A Generalized-Rate Model for Describing and Scaling Redox Kinetics

Mar 30, 2018 - Batch experiments were performed to characterize the rates of Cr(VI) reduction in four Fe(II)-containing sediments and their assemblage...
0 downloads 9 Views 859KB Size
Subscriber access provided by - Access paid by the | UCSB Libraries

Environmental Modeling

A Generalized Rate Model for Describing and Scaling Redox Kinetics in Sediments Containing Variable Redox Reactive Materials Fen Xu, Yuanyuan Liu, and Chongxuan Liu Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b06354 • Publication Date (Web): 30 Mar 2018 Downloaded from http://pubs.acs.org on March 31, 2018

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 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 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.

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 32

Environmental Science & Technology

1

A Generalized Rate Model for Describing and Scaling Redox Kinetics in Sediments

2

Containing Variable Redox Reactive Materials

3

Fen Xu1, 2†, Yuanyuan Liu2, 3†, and Chongxuan Liu2,4*

4

1

5

University of Technology, Chengdu, 610059, China

6

2

7

3 Key

8

Engineering, Nanjing University, Nanjing, 210023, China

9

4School

10

State Key Laboratory of Geohazard Prevention and Geoenvironment Protection, Chengdu

Pacific Northwest National Laboratory, Richland, WA 99354 Laboratory of Surficial Geochemistry (Ministry of Education), School of Earth Sciences and

of Environmental Science and Engineering, Southern University of Science and

Technology, Shenzhen, 518055, China

Submit to Environmental Science & Technology

11

12

†These

13

contributed to the modeling part.

14

*

15

of Science and Technology, Shenzhen, China. Email: [email protected]

authors contributed equally. Xu, F. contributed to the experimental part and Liu, Y.

Corresponding author: School of Environmental Science and Engineering, Southern University

1

ACS Paragon Plus Environment

Environmental Science & Technology

16

ABSTRACT: This study developed a generalized modeling approach to describe and scale redox

17

reactions from reactive components to the sediments and their assemblages using Cr(VI) reduction

18

as an example. Batch experiments were performed to characterize the rates of Cr(VI) reduction in

19

four Fe(II)-containing sediments and their assemblages. The experimental data were first used to

20

calibrate a generalized rate model of Cr(VI) reduction with generic rate parameters. The

21

generalized rate model was then used to describe the kinetics of Cr(VI) reduction in the sediment

22

assemblages by linearly scaling the rate parameters from the individual sediments. By comparing

23

with the experimental results, this study found that the generalized rate model with generic rate

24

parameters can describe Cr(VI) reduction in individual sediments and their assemblages with

25

different redox reactivity toward Cr(VI) reduction. The sediment-associated Fe(II) and its

26

reactivity were found to be the key variables in the generalized model for describing the Cr(VI)

27

reduction in the studied sediments. A three-step extraction method was subsequently developed to

28

estimate the rate-specific Fe(II) pools that can facilitate the application of the scaling approach in

29

field systems.

30

2

ACS Paragon Plus Environment

Page 2 of 32

Page 3 of 32

31

Environmental Science & Technology

INTRODUCTION:

32

A sediment is a composite of minerals or reactive components that have different properties

33

of reactivity with respect to mineral dissolution and precipitation, redox transformation, and

34

interaction with metals and contaminants.1-6 The heterogeneity of geochemical reactivity in the

35

sediment poses a significant challenge to scale geochemical reaction parameters from laboratory

36

to field-scale applications when a reactive transport model is used to predict geochemical processes

37

in natural system.2, 4, 6-14 Multi-rate model has been widely used in reactive transport models to

38

address the scaling challenge.2, 3, 5, 7, 15-20 Generalized composite (GC) and component additivity

39

(CA) are two such concepts for scaling sorption processes in heterogeneous sediments.1, 21-28 The

40

GC approach assumes that the sorptive properties of a sediment can be described using generic

41

sorption reactions with parameters determined by fitting experimental data. However, the model

42

parameters obtained from one field site cannot be applied to the others.29-31 The CA approach

43

assumed that sorption to a sediment can be described by linearly adding the sorption to the

44

individual components in the sediment.1, 26, 27, 32, 33 However, the applicability of the CA approach

45

is limited because of the complexity of minerals and their interactions in a sediment that are often

46

difficult to identify in the field. An effective approach to scale sorption process and parameters has

47

yet to be developed.

48

Relative to sorption process, the scaling behavior of redox reaction is understudied. In

49

natural sediments, various reductants (Fe(II), S(II), organic materials, etc.) can coexist that can

50

reduce redox sensitive metals and contaminants such as Cr(VI), Tc(VII), U(VI), and As(V) etc.3,

51

34-40

52

various reactive components,2, 16, 18, 20 or single component with species of different reactivity such

53

as Fe(II).3, 34, 35 These models are mathematically analogue to the GC approach for scaling sorption

Multi-rate models have been used to describe reaction properties in a sediment containing

3

ACS Paragon Plus Environment

Environmental Science & Technology

54

processes. Because of the difficulty to identify all the redox species and their redox reactivity in

55

natural sediments, the rate constants fitted from one sediment are usually not applicable to other

56

sediments, even at the same field site. Development of effective approaches to scale redox

57

processes is therefore critically needed to extrapolate laboratory redox reactions for field-scale

58

applications.

59

In this study, we evaluated a generalized multirate model with linear parameter additivity

60

concept to describe and scale redox kinetics determined from individual sediments to their

61

assemblages using Cr(VI) reduction as an example. Four individual sediments with different redox

62

reactivity collected from the Columbia River hyporheic zone (HZ) at the US DOE’s Hanford Site

63

were first used to experimentally study Cr(VI) reduction, which were then used to calibrate the

64

generalized multirate model with a set of generic rate constants. The individual sediments were

65

then composited with different mass ratios to experimentally study Cr(VI) reduction kinetics in

66

the assemblages. The experimental results were used to evaluate the applicability of the

67

generalized multirate model with the set of generic rate constants from the individual sediments.

68

The results demonstrated the applicability of the generalized rate model. The key variables in the

69

generalized multirate model for the studied sediments are the Fe(II) contents in different rate

70

pools.34 A three-step extraction method was subsequently developed to experimentally quantify

71

these variables so that the generalized model with the generic rate constants can be readily

72

validated or directly used for field-scale applications of Cr(VI) reduction by sediment-associated

73

Fe(II).

74

MATERIALS AND METHODS

75

Sediment and Sample Preparation. Sediments were collected from the Columbia River

76

HZ at the US Department of Energy’s Hanford site 300 Area (Figure 1). The detailed information

4

ACS Paragon Plus Environment

Page 4 of 32

Page 5 of 32

Environmental Science & Technology

77

about the study site and the redox properties of the sediments from our previous study are provided

78

in supporting information (SI).

79

The first sediment, which mainly consists of silt and clay, was frozen-collected in the HZ

80

(Figure 1B, white circle) and was stored at –80 °C until use. The frozen collection method was

81

described elsewhere.36 This sediment sample was named as Sediment A, hereafter.

82

The second sediment was a composite of 20 sediments frozen-collected from 20 different

83

locations in the HZ (Figure 1, red circles). The collected sediments were sieved and C>B>A). The reduction

192

rate in Sediment D was significantly higher than that in Sediment C (Figure 2c and d). Within 20

193

hours, 400 μM Cr(VI) was reduced by Sediment D; while only 40 μM Cr(VI) was reduced by

194

Sediment C. Although HCl-extractable Fe(II) in Sediment D was only 4 times higher than in

195

Sediment C, Cr(VI) reduction rate in Sediment D was 10 times faster than in Sediment C. The

196

result suggested that the freshly generated Fe(II) (biogenic Fe(II)) was more reactive than that in

197

the original sediment. The importance of biogenic Fe(II) on Cr(VI) reduction has been

198

demonstrated in previous studies.34, 35, 42, 48

199

The experimental results of Cr(VI) reduction in the sediments were described using the

200

generalized three-rate model developed in our previous studies (Eqs 1 and 2).34, 35 In the previous

201

study,35 we showed that at least three rates were needed to describe Cr(VI) reduction in the

202

sediments from the Hanford HZ. The fraction of rate component i ( f i ) and its corresponding rate

203

constant (ki and KCr(VI)) were fitted from the experimental data (Figure 2a−d). First, the parameters

204

were estimated by the initial and late time reduction kinetics. Specifically, the fraction of the fastest

205

rate component and its rate constant (k1) were estimated from the initial rate of the measured Cr(VI)

206

reduction profiles (data within the first ~5 h). Then, the fraction of Fe(II) in the slowest rate

207

component and its rate constant (k3) were estimated from the measured Cr(VI) reduction profiles

208

at the late times (after ~20 h). The rest of the Fe(II) in the sediment was assigned to the second

209

rate component with a rate constant k2, which was estimated by fitting the overall Cr(VI) reduction

210

profile. In the generalized rate model, all the sediments share the same set of rate constants, but

211

with different fractions of Fe(II) in different rate components. Consequently, all the parameters in

10

ACS Paragon Plus Environment

Page 10 of 32

Page 11 of 32

Environmental Science & Technology

212

the generalized model were further tuned to minimize the overall least error between the measured

213

and calculated Cr(VI) concentrations for all the individual sediments in Figure 2a-d.

214

As shown in Figure 2a-d, the generalized three-rate model can describe Cr(VI) reduction

215

kinetics in Sediment A-D using the same set of rate constants (ki and KCr(VI), i = 1, 2, and 3 in Eqs

216

1 and 2). However, the fraction of Fe(II) in each rate component ( f i m , i = 1, 2, and 3 and m = A,

217

B, C, and D) was different between different sediments, reflecting their overall difference in

218

sediment reactivity toward Cr(VI) reduction. The estimated rate constants (ki =1, 2, 3) for different

219

rate components were different, indicating that each rate component had different reactivity

220

towards Cr(VI) reduction. The rate constant of the fastest component (k1 = 0.4 ± 0.03 M-1∙h-1) was

221

100 times larger than the moderate rate component (k2 = 0.003 ± 0.001 M-1∙h-1), which was 10

222

times larger than the slowest component (k3 = 0.0003 ± 0.0001 M-1∙h-1). The fitted value of KCr(VI)

223

was 1.5 ± 0.2×10-4 M.

224

Cr(VI) reduction rate is proportional to both the initial Fe(II) and its corresponding rate

225

constant for each rate component (Eqs 1 and 2). Since the rate constant for the fastest component

226

was much larger than other components, the overall rate of Cr(VI) reduction quickly decreased

227

when the Fe(II) in the fastest component was exhausted (Figure 2a-d). The average rate constant

228

(< k >) for a sediment was calculated from the following equation: N

229

k   fi k i

(5)

i 1

230

Table 2 shows that the average rate increased with increasing amount of Fe(II) in the fastest

231

component. The overall rate of Cr(VI) reduction in Sediment C was much faster than in Sediment

232

A and B (Figure 2a−c) because Sediment C contained a higher amount of Fe(II) (Table 1) and a

233

larger fraction of Fe(II) was associated with the fastest component than that in Sediment A and B

234

(Table 2). Fresh biogenic Fe(II) in Sediment D is a strong reductant for Cr(VI) reduction.34, 42, 48 11

ACS Paragon Plus Environment

Environmental Science & Technology

235

Model fitting found that 40% of the fresh biogenic Fe(II) in Sediment D was associated with the

236

fastest rate component (Table 2) that was much higher than that in Sediment C (15%).

237

Scaling of Cr(VI) Reduction Kinetics to the Sediment Assemblages. The generalized

238

multirate model was evaluated for scaling Cr(VI) reduction rates with parameters determined from

239

the individual sediments to the sediment assemblages. In the generalized rate model, the rate

240

constants were the same for all the individual sediments so that the rate additivity model (Eq 3)

241

becomes the Fe(II) additivity model for each rate component (Eq 4). With the calculated Fe(II)

242

concentration for each rate component from individual sediments (Tables 1 and 2), equations 1

243

and 2 was used to calculate the rate of Cr(VI) reduction in the assemblages.

244

As a validation of the additivity model for Fe(II), the calculated total Fe(II) in Sediment

245

Assemblage 1, 2, 3, and 4 was compared with HCl-extracted Fe(II) concentrations. The results

246

indicated that the calculated HCl-extractable Fe(II) concentrations generally matched with the

247

measured HCl-extractable Fe(II) concentrations (Table 1).

248

The generalized rate model with linearly scaled Fe(II) concentrations was used to predict

249

Cr(VI) reduction in Sediment Assemblage 1-4 (Figure 2e-h). The predicted Cr(VI) reduction

250

matched well with the experimental results for all the sediment assemblages, indicating that the

251

proposed approach can be applied to scale Cr(VI) reduction with a generic set of rate constants

252

and Fe(II) concentrations scaled from individual sediments. As expected, the rate of Cr(VI)

253

reduction follows an order of Assemblage 4>3>2>1, consistent with their trends in total Fe(II)

254

content and Fe(II) in the fastest rate component. It is also consistent with the average rate constants

255

(k) in the sediment assemblages ranged from 0.114 to 0.152 M-1h-1 in Sediment Assemblage 1 to

256

4 (Table 2).

12

ACS Paragon Plus Environment

Page 12 of 32

Page 13 of 32

Environmental Science & Technology

257

Calculation of Cr(VI) Reduction Rate Using Experimentally Determined Fe(II) in the

258

Sediments. The advantage of the generalized model is that the rate formulation and rate constants

259

can be used to describe the individual sediments and their assemblage. This provides a way to

260

extrapolate the generalized rate for field scale application. The remaining key variables for scaling

261

the model are the total reactive Fe(II) and its fractions in different rate components. These

262

parameters may be difficult to determine in field sediments without extensive characterization of

263

sediment properties. The extraction approach described in the method section is to avoid this

264

problem by directly determining Fe(II) contents in different rate components so that the

265

generalized rate model with rate constants determined from laboratory experiments can be directly

266

applied in field.

267

Three-step extraction method was developed based on the rate of Fe(II) release in the

268

presence of 1 M sodium acetate (Figure S1), which, together with 1 M HCl extraction result, was

269

used to determine Fe(II) concentration for each rate component. The Fe(II) extracted with 1 M

270

sodium acetate includes soluble Fe(II), exchangeable Fe(II), and carbonate-associated Fe(II) that

271

are all reactive toward Cr(VI) reduction.49-52 DI water is often used to extract water soluble fraction

272

of Fe(II), and MgCl2/CaCl2 to extract ion exchangeable Fe(II).45-47 Using these extraction methods,

273

it was found that there was no detectable dissolved Fe(II) in the studied sediments. The

274

exchangeable fractions of Fe(II) were seldom detected in Sediment A, B, and C. The extracted

275

Fe(II) using 1 M sodium acetate (Figure S1) increased with increasing time. The majority of Fe(II)

276

was extracted within the first hour. After 1 h, the extracted Fe(II) concentration only increased

277

slowly until 24 h, and then stabilized after that. Intuitively, Fe(II) extracted by 1 M sodium acetate

278

within the first hour should be readily available for Cr(VI) reduction. The rate-limited release of

279

Fe(II) after the first hour may be due to the effect of chemical heterogeneity (mineral phases that

13

ACS Paragon Plus Environment

Environmental Science & Technology

280

have different reactivity in sodium acetate solution) or physical heterogeneity (mass transfer

281

processes such as diffusion in internal pores of mineral grains). Based on this extraction procedures,

282

Fe(II) extracted within the first hour was used as the Fe(II) in the fastest rate component, and Fe(II)

283

extracted between 1 to 24 h was considered as that in the moderate rate component, and the

284

difference between HCl (1M HCl for 2 h) extractable and sodium acetate (1 M sodium acetate for

285

24 h) extractable Fe(II) was used as the Fe(II) for the slowest rate component. The Fe(II) assigned

286

to three rate components using this method (Table 3) matched well with the additivity model-fitted

287

value (Table 2). Other extraction procedures were tried, but the fractions of Fe(II) determined

288

using the approach as described here matched the best with the model.

289

Using the experimentally determined Fe(II) for each rate component, the simulated results

290

from the generalized rate model with the same rate constants as described before generally matched

291

with the experimental data of Cr(VI) reduction for Sediment A, B, and D (Figure 3a-c). However,

292

the predicted Cr(VI) concentration for Sediment B was lower than experimental results because

293

the experimentally determined fraction of Fe(II) in the fastest component in Sediment B was

294

significantly over-estimated. In general, the average rate constants, , of Sediments A, B, and

295

D predicted using experimentally determined Fe(II) in the rate components (Table 3) was greater

296

than those fitted from Cr(VI) reduction kinetics (Table 2).

297

Cr(VI) reduction in the assemblage sediments was also calculated using the experimentally

298

determined Fe(II) fractions for Sediment A, B, and D and the reactive Fe(II) components fitted

299

from Cr(VI) reduction kinetics for Sediment C. Some of Sediment C was preserved in a stainless

300

steel container and stored in a 4 °C refrigerator for 9 months before extraction method was

301

developed. After 9 month storage in a 4 °C refrigerator, HCl-extractable Fe(II) concentration in

302

Sediment C decreased from 15.42 ± 2.61 to 9.51 ± 0.88 μmol∙g-1 as a result of oxidation by O2.35,

14

ACS Paragon Plus Environment

Page 14 of 32

Page 15 of 32

Environmental Science & Technology

303

40, 53 Since the redox reactivity of the sediment

304

components in Sediment C was fitted from Cr(VI) reduction kinetics. The predicted Cr(VI)

305

reduction in the sediment assemblages matched the major trend of the experimental results of

306

Cr(VI) reduction (Figure 3d-g). However, the prediction over-estimated initial Cr(VI) reduction

307

rate due to over-estimation of Fe(II) in the fastest rate component by the three-step extraction

308

method. The discrepancy increased with the mass ratio of Sediment D because Sediment D

309

contained higher fraction of Fe(II) in the fastest component.

changed during long time storage, the reactive Fe(II)

310

Environmental Implications. Natural sediments usually contain multiple redox reactive

311

species with different reactivity, which requires significant effort to quantify redox kinetics, rate

312

models and rate properties, especially at field site with heterogeneous sediment reaction properties.

313

This study developed a method that used a generalized multirate model concept with a generic set

314

of rate constants to quantitatively describe redox kinetics in sediments with different redox

315

reactivity, and to scale the redox kinetics with parameters determined from individual sediments

316

to sediment assemblages, providing a framework for extrapolating redox properties derived from

317

laboratory studies for field scale applications.

318

Reactive transport models are typically used to simulate and predict solute and contaminant

319

transport in field. In applying a reactive transport model, a simulation domain is typically divided

320

into many numerical grids. The sediment properties within each numerical grid may be

321

heterogeneous, but are often assumed to be homogeneous. Under this condition, the generalized

322

rate model approach described here can be used to calculate the sediment properties within a

323

numerical grid based on the redox properties of a few generic reactive components that can be

324

determined at laboratory. If the mass fractions of such generic reactive components can be

325

measured across the model domain, the redox properties in all the numerical grids can be calculated 15

ACS Paragon Plus Environment

Environmental Science & Technology

326

using the developed approach. In this study, we also developed an extraction approach to

327

determine the mass fractions of the reactive components (i.e., sediment-associated Fe(II) with

328

different reactivity). This would facilitate the application of the generalized approach in reactive

329

transport simulations.

330

The method developed in this study was focused on Cr(VI) reduction by sediment-

331

associated Fe(II) with three generic Fe(II) components. Cr(VI) can be reduced by other

332

reductants such as S(-II),40, 54-56 and organic matters.57, 58 Consequently, the number of the

333

generic reactive components and associated rate constants derived from this study are likely site-

334

specific. However, the concept of the generalized multirate model with a set of generic rate

335

constants to extrapolate laboratory results for field scale applications is applicable for other sites

336

and to other redox sensitive contaminants such as Tc, U, As, etc. For these systems, the

337

generalized rate expression (Eqs 1 and 2) may need to be modified, for example, by increasing

338

the number of rate components to incorporate other redox reaction mechanisms. Further study,

339

especially at the field scale, is needed to further evaluate the applicability of the generalized

340

model framework as proposed here.

341

Supporting Information. The details of experimental method, the chemical composition of

342

SRW, the grain size fraction, and the sodium acetate extraction kinetics are provided in the

343

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

344

http://pubs.acs.org

345

ACKNOWLEDGMENTS

346

This research is supported by the U.S. DOE, Office of Science, Biological and Environmental

347

Research (BER) as part of the Subsurface Biogeochemical Research (SBR) Program through

348

Pacific Northwest National Laboratory (PNNL) SBR Science Focus Area (SFA) Research 16

ACS Paragon Plus Environment

Page 16 of 32

Page 17 of 32

Environmental Science & Technology

349

Project. F Xu, Y Liu, and C Liu also acknowledge the supports from Research Funds from

350

National Natural Science Foundation of China (No.41502233, No. 41572228, NO. 41521001,

351

and No. 41773111) and China Postdoctoral Science Foundation (No. 2015M582305). C Liu also

352

acknowledges the support from Southern University of Science Technology (G61296001) and

353

from Guangdong Provincial Key Laboratory of Soil and Groundwater Pollution Control.

354

17

ACS Paragon Plus Environment

Environmental Science & Technology

355

REFERENCES

356

1. Shang, J.; Liu, C.; Wang, Z.; Zachara, J. M., Effect of grain size on uranium (VI) surface

357

complexation kinetics and adsorption additivity. Environmental Science & Technology 2011, 45,

358

(14), 6025-6031.

359

2. Liu, C.; Shang, J.; Shan, H.; Zachara, J. M., Effect of subgrid heterogeneity on scaling

360

geochemical and biogeochemical reactions: A case of U (VI) desorption. Environmental Science

361

& Technology 2014, 48, (3), 1745-1752.

362

3. Liu, Y.; Liu, C.; Kukkadapu, R. K.; McKinley, J. P.; Zachara, J.; Plymale, A. E.; Miller, M.

363

D.; Varga, T.; Resch, C. T., 99Tc(VII) retardation, reduction, and redox rate scaling in naturally

364

reduced sediments. Environmental Science & Technology 2015, 49, (22), 13403-12.

365

4. Beckingham, L. E.; Mitnick, E. H.; Steefel, C. I.; Zhang, S.; Voltolini, M.; Swift, A. M.;

366

Yang, L.; Cole, D. R.; Sheets, J. M.; Ajo-Franklin, J. B.; DePaolo, D. J.; Mito, S.; Xue, Z.,

367

Evaluation of mineral reactive surface area estimates for prediction of reactivity of a multi-

368

mineral sediment. Geochimica et Cosmochimica Acta 2016, 188, 310-329.

369

5. Zhang, X.; Jiang, B.; Zhang, X., Reliability of the multiple-rate adsorptive model for

370

simulating adsorptive solute transport in soil demonstrated by pore-scale simulations. Transport

371

in Porous Media 2013, 98, (3), 725-741.

372

6. Li, L.; Gawande, N.; Kowalsky, M. B.; Steefel, C. I.; Hubbard, S. S., Physicochemical

373

heterogeneity controls on uranium bioreduction rates at the field scale. Environmental Science &

374

Technology 2011, 45, (23), 9959-9966.

375

7. Liu, C.; Shang, J.; Kerisit, S.; Zachara, J. M.; Zhu, W., Scale-dependent rates of uranyl

376

surface complexation reaction in sediments. Geochimica et Cosmochimica Acta 2013, 105, 326-

377

341. 18

ACS Paragon Plus Environment

Page 18 of 32

Page 19 of 32

Environmental Science & Technology

378

8. Beckingham, L. E., Evaluation of macroscopic porosity-permeability relationships in

379

heterogeneous mineral dissolution and precipitation scenarios. Water Resources Research 2017,

380

53, (12), 10217-10230.

381

9. Steefel, C. I.; DePaolo, D. J.; Lichtner, P. C., Reactive transport modeling: An essential tool

382

and a new research approach for the Earth sciences. Earth and Planetary Science Letters 2005,

383

240, (3), 539-558.

384

10. Li, L.; Peters, C. A.; Celia, M. A., Upscaling geochemical reaction rates using pore-scale

385

network modeling. Adv. Water Resour. 2006, 29, (9), 1351-1370.

386

11. Li, L.; Maher, K.; Navarre-Sitchler, A.; Druhan, J.; Meile, C.; Lawrence, C.; Moore, J.;

387

Perdrial, J.; Sullivan, P.; Thompson, A.; Jin, L.; Bolton, E. W.; Brantley, S. L.; Dietrich, W. E.;

388

Mayer, K. U.; Steefel, C. I.; Valocchi, A.; Zachara, J.; Kocar, B.; McIntosh, J.; Tutolo, B. M.;

389

Kumar, M.; Sonnenthal, E.; Bao, C.; Beisman, J., Expanding the role of reactive transport

390

models in critical zone processes. Earth-Science Reviews 2017, 165, 280-301.

391

12. Li, L.; Steefel, C. I.; Yang, L., Scale dependence of mineral dissolution rates within single

392

pores and fractures. Geochimica et Cosmochimica Acta 2008, 72, (2), 360-377.

393

13. Atchley, A. L.; Navarre-Sitchler, A. K.; Maxwell, R. M., The effects of physical and

394

geochemical heterogeneities on hydro-geochemical transport and effective reaction rates.

395

Journal of Contaminant Hydrology 2014, 165, 53-64.

396

14. Lai, P.; Moulton, K.; Krevor, S., Pore-scale heterogeneity in the mineral distribution and

397

reactive surface area of porous rocks. Chemical Geology 2015, 411, 260-273.

398

15. Haggerty, R.; Gorelick, S. M., Multiple-rate mass transfer for modeling diffusion and

399

surface reactions in media with pore-scale heterogeneity. Water Resources Research 1995, 31,

400

(10), 2383-2400.

19

ACS Paragon Plus Environment

Environmental Science & Technology

401

16. Liu, C.; Shi, Z.; Zachara, J. M., Kinetics of uranium (VI) desorption from contaminated

402

sediments: Effect of geochemical conditions and model evaluation. Environmental science &

403

technology 2009, 43, (17), 6560-6566.

404

17. Fox, P. M.; Davis, J. A.; Hay, M. B.; Conrad, M. E.; Campbell, K. M.; Williams, K. H.;

405

Long, P. E., Rate-limited U(VI) desorption during a small-scale tracer test in a heterogeneous

406

uranium-contaminated aquifer. Water Resources Research 2012, 48, (5).

407

18. Greskowiak, J.; Gwo, J.; Jacques, D.; Yin, J.; Mayer, K. U., A benchmark for multi-rate

408

surface complexation and 1D dual-domain multi-component reactive transport of U(VI).

409

Computational Geoscience 2015, 19, (3), 585-597.

410

19. Berkowitz, B.; Emmanuel, S.; Scher, H., Non-Fickian transport and multiple-rate mass

411

transfer in porous media. Water Resources Research 2008, 44, (3).

412

20. Miller, A. W.; Rodriguez, D. R.; Honeyman, B. D., Upscaling sorption/desorption processes

413

in reactive transport models to describe metal/radionuclide transport: A critical review.

414

Environmental Science & Technology 2010, 44, (21), 7996-8007.

415

21. Filipská H, Š. K., Mathematical modeling of a Cs(I)–Sr(II)–bentonite–magnetite sorption

416

system, simulating the processes taking place in a deep geological repository. Acta Polytechnica

417

2005, 45, (5), 11-18.

418

22. Davis, J. A.; Coston, J. A.; Kent, D. B.; Fuller, C. C., Application of the surface

419

complexation concept to complex mineral assemblages. Environmental Science & Technology

420

1998, 32, (19), 2820-2828.

421

23. Davis, J. A.; Meece, D. E.; Kohler, M.; Curtis, G. P., Approaches to surface complexation

422

modeling of Uranium(VI) adsorption on aquifer sediments. Geochimica et Cosmochimica Acta

423

2004, 68, (18), 3621-3641.

20

ACS Paragon Plus Environment

Page 20 of 32

Page 21 of 32

Environmental Science & Technology

424

24. Kroupova H, S. K., Experimental study and mathematical modeling of Cs(I) and Sr(II)

425

sorption on bentonite as barrier material in deep geological repository. Acta Geodynamica et

426

Geomaterialia 2005, 2, (2), 79-86.

427

25. Payne, T. E.; Brendler, V.; Ochs, M.; Baeyens, B.; Brown, P. L.; Davis, J. A.; Ekberg, C.;

428

Kulik, D. A.; Lutzenkirchen, J.; Missana, T.; Tachi, Y.; Van Loon, L. R.; Altmann, S.,

429

Guidelines for thermodynamic sorption modelling in the context of radioactive waste disposal.

430

Environmental Modelling & Software 2013, 42, 143-156.

431

26. Zhang, X.; Liu, C.; Hu, B. X.; Hu, Q., Grain-size based additivity models for scaling multi-

432

rate uranyl surface complexation in subsurface sediments. Mathematical Geosciences 2015, 48,

433

(5), 511-535.

434

27. Dong, W.; Wan, J., Additive surface complexation modeling of uranium(VI) adsorption onto

435

quartz-sand dominated sediments. Environmental Science & Technology 2014, 48, (12), 6569-

436

77.

437

28. Zheng, Z.; Tokunaga, T. K.; Wan, J., Influence of calcium carbonate on U(VI) sorption to

438

soils. Environmental Science & Technology 2003, 37, (24), 5603-5608.

439

29. Goldberg, S.; Lesch, S. M.; Suarez, D. L., Predicting selenite adsorption by soils using soil

440

chemical parameters in the constant capacitance model. Geochimica et Cosmochimica Acta

441

2007, 71, (23), 5750-5762.

442

30. Kantar, C.; Ikizoglu, G.; Koleli, N.; Kaya, O., Modeling Cd(II) adsorption to heterogeneous

443

subsurface soils in the presence of citric acid using a semi-empirical surface complexation

444

approach. Journal of Contaminant Hydrology 2009, 110, (3), 100-109.

21

ACS Paragon Plus Environment

Environmental Science & Technology

445

31. Petrangeli Papini, M.; Saurini, T.; Bianchi, A.; Majone, M.; Beccari, M., Modeling the

446

competitive adsorption of Pb, Cu, Cd, and Ni onto a natural heterogeneous sorbent material

447

(Italian “Red Soil”). Industrial & Engineering Chemistry Research 2004, 43, (17), 5032-5041.

448

32. Alessi, D. S.; Fein, J. B., Cadmium adsorption to mixtures of soil components: Testing the

449

component additivity approach. Chemical Geology 2010, 270, (1), 186-195.

450

33. Lumsdon, D. G.; Meeussen, J. C. L.; Paterson, E.; Garden, L. M.; Anderson, P., Use of solid

451

phase characterisation and chemical modelling for assessing the behaviour of arsenic in

452

contaminated soils. Applied Geochemistry 2001, 16, (6), 571-581.

453

34. Xu, F.; Liu., Y.; Zachara, J. M.; Bowden, M.; Kennedy, D. W.; Plymale, A. E.; Liu, C.,

454

Redox transformation and reductive immobilization of Cr(VI) in the columbia river hyporheic

455

zone sediment from hanford site. Journal of Hydrology 2017, 555, 278-287.

456

35. Liu, Y.; Xu, F.; Liu, C., Coupled hydro-biogeochemical processes controlling Cr reductive

457

immobilization in Columbia River hyporheic zone. Environmental Science & Technology 2017,

458

51, (3), 1508-1517.

459

36. Moser, D. P.; Fredrickson, J. K.; Geist, D. R.; Arntzen, E. V.; Peacock, A. D.; Li, S. M. W.;

460

Spadoni, T.; McKinley, J. P., Biogeochemical processes and microbial characteristics across

461

groundwater−surface water boundaries of the Hanford Reach of the Columbia River.

462

Environmental science & technology 2003, 37, (22), 5127-5134.

463

37. Hu, N.; Ding, D.-x.; Li, S.-m.; Tan, X.; Li, G.-y.; Wang, Y.-d.; Xu, F., Bioreduction of

464

U(VI) and stability of immobilized uranium under suboxic conditions. Journal of Environmental

465

Radioactivity 2016, 154, 60-67.

22

ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32

Environmental Science & Technology

466

38. Peretyazhko, T.; Zachara, J. M.; Heald, S. M.; Jeon, B. H.; Kukkadapu, R. K.; Liu, C.;

467

Moore, D.; Resch, C. T., Heterogeneous reduction of Tc (VII) by Fe (II) at the solid–water

468

interface. Geochimica et Cosmochimica Acta 2008, 72, (6), 1521-1539.

469

39. Chen, X.; Zeng, X.-C.; Wang, J.; Deng, Y.; Ma, T.; Guoji, E.; Mu, Y.; Yang, Y.; Li, H.;

470

Wang, Y., Microbial communities involved in arsenic mobilization and release from the deep

471

sediments into groundwater in Jianghan Plain, Central China. Science of The Total Environment

472

2017, 579, 989-999.

473

40. Wadhawan, A. R.; Stone, A. T.; Bouwer, E. J., Biogeochemical controls on hexavalent

474

chromium formation in estuarine sediments. Environmental Science & Technology 2013, 47,

475

(15), 8220-8228.

476

41. Stookey, L. L., Ferrozine-a new spectrophotometric reagent for iron. Analytical Chemistry

477

1970, 42, (7), 779-781.

478

42. Brookshaw, D. R.; Coker, V. S.; Lloyd, J. R.; Vaughan, D. J.; Pattrick, R. A., Redox

479

interactions between Cr(VI) and Fe(II) in bioreduced biotite and chlorite. Environmental Science

480

& Technology 2014, 48, (19), 11337-42.

481

43. Liu, C.; Zachara, J. M., Uncertainties of Monod kinetic parameters nonlinearly estimated

482

from batch experiments. Environmental Science & Technology 2001, 35, (1), 133-141.

483

44. Beck, J. V.; Arnold, K. J., Parameter estimation in engineering and science. John Wiley and

484

sons: New York, 1977.

485

45. Tessier, A.; Campbell, P. G. C.; Bisson, M., Sequential extraction procedure for the

486

speciation of particulate trace metals. Analytical Chemistry 1979, 51, (7), 844-851.

23

ACS Paragon Plus Environment

Environmental Science & Technology

487

46. Poulton, S. W.; Canfield, D. E., Development of a sequential extraction procedure for iron:

488

implications for iron partitioning in continentally derived particulates. Chemical Geology 2005,

489

214, (3-4), 209-221.

490

47. Gleyzes, C.; Tellier, S.; Astruc, M., Fractionation studies of trace elements in contaminated

491

soils and sediments: a review of sequential extraction procedures. TrAC Trends in Analytical

492

Chemistry 2002, 21, (6), 451-467.

493

48. Bishop, M. E.; Glasser, P.; Dong, H.; Arey, B.; Kovarik, L., Reduction and immobilization

494

of hexavalent chromium by microbially reduced Fe-bearing clay minerals. Geochimica et

495

Cosmochimica Acta 2014, 133, 186-203.

496

49. Fendorf, S. E.; Li, G., Kinetics of chromate reduction by ferrous iron. Environmental

497

Science & Technology 1996, 30, (5), 1614-1617.

498

50. Eary, L. E.; Rai, D., Chromate removal from aqueous wastes by reduction with ferrous ion.

499

Environmental Science & Technology 1988, 22, (8), 972-977.

500

51. Erdem, M.; Gur, F.; Tumen, F., Cr(VI) reduction in aqueous solutions by siderite. Journal of

501

hazardous materials 2004, 113, (1-3), 217-22.

502

52. McBeth, J. M.; Lloyd, J. R.; Law, G. T. W.; Livens, F. R.; Burke, I. T.; Morris, K., Redox

503

interactions of technetium with iron-bearing minerals. Mineralogical Magazine 2011, 75, (4),

504

2419-2430.

505

53. Morgan, B.; Lahav, O., The effect of pH on the kinetics of spontaneous Fe(II) oxidation by

506

O2 in aqueous solution--basic principles and a simple heuristic description. Chemosphere 2007,

507

68, (11), 2080-4.

508

54. Arias, Y. M.; Tebo, B. M., Cr(VI) reduction by sulfidogenic and nonsulfidogenic microbial

509

consortia. Applied and Environmental Microbiology 2003, 69, (3), 1847-1853.

24

ACS Paragon Plus Environment

Page 24 of 32

Page 25 of 32

Environmental Science & Technology

510

55. Graham, A. M.; Bouwer, E. J., Rates of hexavalent chromium reduction in anoxic estuarine

511

sediments: pH effects and the role of acid volatile sulfides. Environmental Science & Technology

512

2009, 44, (1), 136-142.

513

56. Fendorf, S.; Wielinga, B. W.; Hansel, C. M., Chromium transformations in natural

514

environments: The role of biological and abiological processes in chromium(VI) reduction.

515

International Geology Review 2000, 42, (8), 691-701.

516

57. Wittbrodt, P. R.; Palmer, C. D., Reduction of Cr (VI) in the presence of excess soil fulvic

517

acid. Environmental Science & Technology 1995, 29, (1), 255-263.

518

58. Tokunaga, T. K.; Wan, J.; Firestone, M. K.; Hazen, T. C.; Schwartz, E.; Sutton, S. R.;

519

Newville, M., Chromium diffusion and reduction in soil aggregates. Environmental Science &

520

Technology 2001, 35, (15), 3169-3174.

521

25

ACS Paragon Plus Environment

Environmental Science & Technology

B

A

522 523

Figure 1. Sediment sampling locations (the right panel) at the HZ of the U.S. DOE’s Hanford

524

Site in Washington State. The top left panel was the study area. The bottom left panel was the

525

schematic diagram of the geologic formation of the HZ at the site. Sediment A was frozen

526

collected from an impermeable zone (white circle in the right panel), Sediment B was

527

homogenized from the sediments frozen collected from 20 different locations (red circles), and

528

Sediment C was fresh collected from the shore of the Columbia River (white cross).

26

ACS Paragon Plus Environment

Page 26 of 32

Page 27 of 32

Environmental Science & Technology

529

Table 1. The mass ratios of Sediment A−D in the sediment assemblages and HCl-extractable

530

Fe(II) content in the sediment samples.

Sediment A Sediment B Sediment C Sediment D Sediment Assemblage 1 Sediment Assemblage 2 Sediment Assemblage 3 Sediment Assemblage 4

A:B:C:D 1:0:0:0 0:1:0:0 0:0:1:0 0:0:0:1 1:2:2:1 1:1:1:1 0:1:1:2 0:1:1:4

Measured Fe(II) (μmol∙g-1) 1.52 ± 0.08 4.58 ± 0.35 15.42 ± 2.61 62.28 ± 3.22 15.18 ± 2.02 20.24 ± 1.37 33.25 ± 2.11 38.30 ± 1.42

Predicted Fe(II) (μmol∙g-1) NA NA NA NA 17.15 ± 1.54 20.90 ± 1.57 35.80 ± 2.34 44.45 ± 2.63

531

Measured Fe(II) in the sediments were experimentally determined; predicted Fe(II) in sediment

532

assemblages were calculated from mass ratios and Fe(II) concentrations of the individual

533

sediments.

534

27

ACS Paragon Plus Environment

40 35 30 25 20 15 10 5 0

a) Sediment A

Page 28 of 32

b) Sediment B 60 Cr(VI) (M)

Cr(VI) (M)

Environmental Science & Technology

0

100

200 Time(h)

20

0

300

0

50 Time(h)

1500

c) Sediment C

300

40

100

d) Sediment D

Cr(VI) (M)

Cr(VI) (M)

250 200 150 100

1000

500

50 0

0

100

200

0

300

0

200

Time(h) 800

e) SA1_1:2:2:1

600

Cr(VI) (M)

Cr(VI) (M)

800

400 200 0

600 400 200

0

100

200

300

0

400

0

100

200

300

Time(h) 800

g) SA3_0:1:1:2

600

Cr(VI) (M)

Cr(VI) (M)

800

400 200

535

600

f) SA2_1:1:1:1

Time(h)

0

400 Time(h)

h) SA4_ 0:1:1:4

600 400 200

0

50

100

150

200

250

0

0

100

Time(h)

200

300

Time(h)

536

Figure 2. Measured (symbols) and simulated (lines) Cr(VI) reduction in Sediment A−D and

537

Sediment Assemblage 1−4.

28

ACS Paragon Plus Environment

Page 29 of 32

Environmental Science & Technology

538

Table 2. Fractions of the three redox reactive components with different rate constants in the

539

sediment samples.

Sediment A Sediment B Sediment C Sediment D Sediment Assemblage 1 Sediment Assemblage 2 Sediment Assemblage 3 Sediment Assemblage 4 540 541 542

f1 0.01±0.005 0.01±0.002 0.15±0.005 0.40±0.010 0.28 0.32 0.36 0.38

f2 0.20±0.10 0.20±0.10 0.55±0.17 0.15±0.04 0.28 0.23 0.20 0.17

f3 0.79±0.10 0.79±0.10 0.30±0.17 0.45±0.05 0.44 0.45 0.44 0.45

0.005 0.005 0.062 0.161 0.114 0.131 0.145 0.152

f1, f2, and f3 are the fractions of the fastest, moderate, and slowest rate components. Fraction values in Sediment A-D were model-fitted, and in Sediment Assemblage1 to 4 were calculated from Eq 4.

543

29

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 32

544

Table 3. Fractions of the three reactive Fe(II) components determined by the three-step

545

extraction method and the average rate constants calculated from the experimental determined

546

fractions.

Sediment A Sediment B Sediment C* Sediment D 547 548 549

f1 0.03±0.01 0.07±0.06 0.15±0.005 0.44±0.11

f2 0.16±0.01 0.30±0.09 0.55±0.17 0.27±0.11

f3 0.81±0.01 0.63±0.04 0.30±0.17 0.29±0.03

0.013 0.029 0.062 0.177

f1, f2, and f3 are the fractions of the fastest, moderate, and slowest rate components. *Fractions of Sediment C were fitted from Cr(VI) reduction kinetics because the redox reactivity of the sediment changed during long time storage.

30

ACS Paragon Plus Environment

Page 31 of 32

Environmental Science & Technology

10 0

0

100

200

1000

40

20

0

300

Cr(VI) (M)

20

0

20

Time(h)

60

600 400 200

0

100

300

551 552 553 554

0

200

400 Time(h)

600

e) SA2_1:1:1:1

600 400

0

400

0

100

800

f) SA3_0:1:1:2

600

Cr(VI) (M)

Cr(VI) (M)

200 Time(h)

400 200

550

0

100

200

800

0

80

800

d) SA1_1:2:2:1 Cr(VI) (M)

Cr(VI) (M)

40

500

Time(h)

800

0

c) Sediment D

60 Cr(VI) (M)

Cr(VI) (M)

30

1500

b) Sediment B

a) Sediment A

40

200 Time(h)

300

g) SA4_0:1:1:4

600 400 200

0

50

100

150

200

250

0

Time(h)

0

100

200

300

Time(h)

Figure 3. Experimental and predicted Cr(VI) reduction in Sediment A, B, and D and Sediment Assemblage 1−4. The fractions of the Fe(II) components in Sediment A, B, and D were determined by the three-steps extraction method, and in Sediment Assemblage1 to 4 were calculated from Eq 4.

555

31

ACS Paragon Plus Environment

Environmental Science & Technology

t

3

i  3 ki CFe(II) i 1

800

C

A,i Fe(II)

K Cr(VI)  CCr(VI)

  C m

m ,i Fe(II)

40

Sed. A

20 0

m

0

100 200 300

Time(h)

400 0

0

200

Time(h)

400

Cr(VI) (M)

Cr(VI) (M)

Sed. A:B:C:D=1:2:2:1

CCr(VI)

Cr(VI) (M)

CCr(VI)

Page 32 of 32

1500 1000

Sed. D

500 0

0 200 400 600

Time(h)

ACS Paragon Plus Environment