicted sharper peaks than Model

365

III, potentially indicative of BC being at or near sorption capacity. For atrazine, the

366

uncertainty bounds accounted for most deviations from the predictions for both models.

367

For benzotriazole, which showed the greatest nonlinearity in the relevant Cw range, only

368

Model IV predicted the observed sharp peak, though a comparatively larger portion of the

369

observed curve fell outside of the uncertainty bounds. These deviations may be due to

370

differences in the ionized fraction of benzotriazole between batch and column

371

experiments, since the pH was near its pKa of 8.2.

372

Case Study Simulation. Figure 6 shows the results of the case study predicting atrazine

373

breakthrough times for an infiltration basin with varying doses of F300-AC and MCG-

374

biochar, assuming that no biodegradation occurs and infiltration rate is determined by BC

375

dose. When BC dose is increased, infiltration rate decreases and atrazine breakthrough

ACS Paragon Plus Environment

16

Page 17 of 29

Environmental Science & Technology

376

time increases, representing the performance trade-off associated with increased dose.

377

Infiltration basins amended with both MCG-biochar and F300-AC were predicted to meet

378

the minimum design criteria set for Case 1 (1 in/hr minimum infiltration rate) and Case 2

379

(1 year minimum breakthrough time). For F300-AC, a maximum breakthrough time of 54

380

years at a dose of 12.3 wt% AC was estimated for Case 1, and a maximum infiltration

381

rate of 4.3 in/hr at a dose of 4.9 wt% was estimated for Case 2. For MCG-biochar, a

382

maximum breakthrough time of 5.8 years at a dose of 4.5 wt% was estimated for Case 1,

383

and a maximum infiltration rate of 3.0 in/hr at a dose of 2.9 wt% was estimated for Case

384

2.

385

These results suggest that both biochar and AC could obtain atrazine

386

breakthrough times greater than five years by sorption only, and that AC may prevent

387

breakthrough for multiple decades. However, to dose a 100 m2 infiltration basin at 12.3

388

wt% AC would require approximately 11.8 tons, which may be cost prohibitive with AC

389

prices near $1500/ton.47 A dose of 5 wt% (4.8 tons) may be more reasonable. Further,

390

application of biochar may have economic benefits, as break-even costs below $250/ton

391

have been estimated for production from waste material.17 Additionally, in light of the

392

superior metals retention and carbon sequestration benefits of biochar,15,16 it is unclear if

393

this additional predicted retention makes AC more advantageous. If sufficient space were

394

available to construct a basin to drain at 1 in/hr, the longer hydraulic retention time of the

395

biochar-amended media may allow increased biodegradation and reduced kinetic effects.

396

Environmental Implications. A significant outcome of this study was the verification of

397

a forward-prediction model to describe sorption-retarded transport of TOrCs in saturated

398

BC-amended sand. This model may broadly enable the design of treatment systems for

ACS Paragon Plus Environment

17

Environmental Science & Technology

Page 18 of 29

399

targeted removal of a wide variety of TOrCs under diverse operating conditions, as

400

calibration requires only batch data. Initial case study simulations predicted that either

401

biochar or AC could effectively enhance polar TOrC retention in stormwater infiltration

402

basins, providing a potentially cost-effective means of mitigating contamination of urban

403

receiving waters. Further, the described methodology for batch screening experiments

404

may provide a reproducible means of screening BCs for diverse applications. This

405

outcome may also be especially significant in light of the wide performance variability of

406

commercially available biochars evidenced in this study.

407

Additional work is necessary to address several potential opportunities and

408

challenges. Frequent media replacement and disposal may be avoided if retained TOrCs

409

can be degraded.48 Recent work has suggested that amendment of soils with biochar

410

enhances the biodegradation of some TOrCs,49 motivating further studies with

411

biologically active media. Further, difficulties predicting the effects of long-term fouling

412

by DOC, clogging, and channeling may present engineering challenges, and field

413

validation experiments will be necessary to ensure that adequate TOrC removal can be

414

achieved under actual environmental conditions.

415

Supporting information. This material is available free of charge via the Internet at

416

http://pubs.acs.org, and contains a comprehensive summary of batch and column data,

417

model outputs, calculations, and methodology details, as is referenced throughout the

418

text.

419

Author information

420

Corresponding author. Christopher P. Higgins

ACS Paragon Plus Environment

18

Page 19 of 29

Environmental Science & Technology

421

Acknowledgements.

422

Science Foundation Graduate Research Fellowship under Grant No. DGE-1057607 and

423

the National Science Foundation Engineering Research Center Program under

424

Cooperative Agreement EEC-1028968 (ReNUWIt). Collaboration with Newcastle

425

University was facilitated by EPSRC grant EP/K021737/1. The authors also thank Dr.

426

David Rutherford for SA analyses and helpful discussions.

427

REFERENCES

428 429 430

(1)

Hering, J. G.; Waite, T. D.; Luthy, R. G.; Drewes, J. E.; Sedlak, D. L. A changing framework for urban water systems. Environ. Sci. Technol. 2013, 47, 10721– 10726.

431 432

(2)

Tsihrintzis, V. A.; Hamid, R. Modeling and Management of Urban Stormwater Runoff Quality: A Review. Water Resour. Manag. 1997, 11, 136–164.

433 434

(3)

Ellis, J. B.; Hvitved-Jacobsen, T. Urban drainage impacts on receiving waters. J. Hydraul. Res. 1996, 34, 771–783.

435 436

(4)

Makepeace, D. K.; Smith, D. W.; Stanley, S. J. Urban stormwater quality: Summary of contaminant data. Crit. Rev. Environ. Sci. Technol. 1995, 25, 93–139.

437 438 439

(5)

Pitt, R. E.; Maestre, A. Stormwater quality as described in the National Stormwater Quality Database ( NSQD ). In 10th International Conference on Urban Drainage; Copenhagen, Denmark, 2005.

440 441 442

(6)

Grebel, J. E.; Mohanty, S. K.; Torkelson, A. A.; Boehm, A. B.; Higgins, C. P.; Maxwell, R. M.; Nelson, K. L.; Sedlak, D. L. Engineered Infiltration Systems for Urban Stormwater Reclamation. Environ. Eng. Sci. 2013, 30, 437–454.

443 444 445

(7)

Bucheli, T. D.; Müller, S. R.; Heberle, S.; Schwarzenbach, R. P. Occurrence and behavior of pesticides in rainwater, roof runoff, and artificial stormwater infiltration. Environ. Sci. Technol. 1998, 32, 3457–3464.

446 447 448

(8)

Sidhu, J.; Gernjak, W.; Toze, S. Health Risk Assessment of Urban Stormwater Edited by : Urban Water Security Research Alliance Technical Report No . 102. 2012.

449 450

(9)

Huang, X.; Fong, S.; Deanovic, L.; Young, T. M. Toxicity of herbicides in highway runoff. Environ. Toxicol. Chem. 2005, 24, 2336–2340.

This material is based upon work supported by the National

ACS Paragon Plus Environment

19

Environmental Science & Technology

Page 20 of 29

451 452 453

(10)

Fletcher, T. D.; Deletic, A.; Mitchell, V. G.; Hatt, B. E. Reuse of Urban Runoff in Australia: A Review of Recent Advances and Remaining Challenges. J. Environ. Qual. 2008, 37, S – 116.

454 455

(11)

Davis, A. P.; Hunt, W. F.; Traver, R. G.; Clar, M. Bioretention Technology: Overview of Current Practice and Future Needs. J. Environ. Eng. 2009, 109–117.

456 457

(12)

Pitt, R.; Clark, S.; Field, R. Groundwater contamination potential from stormwater infitration practices. Urban Water 1999, 1, 217–236.

458 459 460

(13)

Zhang, K.; Randelovic, A.; Page, D.; McCarthy, D. T.; Deletic, A. The validation of stormwater biofilters for micropollutant removal using in situ challenge tests. Ecol. Eng. 2014, 67, 1–10.

461 462 463

(14)

Barnes, R. T.; Gallagher, M. E.; Masiello, C. a; Liu, Z.; Dugan, B. Biocharinduced changes in soil hydraulic conductivity and dissolved nutrient fluxes constrained by laboratory experiments. PLoS One 2014, 9, e108340.

464 465 466 467

(15)

Beesley, L.; Moreno-Jiménez, E.; Gomez-Eyles, J. L. Effects of biochar and greenwaste compost amendments on mobility, bioavailability and toxicity of inorganic and organic contaminants in a multi-element polluted soil. Environ. Pollut. 2010, 158, 2282–2287.

468 469

(16)

Lehmann, J.; Gaunt, J.; Rondon, M. Bio-char sequestration in terrestrial ecosystems - A review. Mitig. Adapt. Strateg. Glob. Chang. 2006, 11, 403–427.

470 471 472

(17)

Ahmad, M.; Rajapaksha, A. U.; Lim, J. E.; Zhang, M.; Bolan, N.; Mohan, D.; Vithanage, M.; Lee, S. S.; Ok, Y. S. Biochar as a sorbent for contaminant management in soil and water: a review. Chemosphere 2014, 99, 19–33.

473 474 475

(18)

Cornelissen, G.; Haftka, J.; Parsons, J.; Gustafsson, O. Sorption to black carbon of organic compounds with varying polarity and planarity. Environ. Sci. Technol. 2005, 39, 3688–3694.

476 477 478

(19)

Reemtsma, T.; Miehe, U.; Duennbier, U.; Jekel, M. Polar pollutants in municipal wastewater and the water cycle: occurrence and removal of benzotriazoles. Water Res. 2010, 44, 596–604.

479 480 481 482

(20)

Ding, L.; Snoeyink, V. L.; Mariñas, B. J.; Yue, Z.; Economy, J. Effects of powdered activated carbon pore size distribution on the competitive adsorption of aqueous atrazine and natural organic matter. Environ. Sci. Technol. 2008, 42, 1227–1231.

483 484 485

(21)

Zhu, D.; Kwon, S.; Pignatello, J. J. Adsorption of single-ring organic compounds to wood charcoals prepared under different thermochemical conditions. Environ. Sci. Technol. 2005, 39, 3990–3998.

ACS Paragon Plus Environment

20

Page 21 of 29

Environmental Science & Technology

486 487 488

(22)

Lattao, C.; Cao, X.; Mao, J.; Schmidt-Rohr, K.; Pignatello, J. J. Influence of molecular structure and adsorbent properties on sorption of organic compounds to a temperature series of wood chars. Environ. Sci. Technol. 2014, 48, 4790–4798.

489 490 491

(23)

Chen, B.; Zhou, D.; Zhu, L. Transitional adsorption and partition of nonpolar and polar aromatic contaminants by biochars of pine needles with different pyrolytic temperatures. Environ. Sci. Technol. 2008, 42, 5137–5143.

492 493 494 495

(24)

Kwon, S.; Pignatello, J. J. Effect of natural organic substances on the surface and adsorptive properties of environmental black carbon (char): pseudo pore blockage by model lipid components and its implications for N2-probed surface properties of natural sorbents. Environ. Sci. Technol. 2005, 39, 7932–7939.

496 497 498

(25)

Kasozi, G. N.; Zimmerman, A. R.; Nkedi-Kizza, P.; Gao, B. Catechol and humic acid sorption onto a range of laboratory-produced black carbons (biochars). Environ. Sci. Technol. 2010, 44, 6189–6195.

499 500 501 502

(26)

De Ridder, D. J.; Verliefde, a R. D.; Heijman, S. G. J.; Verberk, J. Q. J. C.; Rietveld, L. C.; van der Aa, L. T. J.; Amy, G. L.; van Dijk, J. C. Influence of natural organic matter on equilibrium adsorption of neutral and charged pharmaceuticals onto activated carbon. Water Sci. Technol. 2011, 63, 416–423.

503 504 505 506

(27)

Pignatello, J. J.; Kwon, S.; Lu, Y. Effect of natural organic substances on the surface and adsorptive properties of environmental black carbon (char): attenuation of surface activity by humic and fulvic acids. Environ. Sci. Technol. 2006, 40, 7757–7763.

507 508 509

(28)

Werner, D.; Karapanagioti, H. K.; Sabatini, D. A. Assessing the effect of grainscale sorption rate limitations on the fate of hydrophobic organic groundwater pollutants. J. Contam. Hydrol. 2012, 129-130, 70–79.

510 511 512

(29)

Lefevre, G. H.; Novak, P. J.; Hozalski, R. M. Fate of naphthalene in laboratoryscale bioretention cells: implications for sustainable stormwater management. Environ. Sci. Technol. 2012, 46, 995–1002.

513 514 515

(30)

Gidley, P. T.; Kwon, S.; Yakirevich, A.; Magar, V. S.; Ghosh, U. Advection dominated transport of polycyclic aromatic hydrocarbons in amended sediment caps. Environ. Sci. Technol. 2012, 46, 5032–5039.

516 517

(31)

Ball, W. P.; Roberts, P. V. Long-term sorption of halogenated organic chemicals by aquifer material. 1. Equilibrium. Environ. Sci. Technol. 1991, 25, 1223–1237.

518 519 520

(32)

Werner, D.; Ghosh, U.; Luthy, R. G. Modeling polychlorinated biphenyl mass transfer after amendment of contaminated sediment with activated carbon. Environ. Sci. Technol. 2006, 40, 4211–4218.

ACS Paragon Plus Environment

21

Environmental Science & Technology

Page 22 of 29

521 522

(33)

Lippens, B. C.; De Boer, J. H. Studies on pore systems in catalysts V. The t method. J. Catal. 1965, 4, 319–323.

523 524 525

(34)

Vanderford, B. J.; Snyder, S. A. Analysis of pharmaceuticals in water by isotope dilution liquid chromatography/tandem mass spectrometry. Environ. Sci. Technol. 2006, 40, 7312–7320.

526 527

(35)

Sander, M.; Lu, Y.; Pignatello, J. J. A thermodynamically based method to quantify true sorption hysteresis. J. Environ. Qual. 2005, 34, 1063–1072.

528 529 530

(36)

Sickman, J. O.; Zanoli, M. J.; Mann, H. L. Effects of urbanization on organic carbon loads in the Sacramento River, California. Water Resour. Res. 2007, 43, 1– 15.

531 532

(37)

WaveMetrics Support: Demos http://www.wavemetrics.com/support/demos.htm (accessed Apr 16, 2015).

533 534 535

(38)

Grathwohl, P. Section 3.1: Intraparticle Diffusion. In Diffusion in Natural Porous Media: Contaminant Transport, Sorption/Desorption and Dissolution Kinetics; Springer Science+Business Media, LLC: New York, NY, 1998; pp. 44–67.

536 537 538

(39)

Lang, J. S.; Giron, J. J.; Hansen, A. T.; Trussell, R. R.; Hodges, W. E. Investigating filter performance as a function of the ratio of filter size to media size. J. / Am. Water Work. Assoc. 1993, 85, 122–130.

539 540

(40)

UDFCD. Urban Storm Drainage Criteria Manual Volume 3 - Best Mangement Practices; Water Resources Publications, LLC: Denver, CO, 2010.

541 542

(41)

Campbell, G. A Simple Method For Determining Uunsaturated Conductivity from Moisture Retention Data. Soil Sci. 1974, 117, 311–314.

543 544

(42)

Clapp, R. B.; Hornberger, G. M. Empirical Equations for Some Soil Hydraulic Properties. Water Resour. Res. 1978, 14, 601–604.

545 546

(43)

Yang, Y.; Sheng, G. Pesticide adsorptivity of aged particulate matter arising from crop residue burns. J. Agric. Food Chem. 2003, 51, 5047–5051.

547 548

(44)

Padilla, I. Y.; Yah, T. J.; Conklin, M. H. The effect of water content on solute transport in unsaturated porous media. Water Resour. Res. 1999, 35, 3303–3313.

549 550 551

(45)

Hale, S. E.; Tomaszewski, J. E.; Luthy, R. G.; Werner, D. Sorption of dichlorodiphenyltrichloroethane (DDT) and its metabolites by activated carbon in clean water and sediment slurries. Water Res. 2009, 43, 4336–4346.

552 553

(46)

Cornelissen, G.; Gustaffsson, O.; Bucheli, T. D.; Jonker, M. T. O.; Van Noort, P. C. M. Extensive Sorption of Organic Compounds to Black Carbon , Coal , and

ACS Paragon Plus Environment

22

Page 23 of 29

Environmental Science & Technology

554 555 556

Kerogen in Sediments and Soils: Mechanisms and Consequences for Distribution, Bioaccumulation, and Biodegradation. Environ. Sci. Technol. 2005, 39, 6881– 6895.

557 558 559

(47)

USEPA. Estimating Total Capital Investment. In EPA Air Pollution Control Cost Manual; United States Environmental Protection Agency: Research Triangle Park, North Carolina, 1999; pp. 1–24.

560 561 562

(48)

Lefevre, G. H.; Hozalski, R. M.; Novak, P. J. The role of biodegradation in limiting the accumulation of petroleum hydrocarbons in raingarden soils. Water Res. 2012, 46, 6753–6762.

563 564 565

(49)

Tong, H.; Hu, M.; Li, F. B.; Liu, C. S.; Chen, M. J. Biochar enhances the microbial and chemical transformation of pentachlorophenol in paddy soil. Soil Biol. Biochem. 2014, 70, 142–150.

566

FIGURES

567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584

Figure 1. Representative results from batch screening tests in clean and standard synthetic stormwater (SSW). (A) Five-day LogKd versus LogKow for sorption of eight TOrCs to MCG-biochar. (B) Five-day LogKd versus N2 total pore volume for sorption of fipronil to eight BCs. (C) Five-day LogKd versus mesopore volume for sorption of TCPP to eight BCs.

ACS Paragon Plus Environment

23

Environmental Science & Technology

585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609

Page 24 of 29

Figure 2. Representative model calibration results for benzotriazole (complete results in Figures S1 and S2). (A) Sorption-retarded diffusion model fit to batch kinetic data. (B) Langmuir isotherm fit to batch equilibrium data. Dotted lines indicate a 95% confidence interval for the isotherm fit.

610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630

ACS Paragon Plus Environment

24

Page 25 of 29

631 632 633 634 635 636 637 638 639

Environmental Science & Technology

Table 1. Parameters used for model predictions, as determined from batch calibration experiments (ATZ = atrazine, BTA = benzotriazole, PRN = Prometon). The standard deviation values used for the Monte Carlo uncertainty analysis are shown in parentheses. The SSR is for the fit of the sorption-retarded diffusion model to experimental LogKd,app, and the RMSE is for the fit of the Langmuir model to experimental Cs. KINETIC

τ [-]

SSR [L/kg]

LINEAR LANGMUIR ISOTHERM LogKd

KL

Cs,max

RMSE

[L/kg]

[L/µg]

[g/kg]

[g/kg]

3.8 (0.2) 4.5 (0.2) 3.5 (0.2) 3.6 (0.1)

0.059 (0.018) 0.290 (0.087) 0.049 (0.019) 0.013 (0.010)

0.15 (0.05) 0.23 (0.07) 0.12 (0.04) 0.37 (0.11)

4.9 (0.3) 5.0 (0.2) 4.6 (0.2) 4.6 (0.1)

0.145 (0.043) 0.160 (0.036) 0.031 (0.009) 0.0027 (0.001)

1.13 (0.34) 1.14 (0.34) 1.83 (0.54) 14.8 (4.4)

5.7 (0.3) 5.5 (0.2) 5.4 (0.2) 5.4 (0.2)

0.099 (0.030) 0.113 (0.034) 0.065 (0.020) 0.068 (0.020)

6.5 (2.0) 5.1 (1.5) 5.7 (1.7) 14.0 (4.2)

COLUMN -4

Edisp x10

PARTICLE

[cm2/s]

n

pBC

[-]

[-]

[g/cm3]

dBC

1.4 (0.1)

0.36 (0.01)

0.6 (0.1)

1.5 (0.1)

3.1 (0.2)

0.34 (0.01)

0.56 (0.1)

1.7 (0.1)

3.4 (0.2)

0.35 (0.01)

0.77 (0.1)

2.0 (0.1)

BN-BIOCHAR ATZ BTA PRN TCPP

636 (191) 1546 (464) 382 (115) 442 (133)

0.0055 0.0334 0.0048 0.0500

0.010 0.010 0.012 0.144

MCG-BIOCHAR 17.1 (5.1) 28.7 (8.6) 16.4 (4.9) 8.2 (2.4)

ATZ BTA PRN TCPP

0.0080 0.0067 0.0020 0.0538

0.087 0.083 0.073 0.117

F300-AC ATZ BTA PRN TCPP

96.4 (28.9) 11.7 (3.5) 94.2 (28.3) 57.2 (17.2)

0.0848 0.0195 0.0122 0.0296

0.525 0.371 0.414 0.619

640

ACS Paragon Plus Environment

25

Environmental Science & Technology

Page 26 of 29

641 642 643 644 645 646

Figure 3. Representative pulsed breakthrough curves for columns packed with (A) 1.0 wt% BN-biochar and (B) 0.2 wt% MCG-biochar, for a 100 pore volume pulse of 20 μg/L TOrCs, followed by a 328 pore volume flush. Column experiments were conducted in duplicate or triplicate, results are shown for a single experiment to improve clarity.

647 648 649 650 651

Figure 4. Observed prometon effluent profiles for columns packed with (A) 0.2 wt% MCG-biochar and (B) 1.0 wt% BN-biochar, shown with predictions using mean parameters from Models I (equilibrium), II (first-order kinetic), III (linear, diffusionlimited), and IV (Langmuir, diffusion-limited).

ACS Paragon Plus Environment

26

Page 27 of 29

652 653 654 655 656

Environmental Science & Technology

Figure 5. MCG-biochar column breakthrough curves and uncertainty analysis results for Models III (linear, diffusion-limited) and IV (Langmuir, diffusion-limited), as determined for TCPP (A,B), atrazine (C,D), and benzotriazole (E,F). Solid lines and dotted lines indicate the mean-parameter output and uncertainty bounds, respectively.

ACS Paragon Plus Environment

27

Environmental Science & Technology

657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675

Page 28 of 29

Figure 6. Results from a case study on atrazine breakthrough in BC-sand beds with depths between 1 foot and 3 feet for (A) MCG-biochar and (B) F300-AC. Imax and Imin are the maximum and minimum infiltration rates, and BTmax and BTmin are the maximum and minimum breakthrough times, respectively. Infiltration rates are set by BC dose, assuming that hydraulic conductivity decreases with increased BC dose. For Case 1, the solid lines represent the design constraint of a 1 in/hr Imin, and the dotted lines indicate the corresponding predicted BTmax. For Case 2, the solid lines represent the design constraint of a 1 year BTmin, and the dotted lines indicate the corresponding BC dose predicted for Imax. The shaded area indicates the feasible design region.

ACS Paragon Plus Environment

28

Page 29 of 29

676 677

Environmental Science & Technology

TOC Image.

ACS Paragon Plus Environment

29