hydroxides on sorption and speciation

hydroxides on sorption and speciation

Accepted Manuscript Np(V) uptake by bentonite clay: Effect of accessory Fe oxides/hydroxides on sorption and speciation Parveen K. Verma, Anna Yu Roma...

5MB Sizes 0 Downloads 19 Views

Accepted Manuscript Np(V) uptake by bentonite clay: Effect of accessory Fe oxides/hydroxides on sorption and speciation Parveen K. Verma, Anna Yu Romanchuk, Irina E. Vlasova, Victoria V. Krupskaya, Sergey V. Zakusin, Alexey V. Sobolev, Alexander V. Egorov, Prasanta K. Mohapatra, Stepan N. Kalmykov PII:

S0883-2927(16)30273-6

DOI:

10.1016/j.apgeochem.2016.12.009

Reference:

AG 3774

To appear in:

Applied Geochemistry

Received Date: 31 August 2016 Revised Date:

7 December 2016

Accepted Date: 14 December 2016

Please cite this article as: Verma, P.K., Romanchuk, A.Y., Vlasova, I.E., Krupskaya, V.V., Zakusin, S.V., Sobolev, A.V., Egorov, A.V., Mohapatra, P.K., Kalmykov, S.N., Np(V) uptake by bentonite clay: Effect of accessory Fe oxides/hydroxides on sorption and speciation, Applied Geochemistry (2017), doi: 10.1016/ j.apgeochem.2016.12.009. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

Np(V) uptake by bentonite clay: effect of accessory

2

Fe oxides/hydroxides on sorption and speciation

3

Parveen K. Verma1, Anna Yu. Romanchuk2*, Irina E. Vlasova2, Victoria V. Krupskaya2,3,

4

Sergey V. Zakusin2,3, Alexey V. Sobolev2, Alexander V. Egorov2, Prasanta K. Mohapatra1,

5

Stepan N. Kalmykov2

SC

2

7 8

Bhabha Atomic Research Centre, Mumbai – 400 085, Mumbai, India

3

M AN U

1

6

RI PT

1

Lomonosov Moscow State University, Moscow, Russia

Institute of Geology of Ore Deposits, Petrography, Mineralogy and Geochemistry, Russian Academy of Science (IGEM RAS), Moscow, Russia

10

KEYWORDS: Neptunium, interfacial behavior, sorption, speciation, bentonite clays, iron

11

oxides, goethite nanoparticles, accessory minerals, thermodynamic modeling, surface

12

complexation modeling

13

ABSTRACT

14

Batch sorption experiments on thoroughly characterized bentonites and thermodynamic

15

modeling studies were conducted to reveal the role of iron-containing accessory phases on the

16

interfacial behavior of Np(V). Bentonite clays from different industrial deposits with varying

17

total iron contents were selected for the studies. The samples were characterized by XRD,

18

Mossbauer spectroscopy, XRF, HRTEM, SEM-EDX and other techniques, and wherever

19

possible, the accessory iron phases were identified and quantified. Thermodynamic modeling

20

using available surface complexation data revealed the dominant role of the goethite accessory

21

phase, which was present as nanoparticles, in Np(V) sorption at trace level concentrations (10-14

AC C

EP

TE D

9

1

ACCEPTED MANUSCRIPT

M). This fact is independently supported by the combination of SEM-EDX and α-track

23

radiography. These studies illustrate the important role that accessory minerals can play in

24

radionuclides sorption data. This is important to the understanding and modeling of the

25

molecular-level speciation of radionuclides.

26

INTRODUCTION

27

The wide distribution of radionuclides in the subsurface environment from nuclear weapons

28

testing and/or from accidental events, such as Fukushima (2011) and Chernobyl (1986) is of

29

concern for humanity currently and in the future (Buesseler et al., 2011; Cuddihy et al.,

30

1989;Kashkarov et al., 2001;Steinhauser et al., 2015). Neptunium (237Np) is a transuranic

31

element of particular concern because of its long half-life (2.1·106 years) and high geochemical

32

mobility under oxic conditions.

33

time in deep geological repositories (Nash, 2006). A large volume of R&D work is ongoing to

34

understand the sorption, migration and speciation behavior of different radionuclides of several

35

contaminated sites, such as the Hanford Site (Cantrell and Felmy, 2012;Felmy et al., 2010),

36

Nevada Test Site (Kersting et al., 1999;Kersting and Zavarin, 2011;Smith et al., 2003), Savannah

37

River Site (Xu et al., 2014) and Mayak Site (Batuk et al., 2015;Novikov et al., 2006). The

38

understanding of the long-term behavior of the different radionuclides under various chemical

39

and physical conditions helps to better assess their migration and associated contamination or

40

radiation doses that the population might receive from the exposure to these radionuclides.

41

The long-term safety of geological repository sites for radioactive waste disposal is not only

42

related to the capacity to confine large volumes of radiotoxic waste but also to how well the

43

stored high-level radioactive waste can be isolated from the biosphere for thousands of years. For

44

this reason, deep geological repositories are conceptualized and engineered with buffer/backfill

45

materials to block radioactivity release from anthropogenic or environmental mishaps in the

46

future. Bentonite, an aluminosilicate, was found to be a promising candidate for backfill

Np will be a significant radiotoxicity contributor for a long

AC C

EP

TE D

237

M AN U

SC

RI PT

22

2

ACCEPTED MANUSCRIPT

material in engineered barrier systems due to its favorable chemical and physical properties

48

under deep geological conditions, with efficient radionuclide retention capacity (Akgun,

49

2010; Holopainen, 1985; NEA, 2003;).

50

Sorption reactions play an important role in the immobilization/retention of radionuclides (Davis

51

et al., 2005;Geckeis and Rabung, 2008;Goldberg et al., 2007). The majority of the literature on

52

SCM describes the adsorption of metal ions by different pure or synthetic mineral phases under

53

well-controlled laboratory conditions. The complexity of natural soil systems makes it very

54

difficult to apply the concept of SCM to soil and/or natural sediments. To simulate adsorption

55

onto well-characterized natural samples, two basic modeling approaches were proposed: (a) the

56

component additivity (CA) approach and (b) the generalized composite (GC) approach (Davis et

57

al., 1998; Davis et al., 2005). The component additivity approach is based on summing the

58

sorption of individual components in the complex mixture. To simplify the model, it is important

59

to determine the mineral components that dominate sorption (Payne et al., 2013). For example,

60

Payne et al. (2004) found that the presence of trace amounts of anatase determine sorption of

61

U(VI) onto Georgia kaolinite. Kalmykov et al. (2015) demonstrated the dominating effect of

62

Fe/Cr corrosion products (oxides) on plutonium (Pu) partitioning in sedimentary rocks.

63

However, the identification of mineral components that dominate sorption in natural sediments is

64

a challenging task. Even if quantitative and/or qualitative data analysis was performed, further

65

understanding of the combined behavior toward radionuclide sorption under the given physical

66

or chemical conditions is complicated. Furthermore, the components can be distributed quite

67

unevenly. For example, iron oxide/hydroxide minerals may coat clay minerals, (Stubbs et al.,

68

2009) which can influence the sorption process. Prieve et al. (1978) showed that the nature of the

69

electrical double layer was affected by the presence of other heterogeneous particle surfaces.

70

Np(V) sorption has been previously studied on pure or pretreated clay minerals (Benedicto et al.,

71

2014; Bertetti et al., 1998; Bradbury and Baeyens, 1997; Fröhlich, 2015; Kozai, et al., 1996; Li

AC C

EP

TE D

M AN U

SC

RI PT

47

3

ACCEPTED MANUSCRIPT

et al., 2015; Marsac, et al., 2015; Turner et al., 1998; Zavari et al., 2012). Observations of its

73

sorption on natural raw bentonite clays (without treatment) are still lacking (Bradbury and

74

Baeyens, 2011; Fernandes et al., 2015). The present paper is intended to provide insight in that

75

direction. Three samples of raw bentonite clay collected from industrial deposits with varying

76

mineralogical compositions, i.e., different accessory phase compositions and iron speciation, are

77

used for sorption experiments with trace level Np(V). The experimental data are interpreted in

78

terms of the dominating component additivity approach.

79

MATERIALS AND METHODS

80

Characterization of bentonite

81

Bentonite samples from the Khakassia (Russia), Rajasthan and Kutch (India) industrial deposits

82

were used for this study. The Rajasthan and Kutch bentonites were crushed in a ball mill, sieved

83

to size fractions <75 µm and were used without chemical treatment. Khakassia bentonite was

84

used in the Na-form after treatment with Na2CO3.

85

Chemical analysis was performed using XRF in accordance with the standard procedure using an

86

AxiosmAX spectrometer (PANalytical, The Netherlands). Samples were dried at 110°C and

87

were prepared by fusion with lithium borate at 1200°C. The iron content in the samples was

88

determined in the form of total Fe2O3, regardless of the actual valence state.

89

XRD was used to investigate the mineral composition of clay samples, as well as colloidal clay

90

fractions (<0.5 µm), that was selected from an water suspension according to Stokes' law,

91

without using of chemical dispersants. X-ray diffraction patterns were obtained with an Ultima-

92

IV X-ray diffractometer (Rigaku) acquired in a frame of the Moscow State University

93

Development Program. The measurement conditions were as follows: Cu-Kαradiation, D/Tex-

94

Ultra 1D-detector, scan range: 3.6-65° 2θ, scan speed: 5°/min, step: 0.02° 2θ, max intensity: ~

95

25000 counts. Identification of the mineral composition was performed in the textural (oriented)

96

samples prepared from a bentonite suspension in the air-dried and ethylene glycol solvated states

AC C

EP

TE D

M AN U

SC

RI PT

72

4

ACCEPTED MANUSCRIPT

97

(Moore and Reynolds, 1997). The quantitative mineral composition was calculated by the

98

Rietveld method (Bish and Post, 1993) using the BGMN program.

99

Mössbauer spectra at the

57

Fe nuclei were obtained by a MS-1104Em-type spectrometer

operating in constant acceleration mode. Spectra were recorded using a JANIS closed-cycle

101

cryostat and were processed using the SpectrRelax program (Matsnev and Rusakov, 2012). All

102

isomer shifts are given relative to the Mössbauer spectra of α-Fe at room temperature.

103

The specific surface areas of the samples were determined using the BET technique (ASAP

104

2010N, Micrometrics). The HRTEM images were obtained with an aberration-corrected JEOL

105

2100F operated at 200 kV.

106

The original Kutch bentonite was also treated with 1 M NaClO4 for approximately 20 h to

107

convert it to the Na form before centrifuging. The procedure was repeated thrice, and the

108

resultant solid was washed thoroughly with Millipore water to obtain the modified bentonite,

109

which is further referred to as Na-form Kutch (see “Sorption of Np(V) onto clay samples under

110

the steady state” section for details).

111

Sorption experiments

112

The 239Np (4·10-14 M) used for the sorption experiments was collected from the parent 243Am; the

113

extraction and stripping details can be found elsewhere (Sill, 1966). Its purity was ascertained by

114

radiometry. The suspensions of bentonite (0.5 g/L) were prepared in 0.01 M or 1 M NaClO4 for

115

all sorption experiments, which were performed under ambient atmospheric conditions at 25 ±

116

3°C with continuous shaking. The suspensions were conditioned for 48 hours in the background

117

electrolyte solution prior to the addition of the

118

suspensions was adjusted by aliquots of dilute solutions of HClO4 (ultra-pure grade) or NaOH

119

(ultra-pure grade). The proton concentration (molal scale) was measured using a pH electrode (In

120

lab, Mettler Toledo) calibrated against standard pH. At high ionic strength (1M), the measured

AC C

EP

TE D

M AN U

SC

RI PT

100

239

Np(V) radiotracer. The pH of the different

5

ACCEPTED MANUSCRIPT

121

pH was corrected by adding the correction factor (Altmaier et al., 2003). The correction factor

122

was determined by measuring a series of pH samples of known concentration of H+ at both low

123

(0.01 M) and high (1M) ionic strength, and the difference in the two measured pH values was

124

taken as the correction factor for further experiments (0.65 pH units). Aliquots of the supernatant

125

solutions were mixed with scintillation cocktail (Ultima gold) for the

126

scintillation counting using a Quantulus-1220, Perkin Elmer) after centrifugation at 30,000 g for

127

20 min (Allegra 64R, Beckman Coulter). The sorption was calculated based on the difference

128

between the radioactivity of the

129

remaining in the supernatant after equilibration (the measured activity was decay-corrected).

130

A blank experiment (without bentonite) was also performed under the same experimental

131

conditions to account for the sorption of 239Np onto the vial used for the sorption experiments.

132

α-Track radiography and SEM-EDX

133

α-Track radiography and SEM-EDX studies were conducted to identify the local distribution of

134

Np on bentonite samples at the micrometer scale. α-Emitter

135

with the bentonite samples at pH ~ 9. The concentration of the solid phase was 0.2 g/L. The solid

136

phase was separated from the solution by one centrifugation (30,000 g for 20 min). As a result of

137

centrifugation, two fractions with different colors were collected and named the “dark” and

138

“light” fractions, depending upon their relative color intensities. Both fractions were placed on

139

glass slides and covered with CR-39 type polycarbonate film (TASTRAK PADC) for α-

140

radiography. The polycarbonate film was removed after 17 h and was etched in 6 M NaOH at

141

75±1°C for 6 h. The α-track images along with the corresponding areas of the solid samples were

142

observed using an Olympus BX-51 optical microscope. The elemental composition of the solid

143

phase was determined by SEM-EDX (JEOL JSM-6380 LA with a JED 2300 analyzer).

144

Surface complexation modeling

RI PT

Np assay (liquid

Np initially added to the suspension and the radioactivity

237

Np (2·10-6 M) was equilibrated

AC C

EP

TE D

M AN U

SC

239

239

6

ACCEPTED MANUSCRIPT

Sorption was modeled using geochemical speciation software PHREEQC (USGS, 2015).

146

Thermodynamic constants for Np aqueous speciation were taken from the NEA thermodynamic

147

database (Guillaumont et al., 2003). The all present modeling results are calculated at 0.01 M

148

ionic strength.

149

RESULTS AND DISCUSSION

150

Characterization of bentonite clay from different origins

151

Samples from three bentonite industrial deposits were characterized with respect to their

152

elemental composition, mineral phase composition, including accessory minerals, and Fe-

153

speciation. The elemental compositions of the studied samples are presented in Table 1. The high

154

contents of Si and Al are due to the clay minerals present in the tetrahedral and octahedral sheets

155

of smectites (montmorillonite), as well as admixture of kaolinite and chlorite. The higher amount

156

of Si in Khakassia clay compared to the other samples is due to the higher quartz content, which

157

was supported by the XRD results (Figure 1). The presence of relatively high concentration of

158

Ca and Mg can be partly attributed to the presence of carbonate minerals (calcite and dolomite),

159

as indicated by XRD.

160

The Rajastan and Kutch clays had high contents of iron and titanium. Titanium is associated with

161

the presence of anatase, which was identified by XRD. Iron can be included in the octahedral

162

sheets of clay minerals, mostly smectites, and can also be found as separated accessory iron

163

phases.

164

Table 1. XRF data of the chemical compositions of the studied bentonite clays

AC C

EP

TE D

M AN U

SC

RI PT

145

Sample

LOI

Rajastan 20.06 clay Kutch 21.43 clay Khakassia 12.74 clay

Na2O MgO Al2O3

Composition, % SiO2 K2O CaO TiO2 MnO Fe2O3 P2O5

1.15

4.64

16.28 35.12 0.30 6.67

3.09

0.16

12.27

0.11

0.72

2.02

12.33 42.52 0.11 2.77

1.42

0.19

15.58

0.27

2.98

2.48

15.19 58.36 0.97 3.05

0.63

0.09

3.42

0.11 7

ACCEPTED MANUSCRIPT

165

M AN U

SC

RI PT

166

167

Figure 1. X-Ray diffraction patterns collected for the non-oriented natural clays of Rajastan,

169

Kutch and Khakassia bentonite clays. M – montmorillonite, K – kaolinite, Q – quartz, C –

170

calcite, D – dolomite, At – anatase, G – gypsum, Gt – goethite, H – hematite, F – feldspars.

TE D

168

The X-Ray diffraction patterns for all the studied samples have a (001) reflection (Figure

172

1) with d=12.8 Å for Rajastan, reported mainly as a mixture of Na+ and Ca,Mg2+ forms, 15.2 Å

173

for Kutch, interpreted as a mixture of Na+ and Ca,Mg2+ forms, and d=12.6 Å for Khakassia clay,

174

due to the predominant of Na+ form of smectite. The non-basal reflection has the same value for

175

all studied clays: d=4.47 Å, 2.56 Å, 1.69 Å for Rajastan, Kutch and Khakassia, respectively, as

176

well as some other small non-basal reflections. The values of the (060) reflection of the Kutch

177

and Rajastan clays are higher than that of the Khakassia clay, which indicates the presence of Fe

178

in the 2:1 layer. The strongest reflections of the admixed clay and non-clay minerals are shown

179

in Figure 1.

AC C

EP

171

180

The compositions of colloidal fractions <0.5 µm for all clay samples are enriched in Na-

181

montmorillonite (Fig. S1), with basal reflections at d(001) = 12.5 Å, d(002) = 6.11-6.23 Å, and d(004) 8

ACCEPTED MANUSCRIPT

= 3.12-3.13 Å, d(006) = 2.07 Å. The admixture of kaolinite in Rajastan clay was identified by a

183

series of basal reflections at d(001) = 7.16 Å, d(002) = 3.57 Å, d(003) = 2.38 Å, and d(004) = 1.79 Å.

184

The presence of goethite in the Rajastan and Kutch clay was identified by the small reflection at

185

d = 4.19 Å, and the admixture of gypsum in the Kutch sample was identified based on the small

186

reflection at d = 7.62 Å.

187

In short, all the studied clays are enriched with smectite (montmorillonite), with some admixture

188

of other clay minerals (chlorite, kaolinite), as well as carbonates, quartz, different iron oxides

189

and hydroxides and other minerals in relatively small amounts. Due to the different position of

190

the (001) reflection, the Na+ and Ca2+ forms were identified and estimated.

191

Mössbauer spectroscopy was used for speciation of Fe in the clay samples. The Mössbauer

192

spectra of the

193

clay samples present a superposition of a broadened quadrupole doublet and broadened magnetic

194

component, which can be attributed to a magnetically ordered phase. The minor components of

195

the Khakassia clay sample are presented in the Figure 2. The best fit hyperfine parameters of the

196

de-convoluted spectra at different temperatures are shown in Table S1. From these studies, one

197

may conclude that all the bentonite clay samples consist of ferric atoms in the high-spin state

198

within an octahedral oxygen environment, in accordance with the isomer shift values (Menil,

199

1985), which can be attributed to iron atoms within nonmagnetic minerals. In addition, the

200

Khakassia clay sample also consists of two nonmagnetic quadrupole doublets with large isomer

201

shifts and quadrupole splitting (see Table S1), indicating the presence of Fe2+ sites. The

202

hyperfine parameters are similar to those for Si and Al minerals (Murad, 1988) but cannot be

203

attributed to a specific source.

204

The hyperfine parameter analysis shows non-significant changes in the sub-spectra areas versus

205

temperature. One may conclude that the ferric atoms within all clay samples are separated into

206

two essentially different states. The first can be presented as a distribution of quadrupole

Fe nuclei (Figure 2, S2) measured at different temperatures for three bentonite

AC C

EP

TE D

57

M AN U

SC

RI PT

182

9

ACCEPTED MANUSCRIPT

doublets, but it is impossible to fit it correctly with several quadrupole doublets of natural line

208

width (Figure 2). The average isomer shift of this distribution is illustrative of high-spin Fe3+

209

atoms within an octahedral oxygen environment. The wide distribution of quadrupole splitting

210

values (Fig. S3A, C, E) is related to different forms of ferric hydroxides, surface atoms, and solid

211

solutions of nonmagnetic Al and Si phases (Murad, 1988). It can be observed that the profiles of

212

the quadrupole splitting distribution are similar for all samples and do not change with

213

temperature.

AC C

EP

TE D

M AN U

SC

RI PT

207

10

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 2. Mössbauer spectra measured at 77 K together for the (A) – Rajastan, (B) – Kutch and (C) – Khakassia clays.

11

ACCEPTED MANUSCRIPT

The second is a broadened Zeeman component that is present in all samples, which could appear

215

due to the size effects of nanocrystalline particles (Morup and Knudsen, 1985). This hypothesis

216

is logical because the magnetic component becomes more “classic” when the temperature

217

decreases (Figure 2, S2). These components were fitted (Figure S3B, D, F) as a distribution of

218

Zeeman sextets with different frequencies of slow superparamagnetic relaxation (Morup and

219

Knudsen, 1985), which is a manifestation of the particle size effect. The hyperfine parameters of

220

these distributions (Table S1) correspond to major goethite nanocrystalline phase, for which the

221

absolute value of the quadrupole shift ε is slightly higher than for the α-Fe2O3 hematite phase

222

(Vandenberghe et al., 1986). Thus, the Kutch and Rajastan samples contain only a small amount

223

(~3-6%) of “bulk” α-Fe2O3. The weak changes in the quadrupole shift average values (Table S1)

224

are related to the Morin transition (TM ≈ 260 K) of the hematite phase (Nininger and Schroeer,

225

1978).

226

The presence of iron oxide/hydroxide nanoparticles was confirmed by HRTEM. Figure 3 shows

227

typical images of goethite nanoparticles in the Kutch clay samples. The particle size varies in the

228

range from 2 to 10 nm.

AC C

EP

TE D

M AN U

SC

RI PT

214

Figure 3. (A), (B) HRTEM images of goethite nanoparticles from Kutch clay samples. Selected area electron diffraction are presented on the insert, white lines correspond to the position of the goethite reflexes.

12

ACCEPTED MANUSCRIPT

The X-ray diffraction patterns of the Rajastan and Kutch clay samples showed the presence of

230

amorphous phases. Based on the detailed analysis of the Mossbauer spectra, we assume this

231

amorphous phase can be attributed to goethite nanoparticles. Therefore, the concentrations of

232

various accessory iron-containing phases were calculated using the Mossbauer and XRF data.

233

The amounts of other mineral phases were based on the XRD data. All data for the clay samples

234

studied in the present work are listed in Table 2.

Table 2. Composition and related properties of the studied bentonite samples

Na+-smectite 2+

23.6 66.4

3.5

1.6

EP

236

Khakassia clay

32.8

65

14.4

4.5 1.3 2.2 15.9 6.3 3.8 0.5 0.5 -

26.4 1.9

3.2 11.9 1.3 1.1 2.9 3.0 4.5 0.7 0.7 2 Surface area, m /g 115 50

TE D

Ca,Mg -smectite Illite Kaolinite Chlorite Quartz Albite Calcite Dolomite Gypsum Anatase Goethite Hematite

Kutch clay Rajastan clay Composition, mass-%

SC

Sample

M AN U

235

RI PT

229

15

Kinetics of Np(V) sorption onto bentonite

238

The Np(V) uptake studies were performed as described above. The kinetics of Np(V) sorption

239

onto bentonite of different origins was measured from 0.01 M and 1 M NaClO4 at pH ~8.5

240

(Figure 4). Various theoretical kinetic models exist in the literature for metal ion absorption at

241

mineral-water interfaces (Azizian, 2004;Ho and McKay, 1999;Ho, 2006;Plazinski et al., 2009).

242

For the present system, different kinetic models (i.e., pseudo-first-order and pseudo-second-

243

order) along with the diffusion model were tested for Np(V) sorption onto different bentonite

AC C

237

13

ACCEPTED MANUSCRIPT

samples at different ionic strengths. The different kinetic equations and their linear forms used to

245

model the sorption data are presented in Table S2.

246

The best fit was obtained by the pseudo-second-order based on the linear fitting statistical

247

analysis (R2 value) for all the studied systems. The derived kinetics parameters for different

248

bentonite clay samples are listed in Table S3. The steady-state conditions for Np(V) sorption

249

were reached during the first 24 hours for all bentonite samples at different ionic strengths

250

(Figure 4). The sorption of Np(V) onto Khakassia clay was faster but reached a lower value.

251

Although there were mineralogical differences in the bentonite from Rajasthan and Kutch, the

252

sorption and kinetics of Np(V) on these samples were similar. The difference in the rate of

253

Np(V) sorption onto Khakassia bentonite is attributed to differences in their mineralogy and/or

254

mechanism of sorption and will be discussed below. 100

B 100 80

60 40

TE D

Sorption, %

80

Kutch clay Rajasthan clay Khakassia clay

20 0 10

20

30

EP

0

40

Sorption, %

A

M AN U

SC

RI PT

244

60 40

Kutch clay Rajasthan clay Khakassia clay

20 0

50

0

Time, h.

10

20

30

40

50

Time, h.

AC C

Figure 4. Kinetics of Np(V) sorption onto bentonite of different origins at (A) 0.01 M and (B) 1 M ionic strength (pHinitial = 8.5, [solid phase] = 0.5 g/L, [Np] = 4·10-14 M).

255

Sorption of Np(V) onto clay samples under the steady state

256

The sorption of

257

pH in 0.01 M and 1 M NaClO4 (Figure 5). In all cases, sorption increased with increasing pH at a

258

given ionic strength. The shape and position of the pH sorption edges is different for the studied

259

bentonite samples — Khakassia bentonite has the lowest sorption. The conditions of our

260

experiments correspond to a high excess of the solid phase compared to the Np(V) concentration,

239

Np(V) (4·10-14 M) onto different bentonite samples was studied with varying

14

ACCEPTED MANUSCRIPT

with a constant solid to liquid ratio of 0.5 g/L. However the values of specific surface area of the

262

studied samples are different. Therefore the changes in pH-edges from sample to sample can be

263

effected by the difference in surface area (see Fig.S5) and mineral composition (Table 2).

AC C

EP

TE D

M AN U

SC

RI PT

261

Figure 5. Thermodynamic modeling and experimental data of the pH dependence of Np(V) sorption onto (A) Kutch, (B) Rajastan and (C) Khakassia clay at different ionic strengths. Black 15

ACCEPTED MANUSCRIPT

and red lines – sorption of Np(V) onto Na+-form of montmorillonite using the constants from Bradbury and Baeyens (1997), blue line – sorption of Np(V) onto goethite using the constants from Kohler et al. (1999), magneta line – sorption of Np(V) onto Na+-form of montmorillonite using the constants from Tachi et al. (2014). The values of concentration of each proposed

RI PT

phases in the legend correspond to the values given in Table 2. Since in 1M NaClO4 smectite is completely converted to the Na+-form (see discussion below), two modeling curves for Np(V) sorption onto montmorillonite are presented for Indian clays – for initial sample and after complete conversion to Na+-form.

SC

([Np(V)] = 4·10-14 M, [solid phase] = 0.5 g/L).

M AN U

264

The differences and similarities in radionuclide sorption at different ionic strengths are indicative

266

of their speciation on the surface (Goldberg et al., 2007). In general, sorption remains the same

267

for inner-sphere complexation with increasing ionic strength, whereas it is decreased at higher

268

ionic strengths for outer-sphere complexation. In the present case, the sorption of Np(V) onto

269

Rajastan and Khakassia bentonite clays at 0.01 M and 1 M ionic strengths is the same,

270

suggesting inner-sphere complexation. Previously it was shown that the sorption mechanism of

271

Np(V) onto clay minerals depends on the pH. At low pH, ion-exchange is dominant (Benedicto

272

et al., 2014;Zavarin et al., 2012), whereas at higher pH, surface complexation is the major

273

mechanism of uptake. In our experiments, sorption at pH < 6 was low and barely changed with

274

increasing ionic strength. This effect may be partially due to the hindered exchange of Np(V)

275

with Ca2+ and Mg2+, especially in the case of Kutch and Rajastan clays, as well as the

276

competition with these cations for sorption sites. These cations are found in solution due to the

277

partial dissolution of accessory minerals (calcite, dolomite and others) and the exchange of

278

interlayer cations in the clay minerals. The presence of Ca2+ and Mg2+ in solution may affect the

279

ion exchange of Np(V) (see Table S4).

AC C

EP

TE D

265

16

ACCEPTED MANUSCRIPT

In the case of Kutch clay, the sorption differs at varying ionic strengths, but the trend is

281

unexpected. The experimental results demonstrate higher sorption of Np(V) on Kutch clay at

282

higher ionic strength. Kutch clay mostly contains Mg2+- and Ca2+-substituted smectite (Table 2),

283

which may influence the sorption of Np(V), consistent with previously published data — Np(V)

284

sorption is higher onto Na+-smectite than Ca2+-smectite (Benedicto et al., 2014;Kozai et al.,

285

1996), which may explain the lower Np(V) sorption onto Kutch clay at lower ionic strength. To

286

determine the impact, an additional experimental approach was taken. The original Kutch

287

bentonite was treated with 1 M NaClO4 to convert the Ca/Mg2+-smectite form of Kutch to the

288

Na+ form. The treated Kutch bentonite (Na+-smectite form), now called Na-form Kutch clay, was

289

further characterized by XRD (Figure S4). The XRD diffraction pattern confirm the smectite

290

modification — Na+-Kutch bentonite has two visible basal reflections with d(001)=12.6 Å and

291

d(002)=6.3 Å, which indicate interlayer cation exchange to Na+.

292

The Np(V) sorption kinetics and pH-edge on Kutch and Na-form Kutch clays were compared

293

(Figure 6). The rate of Np(V) sorption onto Na-form Kutch clay was faster than onto the original

294

Kutch samples, whereas the equilibrium sorption value was similar for both samples (Table S3,

295

Figure 6A). When comparing the pH edges of Np(V) sorption (Figure 6B), it is clear that the

296

sorption of Np(V) onto Na-form Kutch is similar at different ionic strengths and similar to the

297

sorption data for the original Kutch clay at higher ionic strength (1 M NaClO4).

AC C

EP

TE D

M AN U

SC

RI PT

280

17

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 6. (A) Kinetics of Np(V) sorption onto Kutch and Na-form Kutch clays; (B) pH

TE D

dependence of Np(V) sorption onto Kutch and Na-form Kutch clays at different ionic strengths. Modeling of Np(V) sorption onto clays of different origin

299

Among the papers devoted to Np(V) sorption onto bentonites accompanied with thermodynamic

300

modeling, the papers by Bradbury and Baeyens (1997) and Tachi et al.(2014) are highlighted. In

301

both manuscripts, surface complexation modeling of Np(V) sorption onto purified

302

montmorillonite was performed, and the findings were fitted using a linear free-energy

303

relationship (LFER). The notable differences in the above papers were that Bradbury and

304

Baeyens studied the sorption of Np(V) at trace concentrations (<10-13 M) and used two-site

305

protolysis non-electrostatic surface complexation (2SPNE SC/CE), whereas Tachi et al.

306

performed their experiment at higher Np(V) concentrations and used one-site protolysis non-

307

electrostatic surface complexation modeling (1SPNE SC/CE).

AC C

EP

298

18

ACCEPTED MANUSCRIPT

These two models and corresponding surface complexation constants (Table S5) were applied to

309

our experimental data using the concentration of smectite in different bentonites determined

310

from the characterization experiments (Figure 5). Because of the absence of the equilibrium

311

constants for Np(V) sorption onto Ca2+/Mg2+-smectite, modeling was done only for Np(V)

312

sorption onto Na+-smectite. Based on the assumption that in 1M NaClO4 all smectite forms are

313

converted to the Na+-form, two modeling curves for Np(V) sorption onto montmorillonite are

314

presented for Indian clays. Namely, in case of Kutch clay black and red lines in fig.5A calculated

315

for 23% of mass fraction of montmorillonite in the sample (correspond to 0.01 M NaClO4) and

316

90% of mass fraction for montmorillonite in the sample (correspond to 1 M NaClO4). Increasing

317

of Np(V) sorption with mass fraction of montmorillonite that was found by modeling correlates

318

with experimental data on Np(V) sorption onto Kutch clay with increasing of ionic strength.

319

However modeling of Np(V) sorption onto pure montmorillonite did not result in sufficient

320

agreement with experimental data in case of Kutch clay. Np(V) sorption Rajasthan clay does not

321

converge well with the models of Np(V) sorption onto montmorillonite as well. Np(V) sorption

322

onto Khakassia bentonite is satisfactorily fitted by the model of Bradbury and Baeyens. Whereas

323

modeling of sorption with equilibrium constants from Tachi et al. does not give adequate

324

modeling in all cases (shown only for Khakassia clay).

325

We proposed that admixture phases present in the studied clay may influence on Np(V) sorption.

326

The most likely candidates are goethite, hematite and anatase. In this work, sorption onto

327

goethite was taken into account because of its higher concentration than other accessory phases

328

in the samples and its higher Np(V) sorption constant (Kohler et al., 1999) than hematite (Muller

329

et al., 2015;Romanchuk and Kalmykov, 2014) and anatase (Gracheva et al.,2014). The sorption

330

of Np(V) onto goethite was modeled using its total mass fraction from the Table 2 and by

331

assuming that the specific surface area of goethite, i.e. the surface area per mass of goethite, is

332

identical with the specific surface area of the sample as a whole, namely 115 m2/g for Kutch

333

clay, 50 m2/g for Rajastan clay, 15 m2/g for Khakassia clay (Figure 5). The surface area of each

AC C

EP

TE D

M AN U

SC

RI PT

308

19

ACCEPTED MANUSCRIPT

mineral phase within the sample is then proportional to its mass fraction. An experimental

335

determination of the surface area of individual phases in a complex mineral mixture is difficult

336

and remains an open task. The sorption of Np(V) onto Kutch and Rajastan bentonites fit well

337

with Np(V) sorption onto goethite (blue line in Figure 4), indicating that the presence of even a

338

relatively small amount of iron hydroxide/oxide in natural clays may have influence on Np(V)

339

uptake.

340

These findings can explain the difference in the kinetics of Np sorption—the rate of sorption of

341

Np(V) onto the Khakassia clay is faster than Rajastan and Kutch clays (Figure 4). In the sample

342

from Khakassia, Np(V) sorption occurred on the major minerals phases (montmorillonite),

343

whereas for both Rajastan and Kutch clays, Np(V) is presumably sorbed onto both

344

montmorillonite and the accessory mineral goethite, which composes approximately 3% of the

345

mass. Therefore, the diffusion process can explain the slower kinetics of sorption for the

346

Rajastan and Kutch clays.

347

Local distribution of Np(V) on the clays

348

The local distribution of Np(V) on the clays was studied at higher surface loading compared to

349

the batch experiments. The samples were divided into “light” and “dark” fractions by

350

centrifugation.

351

Figure 7 shows typical optical microscope and corresponding α-track analysis images for Np(V)

352

sorbed onto the Rajastan clay. The “light” fraction of Np(V) is evenly distributed, whereas the

353

“dark” fraction shows hot-spots where

354

fraction contains mainly Fe- and Fe/Ti-containing oxides and hydroxides (Figure S6), whereas

355

the “light” fraction predominantly contains clay minerals. Thus, even under high surface loading

356

conditions, the presence of accessory minerals, such as iron oxides/hydroxides, in clays can

357

greatly influence Np(V) sorption.

AC C

EP

TE D

M AN U

SC

RI PT

334

237

Np is concentrated. SEM-EDS showed that the “dark”

20

ACCEPTED MANUSCRIPT

CONCLUSIONS

359

Within the scope of this work we study the uptake mechanisms of tracer amount of Np(V) on the

360

actual clays varying in mineralogical composition and accessory mineral content. We showed

361

that Fe oxides/hydroxide, goethite in this case, present even at low concentrations (few percent)

362

play the important role in Np(V) sorption and speciation. This has influence both on the kinetics

363

of sorption, and especially on pH sorption edges at the steady state and possibly on the

364

reversibility of sorption. It is important to conclude that the surface complexation data are highly

365

important to the modeling of such complex heterogenic systems as soils or bottom sediments.

SC

RI PT

358

AC C

EP

TE D

M AN U

366

Figure 7. Optical microscope images and corresponding α-track analysis images for Np(V) sorbed onto Rajastan clay. 367

ACKNOWLEDGEMENT

21

ACCEPTED MANUSCRIPT

The Np(V) sorption studies and thermodynamic modeling were supported by the Russian

369

Scientific Foundation (project 16-13-00049). The XRD mineralogical studies were supported by

370

the Russian Scientific Foundation (project 16-17-10270).

371 372 373 374

REFERENCES Akgun, H., 2010. Geotechnical characterization and performance assessment of bentonite/sand mixtures for underground waste repository sealing. Appl. Clay Sci. 49, 394-399.

375 376 377

Altmaier, M., Metz,V., Neck,V., Muller, R., Fanghanel, Th., 2003. Solid-liquid equilibria of Mg(OH)2(cr) and Mg2(OH)3Cl·4H2O(cr) in the system Mg-Na-H-OH-Cl-H2O at 25°C. Geochim Cosmochim. Acta 67, 3595-3601.

378 379

Azizian, S., 2004. Kinetic models of sorption: a theoretical analysis. J. Colloid Interface Sci. 276, 47-52.

380 381 382 383 384

Batuk, O.N., Conradson, S.D., Aleksandrova, O.N., Boukhalfa, H., Burakov, B.E., Clark, D.L., Czerwinski, K.R., Felmy, A.R., Lezama-Pacheco, J.S., Kalmykov, S.N., Moore, D.A., Myasoedov, B.F., Reed, D.T., Reilly, D.D., Roback, R.C., Vlasova, I.E., Webb, S.M., Wilkerson, M.P., 2015. Multiscale Speciation of U and Pu at Chernobyl, Hanford, Los Alamos, McGuire AFB, Mayak, and Rocky Flats. Environ. Sci. Technol. 49, 6474-6484.

385 386 387

Benedicto, A., Begg, J.D., Zhao, P., Kersting, A.B., Missana, T., Zavarin, M., 2014. Effect of major cation water composition on the ion exchange of Np(V) on montmorillonite: NpO2+ -Na+K+-Ca2+-Mg2+ selectivity coefficients. Appl. Geochem. 47, 177-185.

388 389 390

Bertetti, F.P., Pabalan, R.T., Almendarez, M.G., 1998. Studies of Neptunium Sorption on Quartz, Clinoptilolite, Montmorillonite, and a-Alumina. In: Jenne, E.A. (Ed.), Adsorption of Metals by Geomedia Academic Press: San Diego, pp. 131-148.

391 392

Bish, D.L., Post, J.E., 1993. Quantitative mineralogical analysis using the Rietveld full-pattern fitting method. American Mineralogist 78, 932-940.

393 394

Bradbury, M.H., Baeyens, B., 1997. A mechanistic description of Ni and Zn sorption on Namontmorillonite Part II: modelling. J. Cont. Hydrol. 27, 223-248.

395 396 397

Bradbury, M.H., Baeyens, B., 2011. Predictive sorption modelling of Ni(II), Co(II), Eu(IIII), Th(IV) and U(VI) on MX-80 bentonite and Opalinus Clay: A “bottom-up” approach. Appl. Clay Sci. 52, 27–33.

398 399

Buesseler, K., Aoyama, M., Fukasawa, M., 2011. Impacts of the Fukushima Nuclear Power Plants on Marine Radioactivity. Environ. Sci. Technol. 45, 9931-9935.

400 401

Cantrell, K.J., Felmy, A.R. Plutonium and Americium Geochemistry at Hanford: A Site-Wide Review . PNNL-21651. 2012. Pacific Northwest National Laboratory .

402 403 404

Cuddihy, R.G., Finch, G.L., Newton, G.J., Hahn, F.F., Mewhinney, J.A., Rothenberg, S.J., Powers, D.A., 1989. Characteristics of radioactive particles released from the Chernobyl nuclear reactor. Environ. Sci. Technol. 23, 89-95.

405 406 407

Davis, J.A., Coston, J.A., Kent, D.B., Fuller, C.C., 1998. Application of the Surface Complexation Concept to Complex Mineral Assemblages. Environ. Sci. Technol. 32, 28202828.

AC C

EP

TE D

M AN U

SC

RI PT

368

22

ACCEPTED MANUSCRIPT

Davis, J.A., Ochs, M., Olin, M., Payne, T.E., Tweed, C.J. NEA Sorption Project Phase II. Interpretation and Prediction of Radionuclide Sorption onto Substrates Relevant for Radioactive Waste Disposal Using Thermodynamic Sorption Models. 5992. 2005. OECD.

411 412 413

Felmy, A.R., Cantrell, K.J., Conradson, S.D., 2010. Plutonium contamination issues in Hanford soils and sediments: Discharges from the Z-Plant (PFP) complex. Phys. Chem. Earth, Parts A/B/C 35, 292-297.

414 415

Fernandes, M.M., Ver, N., Baeyens, B., 2015. Predicting the uptake of Cs, Co, Ni, Eu, Th and U on argillaceous rocks using sorption models for illite. Appl. Geochem. 59, 189–199.

416 417

Fröhlich, D.R., 2015. Sorption od neptunium on clays and clay minerals – a review. Clays Clay Mineral. 63 (4), 262-276.

418 419

Geckeis, H., Rabung, T., 2008. Actinide geochemistry: From the molecular level to the real system. J. Contam. Hydrol. 102, 187-195.

420 421

Goldberg, S., Criscenti, L.J., Turner, D.R., Davis, J.A., Cantrell, K.J., 2007. AdsorptionDesorption Processes in Subsurface Reactive Transport Modeling. Vadose Zone J. 6, 407-435.

422 423 424

Gracheva, N.N., Romanchuk, A.Y., Smirnov, E.A., Meledina, M.A., Garshev, A.V., Shirshin, E.A., Fadeev, V.V., Kalmykov, S.N., 2014. Am(III) sorption onto TiO2 samples with different crystallinity and varying pore size distributions. Appl. Geochem. 42, 69-76.

425 426 427 428

Guillaumont, R., Fanghanel, T., Neck, V., Fuger, J., Palmer, D.A., Grenthe, I., Rand, M.H. Update on the chemical thermodynamics of uranium, neptunium, plutonium, americium and technecium. 2003. Amsterdam, The Netherlands, OECD Nuclear Energy Agency. Chemical Thermodynamics.

429 430

Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Proc. Biochem. 34, 451-465.

431 432

Ho, Y.S., 2006. Review of second-order models for adsorption systems. J. Hazardous Materials 136, 681-689.

433 434

Holopainen, P., 1985. Crushed aggregate-bentonite mixtures as backfill material for repositories of low- and intermediate level radioactive wastes. Engineering Geology 21, 239-245.

435 436 437

Kalmykov, S.N., Vlasova, I.E., Romanchuk, A.Yu., Zakharova, E.V., Volkova, A.G., Presnyakov, I.A., 2015. Partitioning and speciation of Pu in the sedimentary rocks aquifer from the deep liquid nuclear waste disposal. Radiochim. Acta 103, 175-185.

438 439 440 441

Kashkarov, L.L., Tcherkezian, V.O., Kalinina, G.V., Ivliev, A.I., Shalaeva, T.V., Korovaikov, P.A., 2001. Chernobyl "Hot Particles": Radionuclide Composition and Contribution in the Total Soil Radioactivity. In: Frontasyeva, M., Perelygin, V., Vater, P. (Eds.), Radionuclides and Heavy Metals in Environment Springer Netherlands, pp. 43-48.

442 443

Kersting, A.B., Efurd, D.W., Finnegan, D.L., Rokop, D.L., Smith, D.K., Thompson, J.L., 1999. Migration of plutonium in groundwater at the Nevada Test Site. Nature 397, 56-59.

444 445 446

Kersting, A.B., Zavarin, M., 2011. Colloid-facilitated transport of plutonium at the Nevada Test Site, NV, USA. In: Kalmykov, S.N., Denecke, M.A. (Eds.), Actinide Nanoparticle Research Springer-Verlag, Berlin.

447 448

Kohler, M., Honeyman, B.D., Leckie, J.O., 1999. Neptunium(V) sorption on hematite (a-Fe2O3) in aqueous suspension: The effect of CO2. Radiochim. Acta 85, 33-48.

449 450

Kozai, N., Ohnuko, T., Matsumoto, J., Banba, T., Ito, Y., 1996. A study of the specific sorption of neptunium(V) on smectite in low pH solution. Radiochim. Acta 75, 149-158.

AC C

EP

TE D

M AN U

SC

RI PT

408 409 410

23

ACCEPTED MANUSCRIPT

Li, P., Liu, Z., Ma, F., Shi, Q., Guo, Z., Wu, W., 2015. Effects of pH, ionic strength and humic acid on the sorption of neptunium(V) to Na-bentonite. J. Mol. Liq. 206, 285-292.

453 454 455

Marsac, R., Banik, N.l., Lützenkirchen, J., Marquardt, C.M., Dardenne, K., Schild, D., Rothe, J., Diascorn, A., Kupcik, T., Schäfer, T., Geckeis, H., 2015. Neptunium redox speciation at the illite surface. Geochim. Cosmochim. Acta 152, 39-51.

456 457

Matsnev, M.E., Rusakov, V.S., 2012. SpectrRelax: An application for Mossbauer spectra modeling and fitting. P Conf. Proc. 1489, 178-185.

458 459 460 461

Menil, F., 1985. Systematic trends of the 57Fe Mossbauer isomer shifts in (FeOn) and (FeFn) polyhedra. Evidence of a new correlation between the isomer shift and the inductive effect of the competing bond T-X (-Fe) (where X is O or F and T any element with a formal positive charge). J. Phys. Chem. Solids 46, 763-789.

462 463

Moore, D.M., Reynolds, R.C., 1997. X-Ray Diffraction and the Identification and Analysis of Clay Minerals. Oxford University Press.

464 465

Morup, S., Knudsen, J.E., 1985. Molecular dynamics and spin-lattice relaxation of Fe3+ in a supercooled liquid. J. Phys. C 18, 2943.

466 467 468

Muller, K., Groschel, A., Rossberg, A., Bok, F., Franzen, C., Brendler, V., Foerstendorf, H., 2015. In situ Spectroscopic Identification of Neptunium(V) Inner-Sphere Complexes on the HematiteGAOWater Interface. Environ. Sci. Technol. 49, 2560-2567.

469 470 471

Murad, E., 1988. Properties and behavior of iron oxides as determined by Mossbauer spectroscopy. In: Stucki, J.W., Goodman, B.A., Schwertmann, U. (Eds.), Iron in Soils and Clay Minerals pp. 309-350.

472 473 474

Nash, K.L., 2006. Actinide Solution Chemistry And Chemical Separations: Structure-Function Relationships In The Grand Scheme Of Actinide Separations Science. In: May, I., Bryan, N.D., Alvares, R. (Eds.), Recent Advances In Actinide Science RSC Publishing, Cambridge, pp. 427.

475 476 477

NEA, 2003. Engineered Barrier Systems and the Safety of Deep Geological Repositories: stateof-the-art . Nuclear Energy Agency Organisation for Economic Co-Operation and Development, OECD .

478 479

Nininger, J., Schroeer, D., 1978. Mossbauer studies of the morin transition in bulk and microcrystalline a-Fe2O3. J. Phys. Chem. Solids 39, 137-144.

480 481 482

Novikov, A.P., Kalmykov, S.N., Utsunomiya, S., Ewing, R.C., Horreard, F., Merkulov, A., Clark, S.B., Tkachev, V.V., Myasoedov, B.F., 2006. Colloid Transport of Plutonium in the FarField of the Mayak Production Association, Russia. Science 314, 638-641.

483 484 485 486

Payne, T.E., Brendler, V., Ochs, M., Baeyens, B., Brown, P.L., Davis, J.A., Ekberg, C., Kulik, D.A., Lutzenkirchen, J., Missana, T., Tachi, Y., Van Loon, L.R., Altmann, S., 2013. Guidelines for thermodynamic sorption modelling in the context of radioactive waste disposal. Environ. Model. Software 42, 143-156.

487 488

Payne, T.E., Davis, J.A., Lumpkin, G.R., Chisari, R., Waite, T.D., 2004. Surface complexation model of uranyl sorption on Georgia kaolinite. Appl. Clay Sci. 26, 151-162.

489 490

Plazinski, W., Rudzinski, W., Plazinska, A., 2009. Theoretical models of sorption kinetics including a surface reaction mechanism: A review. Adv. Colloid Interface Sci. 152, 2-13.

491 492

Prieve, D.C., Ruckenstein, E., 1978. The double-layer interaction between dissimilar ionizable surfaces and its effect on the rate of deposition. J. Colloid Interface Sci. 63, 317-329.

493 494 495

Romanchuk, A.Yu., Kalmykov, S.N., 2014. Actinides sorption onto hematite: experimental data, surface complexation modeling and linear free energy relationship. Radiochim. Acta 102, 303310.

AC C

EP

TE D

M AN U

SC

RI PT

451 452

24

ACCEPTED MANUSCRIPT

Sill, C.W., 1966. Preparation of Neptunium-239 Tracer. Anal. Chem. 38, 802-804.

497 498 499

Smith, D.K., Finnegan, D.L., Bowen, S.M., 2003. An inventory of long-lived radionuclides residual from underground nuclear testing at the Nevada test site, 1951GAO1992. J. Environ. Radioactivity 67, 35-51.

500 501 502 503

Steinhauser, G., Niisoe, T., Harada, K.H., Shozugawa, K., Schneider, S., Synal, H.A., Walther, C., Christl, M., Nanba, K., Ishikawa, H., Koizumi, A., 2015. Post-Accident Sporadic Releases of Airborne Radionuclides from the Fukushima Daiichi Nuclear Power Plant Site. Environ. Sci. Technol. 49, 14028-14035.

504 505 506

Stubbs, J.E., Veblen, L.A., Elbert, D.C., Zachara, J.M., Davis, J.A., Veblen, D.R., 2009. Newly recognized hosts for uranium in the Hanford Site vadose zone. Geochim. Cosmochim. Acta 73, 1563-1576.

507 508 509

Tachi, Y., Ochs, M., Suyama, T., 2014. Integrated sorption and diffusion model for bentonite. Part 1: clay-water interaction and sorption modeling in dispersed systems. J. Nucl. Sci. Technol. 51, 1177-1190.

510 511

Turner, D.R., Pabalan, R.T., Bertetti, F.P., 1998. Neptunium(V) sorption on montmorillonite; an experimental and surface complexation modeling study. Clays and Clay Minerals 46, 256-269.

512

USGS, 2015. PHREEQC interactive. [3.2.0.9820].

513 514 515

Vandenberghe, R.E., De Grave, E., De Geyter, G., Landuydt, C., 1986. Characterization of goethite and hematite in a Tunesian soil profile by Moessbauer spectroscopy. Clays and Clay Minerals 34, 275-280.

516 517 518 519

Xu, C., Athon, M., Ho, Y.F., Chang, H.S., Zhang, S., Kaplan, D.I., Schwehr, K.A., DiDonato, N., Hatcher, P.G., Santschi, P.H., 2014. Plutonium Immobilization and Remobilization by Soil Mineral and Organic Matter in the Far-Field of the Savannah River Site, U.S. Environ. Sci. Technol. 48, 3186-3195.

520 521 522

Zavarin, M., Powell, B.A., Bourbin, M., Zhao, P., Kersting, A.B., 2012. Np(V) and Pu(V) Ion Exchange and Surface-Mediated Reduction Mechanisms on Montmorillonite. Environ. Sci. Technol. 46, 2692-2698.

AC C

EP

TE D

M AN U

SC

RI PT

496

25

ACCEPTED MANUSCRIPT Iron oxyhydroxide dominates the Np(V) sorption at trace concentration onto raw clay



Accessory minerals in clay effect on pH edge and kinetics of Np(V) sorption



SCM was applied to describe sorption of Np(V) onto raw clay samples

AC C

EP

TE D

M AN U

SC

RI PT