fe(ii)-initiated reduction of hexavalent chromium in...

22
1 Fe(II)-initiated Reduction of Hexavalent Chromium in 2 Heterogeneous Iron Oxide Suspension 3 4 Jeongyun Choi 1 Yoojin Jung 2 , and Woojin Lee 1* 5 6 1 Department of Civil and Environmental Engineering, Korea Advanced Institute of Science and 7 Technology, 373-1 Guseong-Dong, Yuseong-Gu, Daejeon 305-701, Korea 8 2 Department of Civil and Environmental Engineering, University of Delaware, Newark, DE 19716, 9 USA 10 11 *Corresponding author; phone: 82-42-869-3624; fax; 82-42-869-3610; e-mail: 12 [email protected] 13 14 15

Upload: truonganh

Post on 25-Aug-2018

216 views

Category:

Documents


0 download

TRANSCRIPT

1

Fe(II)-initiated Reduction of Hexavalent Chromium in 2

Heterogeneous Iron Oxide Suspension 3

4

Jeongyun Choi 1 Yoojin Jung

2, and Woojin Lee

1* 5

6

1 Department of Civil and Environmental Engineering, Korea Advanced Institute of Science and 7

Technology, 373-1 Guseong-Dong, Yuseong-Gu, Daejeon 305-701, Korea 8

2 Department of Civil and Environmental Engineering, University of Delaware, Newark, DE 19716, 9

USA 10

11

*Corresponding author; phone: 82-42-869-3624; fax; 82-42-869-3610; e-mail: 12

[email protected] 13

14

15

1

Abstract 16

The characteristics of Fe(II)-initiated reduction of Cr(VI) in iron oxide suspensions were 17

investigated by conducting a series of kinetic experiments. A modified Langmuir-Hinshelwood 18

kinetic model was used to provide the better description of Cr(VI) reduction kinetics which were 19

believed to be occurred on the limited reactive site of reductant. The concentration of magnetite 20

concentration as well as Cr(VI) concentration, significantly affected the reaction kinetics of Cr(VI). 21

The reduction kinetics were improved with increasing magnetite and Cr(VI) concentration. Almost 22

95% of Cr(VI) reduction was achieved within 10 min at the condition of 8 g/L of magnetite and 80 23

mg/L of initial Cr(VI), respectively. The solution pH also affected the reaction rate in the range of 24

5.5 and 8.0 where a lower pH produced a faster reaction rate. The addition of Fe(II) on soil and 25

magnetite showed the capability of improving Cr(VI) reduction kinetics and their reduction kinetics 26

was also well described using a Langmuir-Hinshelwood kinetic model. The experimental results 27

obtained in this research clearly show the advantage of additional reductant for reducing Cr(VI), and 28

they can provide basic knowledge for the development of remediation technology for the treatment 29

of groundwater and soil contaminated with Cr(VI). 30

31

Keywords: Chromium(VI), Modified Langmuir-Hinshelwood kinetics, Magnetite, Reduction, Fe(II) 32

addition 33

34

35

36

2

1. Introduction 37

Chromium generally exists in four oxidation states (Cr(0), Cr(II), Cr(III), and Cr(IV)) in natural 38

and engineered environments, but the trivalent and hexavalent forms are predominant. Most studies 39

have focused on Cr(VI) because of its toxic, carcinogenic, and mutagenic characteristics [1]. In 40

contrast to the harmful characteristics of Cr(VI), Cr(III) that is immobile and insoluble in subsurface 41

environments has been known to be less toxic and even an essential nutrient for plants and animals 42

[2]. Therefore, the redox chemistry reducing Cr(VI) to Cr(III) has been applied to the development of 43

most remedial technologies to treat groundwater and soil contaminated with Cr(VI). Natural 44

attenuation and in-situ redox manipulation are examples of the application of redox chemistry to 45

remediate contaminated sites. Surface-bound natural organic matters (NOMs) and iron-bearing soil 46

minerals such as magnetite, green rust, and zero-valent iron significantly affect the reduction of 47

Cr(VI) [3,4,5]. 48

As one of recent remedial alternatives to expedite natural attenuation, in-situ redox 49

manipulation by injecting reducing agents such as Fe(II) and sulfide into monitoring wells has been 50

studied. The significant reduction of organic compounds has been observed in in-situ redox 51

manipulation system as shown in soil mineral suspensions with reductants [6]. It has been reported 52

that heterogeneous surface reactions cause reductive degradations of organic compounds. Kinetic 53

rate constants for reductive degradation were influenced by the density of reactive Fe(II) bound on 54

the mineral surfaces. Williams and Scherer [7] showed that Fe(II) bound on the mineral surfaces 55

results in enhanced reductive degradation of organic compounds. In contrast, Cr(VI) can be 56

stoichiometrically reduced in the aqueous Fe(II) solution without any solid surfaces (except under 57

high pH conditions) forming reactive precipitates e.g., (CrxFe1-x)(OH)3(s)) [8]. Based on previous 58

experimental results, the Cr(VI) reduction seems to be affected by both aqueous Fe(II) and reactive 59

solids in soil mineral suspensions. 60

Research has been conducted to investigate the reaction kinetics for the reduction of Cr(VI) in 61

soil mineral suspensions with and without Fe(II) addition. However, few research has been carried 62

3

out to fully characterize the reduction of Cr(VI) at the reactive surfaces of soil minerals. This 63

research has investigated the reaction mechanism for the reduction of Cr(VI) at reactive solid 64

surfaces and identified the effect of important physicochemical parameters on the reaction kinetics of 65

Cr(VI). The contribution of aqueous Fe(II) and Fe(II) adsorbed on iron-bearing soil minerals to the 66

reduction of Cr(VI) has been also estimated. Synthesized magnetite and Fe(II) have been used as 67

representative iron-bearing soil mineral and reducing agent, respectively. Nontronite and soil have 68

been used for comparison. 69

70

2. Materials and Methods 71

2.1. Chemicals, Soil Minerals, and Soil 72

Reagents or higher grade chemicals were used as received without further treatment. Cr(VI) 73

(K2Cr2O7, 99.5%, Kanto) and FeSO4·7 H2O (99%, Sigma-Aldrich) were used as a target contaminant 74

and a reductant, respectively. FeCl2·4H2O (99~102%, Wako) and Fe(NO3)3·9H2O (99%, Yakuri) 75

were used to synthesize magnetite. Biological buffers, N-(2-hydroxyethyl)-piperazine-N-2-76

ethanosulfunic acid (HEPES, 99.5%, Sigma), 2-(N-Morpholino)ethanesulfonic acid (MES, 99.5%, 77

Sigma), and a mixture of Tris(hydroxymethyl)aminomethane (99.8%, Sigma-Aldrich) and 78

Tris(hydroxymethyl)aminomethane hydrochloride (99%, Sigma) (TRIS buffer), were used to keep 79

the constant pH of aqueous solutions and suspensions. The following chemicals were used for 80

colorimetric analyses: 1,5-Diphenylcarbonohydrazide (Kanto), 3-(2-Pyridyl)-5,6-diphenyl-1,2,4-81

triazine-p,p΄-disulfonic acid monosodium salt hydrate (97%, Aldrich), (NH4)2Fe(SO4)2 (99.9%, 82

Sigma-Aldrich), H2SO4 (95%, Junsei), HNO3 (60%, Junsei), NaOH (96%, Junsei), Na2SO4 (99%, 83

Junsei), ammonium acetate (97%, Yakuri), and hydroxylammonium chloride (97%, Junsei). All 84

chemical reagents were prepared with deaerated deionized water (ddw) by deoxygenating 18 MΩ·cm 85

deionized water with 99.9% nitrogen for two hours. 86

Magnetite was synthesized by modifying the method developed by Taylor at al. [9] and 87

identified using the XRD analysis. The result of XRD was shown in Fig. 1 with the obtained d-space 88

4

vales of the synthesized magnetite which were comparable to the results of magnetite in powder 89

diffraction file. The magnetite was used for a day after the synthesis to preclude aging effect. The 90

preparation and pretreatment of nontronite (Nontrone, France) and soil samples (Seoul, Korea) were 91

conducted by following Lee and Batchelor’s methods [10]. 92

93

2.2. Experimental Procedures 94

Batch kinetic experiments were conducted in 20 mL amber glass vials with open-top cap and 95

silicon septum lined with PTFE film. Magnetite (0.13 g), nontronite (2.5 g), and soil (2 g) were 96

transferred to vials, and 2 mL of 1M HEPES solution was added to adjust suspension pH to 7.2. The 97

total volume of the suspension was 20 mL resulting in concentrations of 6.5, 125, and 100 g/L for 98

magnetite, nontronite, and soil, respectively. An aliquot amount of Cr(VI) stock solution (0.4 mL, 99

1000 mg/L) was spiked to the vials resulting in an initial Cr(VI) concentration of 20 mg/L. 100

Based on the results of the batch kinetic experiments, parametric studies were conducted to 101

identify the effect of important geochemical parameters on the Cr(VI) reduction kinetics. Magnetite 102

was chosen as a representative soil mineral. The effects of magnetite concentration, target compound 103

concentration, and pH were separately investigated. Each parametric study had the following 104

experimental conditions: (1) The concentration of magnetite was varied from 5 to 8 g/L under the 105

initial Cr(VI) concentration of 50 mg L-1

, (2) the concentration of Cr(VI) differed from 20 to 80 106

mg/L, and (3) the pH of magnetite suspensions was adjusted from 5.5 to 8.0 using biological buffers 107

(MES, HEPES, and TRIS, respectively). Based on the study by Alowitz and Scherer, the Cr(VI) 108

reduction by iron metal was dependent on the type of biological buffer [11]. But we think that a 109

biological buffer is better than an inorganic buffer such as phosphate buffer to maintain the pH in 110

this study. Cr(VI) reduction might be less interfered by a biological buffer compared to an inorganic 111

buffer since the chemicals in an inorganic buffer could react with iron and chromium. Sample 112

preparation procedures generally followed that for the batch kinetic experiments described above. 113

5

To identify the effect of Fe(II) addition on the kinetics of Cr(VI) reduction, Fe(II) stock 114

solution was added to magnetite suspensions before spiking Cr(VI) resulting in the desired ratios of 115

ferrous iron to chromate in the range of 0.6 – 3.0. The concentrations of magnetite and Cr(VI) were 116

6.5 g/L and 80 mg/L, respectively. The pH of the mixed suspension was kept constant at 7.2 with a 117

TRIS buffer. Batch kinetic experiments were also conducted to compare the effect of pH in 118

magnetite suspensions with and without Fe(II) in the pH range of 5.5 – 8.0 under the same 119

concentrations of magnetite and Cr(VI). 120

All vials were quickly and tightly capped, mounted on a tumbler, and then completely mixed at 121

7 rpm under room temperature (20 ± 0.5 ℃). After each reaction time, a desorbing agent (0.284g of 122

Na2SO4) was added into the vials to result in 0.1 M sulfate. The suspensions were vigorously mixed 123

on an orbital shaker for 1 day to promote the desorption of Cr(VI) from soil mineral surfaces [12] 124

and centrifuged at 2000 g for five minutes. The supernatant from each vial was used for the 125

measurement of Cr(VI) and iron concentrations in aqueous solution. All samples were prepared in 126

duplicate. Control samples were prepared to check losses due to sorption on walls of vials and 127

septum and volatilization. The control samples were prepared in the same way described above 128

excluding the addition of soil minerals and soil. 129

130

2.3. Analytical Procedures 131

The concentrations of Cr(VI) and iron (Fe(II) and total iron) in the aqueous phase were 132

measured by colorimetric analysis using UV/VIS spectrophotometer (Spectronic® GENESYS™ 2, 133

Thermo Electron Corporation). The Cr(VI) concentration was determined using the 134

diphenylcarbazide method at a wavelength of 540 nm [13]. The concentrations of Fe(II) was directly 135

measured using the ferrozine method at a wavelength of 562 nm [14], and the concentration of total 136

iron was measured as a form of Fe(II) at the same wavelength after reducing Fe(III) to Fe(II) by 137

adding a 10% hydroxylamine solution. The concentration of Fe(III) was calculated by subtracting the 138

Fe(II) concentration from the total iron concentration. 139

6

X-ray diffraction analysis performed using a RigaKu automated diffractometer using Cu Kα 140

radiation. Solids were separated by centrifugation at 2960 g for 5 min and carefully dried in an 141

anaerobic chamber. The samples were scanned between 2.1° to 70° 2θ with a scan speed of 1° 2θ 142

min-1

. 143

144

2.4. Data Treatment for the Reduction Kinetics of Cr(VI) by magnetite 145

Reductive degradation of Cr(VI) by iron-bearing soil minerals and soil were monitored for 146

one hour. Cr(VI) in the soil mineral and soil suspensions showed rapid disappearance initially 147

followed by slow disappearance or constant concentrations depending on the reductive capacity of 148

soil minerals and soil. This indicates that the Cr(VI) reduction kinetics cannot be properly explained 149

by a pseudo-first-order or zero-order rate law. A modified Langmuir-Hinshelwood kinetic model (Eq. 150

1) was used to properly describe Cr(VI) reduction kinetics. This model could give a proper 151

description of kinetic data since the reduction of Cr(VI) might occur on the finite reactive site of 152

magnetite surface [10]. The kinetic model shown below can be obtained by combining a material 153

balance equation in a batch reactor and a rate equation. 154

155

)(

)()(

0

)()(

0

)()(

1

)(

VICr

VICrVICrVICrVICrRCVICrVICr

CK

CCCpCpk

dt

dC

(1) 156

157

where CCr(VI) is the aqueous concentration of Cr(VI) at time t; k is the rate constant for the decay of 158

Cr(VI) at the reactive sites in time-1

; pCr(VI) is the partitioning factor to explain the effect of 159

partitioning of Cr(VI) among aqueous, gas, and solid (soil mineral + vial wall + septum liner) phases 160

assuming instantaneous equilibrium among the three phases; C0

RC is the initial reductive capacity of 161

magnetite for the Cr(VI) in concentration; C0

Cr(VI) is the aqueous concentration of Cr(VI) at time 162

zero; and K is the sorption coefficient of Cr(VI) with units of inverse concentration. 163

164

7

3. Results and Discussion 165

3.1. Reaction Kinetics for the Reduction of Cr(VI) by Iron-bearing Soil Minerals and Soil. 166

Figure 2 shows the reaction kinetics for the reduction of Cr(VI) by magnetite, nontronite, and 167

soil. Cr(VI) was completely degraded in 10 minutes in the magnetite suspension, while more than 168

90% of the initial Cr(VI) was recovered in the soil and nontronite suspensions at the last sampling 169

time. Magnetite having the smallest surface concentration (284.7 m2/L) shows the greatest kinetic 170

rate constant (2.06 min-1

). Although nontronite and soil have approximately 46 and 8 times greater 171

surface concentrations than magnetite, they have smaller rate constants (1.42 min-1

and 1.28 min-1

, 172

respectively). This indicates the surface concentration is not a unique geochemical factor 173

significantly affecting the reduction of Cr(VI) in soil and soil mineral suspensions. On the other hand, 174

the total iron content of magnetite including both Fe(II) (201 mg/g) and Fe(III) (512 mg/g) is 175

approximately 140 and 7 times greater than those of soil and nontronite. Fe(II) content of soil was 176

measured to be less than 1 mg/g, and most iron measured for nontronite was identified as Fe(III). 177

Therefore, the iron content (especially, the surface Fe(II) concentration) could be the more significant 178

factor explaining the great differences in the reduction kinetics of Cr(VI) in the solid suspensions. 179

Peterson et al. showed similar experimental results. Cr(VI) in groundwater was fully reduced to 180

Cr(III) by the soil containing a high concentration of magnetite, but Cr(VI) was rarely reduced by the 181

soil containing natural iron-bearing phyllosilicates [15]. The results obtained in this research and by 182

Peterson et al. [15] confirm the assumption that the reaction kinetics for the reduction of Cr(VI) is 183

significantly influenced by the content of Fe(II) on the surface of soil minerals and soil. The curves 184

fitted by the Langmuir-Hinshelwood kinetic model at Fig. 2 seem to properly explain the reduction 185

kinetics of Cr(VI) in soil mineral and soil suspensions. 186

187

3.2. Effect of Physicochemical Parameters on the Reduction Kinetics of Cr(VI) 188

Figure 3 shows the effect of magnetite concentration on the reduction of Cr(VI) in the 189

magnetite suspension. The removal of Cr(VI) increased 5 times as the content of magnetite increased 190

8

by 1.6 times. This is due to the limited active site and different activity of sites. Almost 95% of 191

reduced Cr(VI) was removed in 0.5 hour from the suspension of different magnetite concentrations, 192

and part of the remaining Cr(VI) was slowly degraded. The result is consistent with previous results 193

[4]. The early rapid removal of Cr(VI) followed by its gradual decrease has been reported to be 194

mainly due to the accumulation of Cr(III) monolayer and subsequent exhaustion of available Fe(II) at 195

the soil mineral surfaces. This assumption was also supported in this study by that the content of 196

Fe(II) in the magnetite concentration of 6.5 g/L was 14 times higher than the stoichiometric amount 197

required for the full Cr(VI) reduction, but 30% of Cr(VI) was still recovered in the aqueous phase. 198

This means that part of Fe(II) measured in the magnetite exists as structural Fe(II), which cannot be 199

easily used for fast kinetic reactions, or that the Cr(III) layer formed hampers additional reactions. 200

These processes can cause a significant deviation from the well-fitted pseudo first-order kinetics for 201

the reduction of Cr(VI) in the soil mineral suspensions. 202

The effect of Cr(VI) concentration on the reaction kinetics is shown in Fig. 4. Cr(VI) at C0

Cr(VI) 203

= 20 mg/L was completely reduced in 10 minutes and well fitted by the first-order rate law with 204

respect to Cr(VI) concentration (R2 = 0.94). As the initial concentration of Cr(VI) increased, 205

different reduction patterns were observed. Although a similar rapid reduction of Cr(VI) at early 206

sampling times was found at the initial Cr(VI) concentration of 50 and 80 mg/L, the reduction 207

kinetics of Cr(VI) under these higher concentrations was not well fitted by the first-order kinetics but 208

properly fitted by the modified Langmuir-Hinshelwood kinetic model. This implies that the reaction 209

kinetics for the reduction of Cr(VI) is not simply dependent on the Cr(VI) concentration but 210

dependent on both concentrations of Cr(VI) and reactive surfaces on magnetite. 211

The pH of magnetite suspensions significantly affected the reduction kinetics of Cr(VI) as 212

shown in Fig. 5. The reduction rates of Cr(VI) increased with the decrease of pH (from 8 to 5.5), 213

which is similar to the experimental result observed in the reduction of aqueous transition metal 214

species on the surfaces of Fe(II)-bearing oxides [3]. The Cr(VI) injected was fully removed under pH 215

6.5, but 30 and 52% of Cr(VI) were recovered at pH 7.2 and 8.0, respectively. The initial rapid drop 216

9

of Cr(VI) concentration was mainly caused by fast Cr(VI) adsorption to magnetite, followed by the 217

slow decrease of Cr(VI) mainly due to Cr(VI) reduction [16]. The rapid adsorption is mainly related 218

with the zero point charge (zpc) of magnetite surface. The zpc of magnetite has been reported in the 219

pH range of 6.4 to 6.85 [17]. If the pH of suspensions is lower or higher than the zpc of magnetite, 220

the magnetite surfaces can be charged positively or negatively. Cr(VI) (Cr2O72-

) may experience 221

different electrostatic forces from the magnetite surfaces depending on the pH, i.e. attraction under 222

the zpc and repulsion above the zpc. Cr(VI) adsorption, the first step for the reduction of Cr(VI) at 223

the surfaces, at pH 7.2 and 8.0 could be interrupted by electrostatic repulsion. Thus, pH is an 224

important factor that controls the reaction kinetics of Cr(VI). In addition, the reduction of Cr(VI) can 225

consume the H+ from solution based on the following equation [11]. Therefore, the reduction of 226

Cr(VI) could be favorable at the lower pH condition and was comparable with the results of this 227

study . 228

O2H3Fe2Cr(OH)10H3Fe2CrO 22

30

4 (2) 229

230

3.3. Reduction of Cr(VI) in the Suspensions of Soil minerals and Soil with Fe(II) 231

Figure 6 shows the reduction kinetics of Cr(VI) in nontronite, soil, and magnetite with Fe(II) 232

(Fe(II)/Cr(VI) = 1.0). The reduction kinetics of Cr(VI) in the aqueous Fe(II) solution and in the 233

magnetite suspension are also shown as controls for comparison. The relative concentrations of 234

Cr(VI) dropped at first sampling time and slowly approached 0.7 during the reaction time. The 235

reduction patterns of Cr(VI) in nontronite and soil with Fe(II) are very similar, however more Cr(VI) 236

was removed by soil than by nontronite. This may be due to the high content of organic matter 237

(0.76%) in soil sample whose influence on the reduction kinetics has been reported [18]. In contrast 238

to the reductive dechlorination of chlorinated organic compounds, more Cr(VI) was removed in the 239

aqueous Fe(II) solution than in the soil and nontronite suspensions with Fe(II) addition. Chlorinated 240

organic compounds have been known to be reductively degraded at the reactive surface sites, but no 241

significant degradation has been reported in homogeneous reductant solutions such as Fe(II) solution. 242

10

This result indicates that the reactive Fe(II) adsorbed on the soil and nontronite surfaces in 243

heterogeneous systems does not effectively reduce Cr(VI) as much as aqueous Fe(II) does in the 244

homogeneous system. 245

The reactivity of magnetite suspension was enhanced by the addition of Fe(II). More Cr(VI) 246

was reduced in the magnetite suspension with Fe(II) (C/C0 at 60 min = 0.2) than in homogeneous 247

Fe(II) solution (0.6) and in the magnetite suspension without Fe(II) (0.4), respectively. This indicates 248

that the reductive capacity of magnetite was enhanced by forming the reactive surface sites (i.e., 249

surface complexes/precipitates) and/or by regenerating the oxidized surface sites by the addition of 250

Fe(II). However, the amount of Cr(VI) removed in magnetite suspension with Fe(II) is far less than 251

that estimated under an assumption that the removal of Cr(VI) in the suspension with Fe(II) is the 252

summation of each removal from aqueous Fe(II) solution and magnetite suspension under the same 253

experimental conditions. Figure 6 shows the estimated Cr(VI) removal completed in 30 min. The 254

part of adsorbed Fe(II) may be oxidized to Fe(III) at the surfaces forming non-reactive complexes 255

and/or precipitates. No measurement has been conducted to identify these non-reactive chemical 256

species on the surfaces in this study. 257

Figure 7 shows the effect of pH on the reduction of Cr(VI) in the magnetite suspension with 258

Fe(II) in the range of suspension pH 5.5 to 8.0. A similar effect shown in magnetite suspension (Fig. 259

5) was also observed in the magnetite suspensions with Fe(II). The reduction rates of Cr(VI) 260

increased with the decrease of suspension pH in both suspensions with and without Fe(II). The 261

magnetite suspension with Fe(II) at pH 5.5 has the greatest reduction rate constant showing the full 262

reduction of Cr(VI) in one hour, followed by those at pH 6.5 > pH 7.2 > pH 8.0 > pH 7.2 without 263

Fe(II). Because of the zpc of magnetite, the surfaces of magnetite at pH 5.5, pH 6.5, and > pH 7.2 are 264

charged with negative (-O-), neutral (-OH), and positive (-OH2

+) surface species, respectively. This 265

significantly affects the behavior of Cr(VI) and Fe(II) in the magnetite suspension. Gregory et al. 266

have reported that Fe(II) adsorbed on magnetite rapidly increases as the suspension pH increases 267

from 6 to 8 [6]. While 30% of initial Fe(II) was adsorbed on the surfaces of magnetite at pH 6, 90% 268

11

of the initial Fe(II) was adsorbed at pH 6.5 in this study. Fe(II) was not detected in the aqueous 269

solution at pH 7.2 and 8. The higher reactivity of magnetite suspension with Fe(II) observed at lower 270

pH (pH 5.5 and 6.5) is due to the significant reduction of Cr(VI) by Fe(II) in aqueous solution and 271

Fe(II) adsorbed on magnetite surfaces, suggesting that the Cr(VI) reduction by aqueous Fe(II) and 272

adsorbed Fe(II) is more reactive than that by Fe(II) adsorbed on magnetite surfaces alone. 273

274

4. Conclusions 275

The reduction of Cr(VI) in soil mineral suspensions was investigated. The concentration of 276

reactive Fe(II) on the magnetite surfaces was a very important factor that significantly affected the 277

reduction kinetics of Cr(VI). Unless the Cr(VI) concentration is much lower than the surface-bound 278

reactive Fe(II), the slow reduction of Cr(VI) followed the rapid initial sorption of Cr(VI) onto the 279

magnetite surfaces. Because of these characteristics of surface reduction, the reaction kinetics of 280

Cr(VI) cannot be properly described by the zero-order or pseudo-first-order rate law, but well 281

described by the modified Langmuir-Hinshelwood kinetic model. 282

The addition of Fe(II) to the magnetite suspensions enhanced the Cr(VI) reduction, which 283

indicates that the reduction can be controlled and accelerated by the simple treatment. Most Fe(II) 284

added into the magnetite suspension was adsorbed on the soil mineral surface and the Cr(VI) 285

reduction rate increased at constant pH (7.2). However the extent of Cr(VI) reduction in the 286

magnetite suspension with Fe(II) did not reach the estimated value, which was a simple summation 287

of the reduced Cr(VI) in the magnetite suspension and that in the Fe(II) solution. An identical result 288

was observed in the soil and the nontronite suspensions. These results showed that the soluble Fe(II) 289

added cannot stoichiometrically convert the surface Fe(III) to Fe(II) and that electrons may transfer to 290

the bulk suspension. The Fe(II)-initiated reduction of Cr(VI) was also significantly affected by 291

suspension pH. The reduction rate of Cr(VI) decreased as the pH of magnetite suspension increased 292

(above the zpc of magnetite). The results can be applied to the development of remediation 293

12

technology and used for the optimal operation of in-situ redox manipulation and natural attenuation 294

to treat groundwater and soil contaminated with Cr(VI). 295

296

Acknowledgements 297

This research was supported by the grants from the Korean Ministry of Environment (Eco-Technopia 298

21). The contents do not necessarily reflect the views and policies of the sponsor nor does the 299

mention of trade names or commercial products constitute endorsement or recommendation for use. 300

301

References 302

References 303

1. Barnhart, J. (1997). Occurrences, uses, and properties of chromium. Regul. Toxicol. Pharm. 304

26, S3-S7. 305

2. Eary, L.E., and Rai, D. (1987). Kinetics of chromium(III) oxidation to chromium(VI) by 306

reaction with manganese dioxide. Environ. Sci. Technol. 21, 1187-1193. 307

3. White, A.F., and Peterson, M.L. (1996). Reduction of aqueous transition metal species on the 308

surfaces of Fe(II)-containing oxides. Geochim. Cosmochim. Acta 60, 3799-3814. 309

4. Williams, A.G.B., and Scherer, M.M. (2001). Kinetics of Cr(VI) reduction by carbonate 310

green rust. Environ. Sci. Technol 35, 3488-3494. 311

5. Song, D., Kim, Y., and Shin, W.S. (2005). A simple mathematical analysis on the effect of 312

sand in Cr(VI) reduction using zero valent iron, Korean J. Chem. Eng., 22, 67-69. 313

6. Gergory, K.B., Larese-Casanova, P., Parkin, G.F., and Scherer, M.M. (2004). Abiotic 314

transformation of hexahydro-1,3,5-trinitro-1,3,5-triazine by Fe(II) bound to magnetite. 315

Environ. Sci. Technol 38, 1408-1414. 316

7. Williams, A.G.B., and Scherer, M.M. (2004). Spectroscopic evidence for Fe(II)-Fe(III) 317

electron transfer at the iron oxide-water interface. Environ. Sci. Technol 38, 4782-4790. 318

13

8. Eary, L.E., and Rai, D. (1988). Chromate removal from aqueous wastes by reduction with 319

ferrous ion. Environ. Sci. Technol 22, 972-997. 320

9. Taylor, R.M., Maher, B.A., and Self, P.G. (1987). Magnetite in soils: I. the synthesis of 321

singledomain and superparamagnetic magnetite. Clay Miner. 22, 411-422. 322

10. Lee, W., and Batchelor, B. (2002). Abiotic reductive dechlorination of chlorinated ethylenes 323

by iron-bearing soil minerals. 1. pyrite and magnetite. Environ. Sci. Technol. 36, 5147-5154. 324

11. Alowitz, M.J., and Scherer M.M. (2002). Kinetics of nitrate, nitrite, and Cr(VI) reduction by 325

iron metal. Environ. Sci. Technol. 36, 299-306. 326

12. Lee, W., Batchelor, B., and Schlautman, M.A. (2000). Reductive capaicty of soils for 327

chromium. Environ. Technol. 21, 953-963. 328

13. Clescerl, L.S., Greenberg, A.E., and Eaton, A.D. (1992). Standard Methods for the 329

Examination of Water and Wastewater (Boston: American Public Health Association). 330

14. Stookey, L.I. (1970). Ferrozine-A new spectrophotometric regent for iron. Anal. Chem. 42, 331

779-781. 332

15. Peterson, M.L., Gordon, E., Brown, J., Parks, G.A., and Stein, C.L. (1997). Differential redox 333

and sorption of Cr(III/VI) on natural silicate and oxide minerals; EXAFS and XANES results. 334

Geochim. Cosmochim. Acta 61, 3399-3412. 335

16 Peterson, M.L., White, A.F., Brown, G.E., and Parks, G.A. (1997). Surface passivation of 336

magnetite by reaction with aqueous Cr(VI): XAFS and TEM Results. Environ. Sci. Technol. 337

31, 1573-1576. 338

17. Regazzoni, A.E., Blesa, M.A., and Maroto, A.J.G. (1982). Interfacial properties of zirconium 339

dioxide and magnetite in water. J. Colloid Interf. Sci. 91, 560-570. 340

18. Jardine, P.M., Fendorf, S.E., Mayes, M.A., Larsen, I.L., Brooks, S.C., and Bailey, W.B. 341

(1999). Fate and transport of hexavalent chromium in undisturbed heterogeneous soil. 342

Environ. Sci. Technol. 33, 2939-2944. 343

344

14

Figure Captions. 345

Fig 1. X-ray diffraction of synthesized magnetite. The values represent the d-space values (Å) 346

Fig 2. Reduction of Cr(VI) by magnetite, nontronite, and soil at pH 7.2. 347

Fig 3. Effect of magnetite dosage on the reaction kinetics of Cr(VI) at pH 7.2. 348

Fig 4. Effect of initial Cr(VI) concentration on the reaction kinetics of Cr(VI) at pH 7.2. 349

Fig 5. Effect of pH on the reaction kinetics of Cr(VI). 350

Fig 6. Reduction of Cr(VI) in solid suspensions with and without Fe(II) and aqueous Fe(II) solution 351

(pH 7.2, Fe(II)/Cr(VI) = 1.0). 352

Fig 7. Effect of pH on the reaction kinetics of Cr(VI) by magnetite with Fe(II). 353

354

355

15

356

Fig 1. 357

358

359

16

360

Time (min)

0 10 20 30 40 50 60

Re

lative

co

nce

ntr

aito

n o

f C

r(V

I) (

C/C

o)

0.0

0.2

0.4

0.6

0.8

1.0

magnetite

nontronite

soil

control

361

362

363

364

365

Fig 2. 366

367

17

368

369

370

371

Time (hrs)

0 2 4 6 8

Re

lative

co

nce

ntr

atio

n o

f C

r(V

I) (

C/C

o)

0.0

0.2

0.4

0.6

0.8

1.05 g/L

6.5 g/L

8 g/L

372

373

374

375

Fig 3. 376

377

18

378

379

380

381

Time (min)

0 10 20 30 40 50 60

Re

lative

co

nce

ntr

aito

n o

f C

r(V

I) (

C/C

o)

0.0

0.2

0.4

0.6

0.8

1.0

20mg/L

50mg/L

80mg/L

382

383

384

Fig 4. 385

386

19

387

388

389

Time (min)

0 10 20 30 40 50 60

Re

lative

co

nce

ntr

atio

n o

f C

r(V

I) (

C/C

o)

0.0

0.2

0.4

0.6

0.8

1.0

pH 8.0

pH 7.2

pH 6.5

pH 5.5

390

391

392

Fig 5. 393

394

20

395

396

397

Time (min)

0 10 20 30 40 50 60

Rela

tive c

oncentr

ation o

f C

r(V

I) (

C/C

0)

0.0

0.2

0.4

0.6

0.8

1.0

Fe(II)+Nontronite

Fe(II)+Soil

Fe(II)+Magnetite

Only Fe(II) addition

Only Magnetite

Expected value for Fe(II)+Magnetite

398

399

400

Fig 6. 401

402

403

404

405

406

21

407

408

409

Time (min)

0 10 20 30 40 50 60

Re

lative

co

nce

ntr

atio

n o

f C

r(V

I) (

C/C

o)

0.0

0.2

0.4

0.6

0.8

1.0

magnetite only at pH 7.2

pH 8.0

pH 7.2

pH 6.5

pH 5.5

410

411

412

Fig 7. 413

414

415