You are on page 1of 36

Accepted Manuscript

Complexation of europium and uranium with natural organic matter (NOM) in highly
saline water matrices analysed by ultrafiltration and inductively coupled plasma mass
spectrometry (ICP-MS)

Ramona Hahn, Christina Hein, Jonas M. Sander, Ralf Kautenburger

PII: S0883-2927(17)30023-9
DOI: 10.1016/j.apgeochem.2017.01.008
Reference: AG 3800

To appear in: Applied Geochemistry

Received Date: 20 June 2016


Revised Date: 9 January 2017
Accepted Date: 10 January 2017

Please cite this article as: Hahn, R., Hein, C., Sander, J.M., Kautenburger, R., Complexation of
europium and uranium with natural organic matter (NOM) in highly saline water matrices analysed
by ultrafiltration and inductively coupled plasma mass spectrometry (ICP-MS), Applied Geochemistry
(2017), doi: 10.1016/j.apgeochem.2017.01.008.

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.
ACCEPTED MANUSCRIPT
1 Complexation of europium and uranium with natural organic matter (NOM)

2 in highly saline water matrices analysed by ultrafiltration and inductively

3 coupled plasma mass spectrometry (ICP-MS)

5 Ramona Hahn, Christina Hein, Jonas M. Sander, Ralf Kautenburger*

PT
6

7 WASTe group, Institute of Inorganic Solid State Chemistry, Saarland University, P.O.

RI
8 Box 151150, D-66041 Saarbrcken, Germany

SC
9

10

U
11
AN
12

13 *Corresponding author. Tel.: +49 681 302 2171; fax: +49 681 302 70652
M

14 E-mail address: r.kautenburger@mx.uni-saarland.de (R. Kautenburger)

15
D

16 Address:
TE

17 PD Dr. Ralf Kautenburger


EP

18 WASTe group, Institute of Inorganic Solid State Chemistry

19 Saarland University, Campus Dudweiler, Zeile 5


C

20 D-66125 Saarbrcken
AC

21 Germany

22

23

24

1
ACCEPTED MANUSCRIPT
25 Abstract

26 Along with general scientific developments for the safe long-term storage as well as for

27 disposal of high level radioactive waste (HLW) an onward improvement in the

28 geochemical process understanding is crucial. Natural organic matter (NOM) can play

29 an important role in the immobilisation or mobilisation of these metal ions due to their

PT
30 complexation and colloid formation tendency. In this study, the complexation behavior

31 of humic acids (HA) and other NOM from different sources and its influence on the

RI
32 mobilisation of europium as homologues of the trivalent actinides like americium and

SC
33 uranium as main component of HLW have been analysed in highly saline solutions of

34 CaCl2 and NaCl solutions up to 1 moL L-1 ion strength. Ultrafiltration (UF) in

U
35 combination with inductively coupled plasma mass spectrometry (ICP-MS) has been
AN
36 used for the evaluation of complex stability constants log . In selected cases capillary

37 electrophoresis (CE) coupled to ICP-MS is used as complementary speciation


M

38 technique to verify the UF results. To determine the complex stability constants a


D

39 simple single site model is used. Depending on the source of the analysed NOM log
TE

40 values in the range of 6.8 4.5 for Eu(III) and 6.3 4.5 for U(VI) (UO22+) can be

41 estimated by UF and CE-ICP-MS experiments. Increasing ionic strength (1 moL L-1


EP

42 NaCl) reduces the available complexation sites of NOM resulting in lower amounts of

43 NOM-complexed Eu(III) and U(VI) (UO22+), respectively. The additional presence of


C

44 calcium (0.5 to 5 mmoL L-1 CaCl2 added to 0.1 moL L-1 NaCl) as a bivalent competing
AC

45 cation leads to lower levels of binding to NOM particularly for europium and Aldrich HA

46 as well as for Eu and HA extracted from Gorleben site. In the case of uranyl, the results

47 are different in comparison to europium. Higher amounts of Ca (5 mmoL L-1) in solution

48 increase the levels of NOM complexation for uranyl. In contrast to the results for HA the

49 used Suwannee river NOM reveals log values in the range of nearly two orders of

50 magnitude lower (4.5 4.1 for Eu3+ and 4.8 4.6 for UO22+). Moreover, the three

2
ACCEPTED MANUSCRIPT
51 examined HA from different sources (soil and river HA extracts) show significant

52 differences in their complexation behavior under the geochemical conditions applied in

53 this study.

54

PT
RI
U SC
AN
M
D
TE
C EP
AC

3
ACCEPTED MANUSCRIPT
55 1. Introduction

56 Along with general scientific developments, an improvement in the process

57 understanding of long-lived radionuclides and other toxic heavy metals interacting with

58 the technical, geotechnical, and geological barrier as well as with the total environment

59 is important for the long-term safety assessment of a high level nuclear waste (HLW)

PT
60 disposal site. In this context the actinides uranium, neptunium, plutonium, americium

RI
61 and curium play an important role mainly due to their inherent radiotoxicity. However,

62 they would not be relevant from the viewpoint of causing dose in a distant future after

SC
63 their radioactive decay. The interaction of metals in a potential host rock formation after

64 their release from the waste container into the aquifer has to be analysed to assess

65
U
possible host formations like clay, salt and granite for their suitability as host rock and
AN
66 to provide the necessary information needed for the required safety case (Altmann,

67 2008; Ewing, 2015; Kautenburger and Beck, 2010; McKinley et al., 2006; Poinssot and
M

68 Gin, 2012).
D

69 The objective of this study was to elucidate the complexing behavior of the actinide
TE

70 uranium (main component of HLW) and of the lanthanide europium (used as chemical

71 analogue of long-lived trivalent actinides like americium or curium) with natural organic
EP

72 matter (NOM) from different sources at different geochemical conditions.


C

73 Special attention in this research study is given to the influence of groundwater and
AC

74 clay porewater with high salinity. The use of porewater with high salinity up to 1 moL L-1

75 (NaCl) aims at validating results from another project (Kautenburger, 2009) studying

76 other potential disposal site conditions that are typical for regions in northern or

77 northwest Germany (Brewitz, 1982; Herbert and Schwandt, 2007). It is also known, that

78 beside rock salt (NaCl) also anhydrite (CaSO4) and polyhalite (K2Ca2Mg(SO4)42H2O)

79 are present in the geological layers relevant for a repository in northern or northwest

80 Germany. Therefore, porewater in this region can also contain relevant concentrations

4
ACCEPTED MANUSCRIPT
81 especially of calcium species. Under the aspect of safety analysis of a future disposal

82 site, higher valent cations like Ca are of increased relevance due to their complexation

83 behavior competing with uranium and europium. This competitive behavior of Ca is

84 therefore a further objective of this study.

85 Another focus of this study is the influence of the source and thus the properties of

PT
86 NOM concerning metal complexation. NOM like humic substances play an important

87 role regarding the aquatic chemistry in the aquifer system. Humic material contains a

RI
88 large and heterogeneous group of macromolecules of different molecular weight and

SC
89 charge density with complex structures (Stevenson, 1994 ; Sutton and Sposito, 2005).

90 They can influence the migration and retardation behavior of actinides by their

U
91 complexing, redox and colloidal properties. In addition, the organic components of the
AN
92 natural clay rock low and high molecular weight compounds - might show similar

93 interactions with actinides. Therefore, the complexation behavior of the mentioned


M

94 metal ions with NOM like humic acid (HA) from different sources has been analysed
D

95 and corresponding complexation constants have been derived. Such complex stability
TE

96 constants are required for geochemical modelling of metal-ion behavior in the

97 environment after collecting them in a thermodynamic database (Marquardt, 2008;


EP

98 Marquardt, 2012). Meanwhile, for the determination of complex stability constants a

99 large number of particularly speculative or inconsistent humic ion-binding models and


C

100 their modifications have been proposed in the literature (Dzombak et al., 1986;
AC

101 Hummel, 1997; Kinniburgh et al., 1996; Kinniburgh et al., 1999; Kirishima et al., 2010;

102 Marsac et al., 2011; Milne et al., 2001; Milne et al., 2003; Perdue and Lytle, 1983;

103 Tipping, 1998; Tipping and Hurley, 1992; Tipping et al., 2011; Van Loon et al., 1992).

104 However, in Radiological Performance Assessment (RPA) studies it is essential to

105 predict long-term radionuclide transport. In spite of the improvements in computing,

106 only rather simple chemical parameters, such as Kd-type distribution coefficients can be

5
ACCEPTED MANUSCRIPT
107 utilised. Therefore, the complex metal-ion humic substance interaction models, such as

108 e.g. Model VI, WHAM, and NICA, cannot be used directly in RPA calculations (Bryan et

109 al., 2012). Therefore, among these different complexation models a simple

110 complexation model based on a 1:1 stoichiometry used in this study to determine

111 complex stability constants of the used metals and NOM.

PT
112 Typically applied analytical tools and methods for the determination of complexation

113 constants log consist of batch experiments alone or in a combination with different

RI
114 chemical separation techniques, e.g. centrifugation, extraction methods, ion exchange,

SC
115 ultrafiltration, thin films or precipitation with sensitive detection methods such as

116 nuclear spectroscopy, soft X-ray spectromicroscopy or mass spectrometry (Drozdzak

U
117 et al., 2016; Frhlich et al., 2015; Kim et al., 2005; Kremleva et al., 2013; Shanbhag
AN
118 and Choppin, 1981; Wrobel et al., 2003). Some of these methods are either not able to

119 separate and quantify different metal or metal-organic species in the presence of high
M

120 salinity or they cannot prevent that kinetically unstable complexes may be modified
D

121 during chemical separation. For quantitative elemental speciation analysis some
TE

122 separation techniques like high performance liquid chromatography (HPLC), ion

123 chromatography (IC), gas chromatography (GC) or capillary electrophoresis (CE) have
EP

124 been used in combination with element-sensitive detectors like inductively coupled

125 plasma (ICP), optical emission (OES) and mass spectrometry (MS) or atomic
C

126 absorption spectroscopy (AAS) (Harrison and Rapsomanikis, 1989; Hein et al., 2014;
AC

127 Kautenburger et al., 2006; Michalke, 1999; Mser et al., 2012; Olesik et al., 1995;

128 Prange and Schaumlffel, 1999; Santos et al., 2010).

129 In the present study, ultrafiltration (UF) experiments (Hahn, 2014; Kautenburger and

130 Beck, 2007; Melin and Rautenbach, 2007; Mizera et al., 2005; Pourret et al., 2007) in

131 combination with CE-ICP-MS speciation analysis are performed to determine complex

132 stability constants. Additionally, complexation levels to NOM are determined by UF for

6
ACCEPTED MANUSCRIPT
133 europium and uranium with NOM from different sources especially in the presence of

134 high salinity up to 1.0 M. In selected cases at low ionic strength (0.1 moL L-1 NaCl) CE-

135 ICP-MS is used as complementary speciation and separation technique to verify the

136 results achieved by UF. The UF technique is neither limited by high ionic strength nor

137 by complexation equilibria and has been applied for the uranium speciation for example

PT
138 in surface water (Guo et al., 2007), in the low-salinity zone of a stable estuary

139 (Andersson et al., 2001) and in coastal seawaters (Singhal et al., 2004). Moreover, the

RI
140 UF technique is generally applied for the decontamination of waste water (Rana et al.,

SC
141 2013). In the here presented study, UF in combination with ICP-MS is used to analyse

142 the complexation behavior of europium and uranium and NOM with different molecular

U
143 weight and structure from various sources. Hereby its influence on the mobility of the
AN
144 lanthanide europium and the actinide uranium in the presence of CaCl2 and NaCl at

145 different ionic strengths is investigated. Finally, levels of binding (Stockdale and Bryan,
M

146 2013) to NOM are estimated for Eu and U.


D

147
TE

148 2. Material and methods

149 2.1 Chemicals and standards


EP

150 All standards and chemicals (see also Table 1) were of p.a. (pro analysis) quality or
C

151 better (e.g. emsure or suprapure) and were obtained from Merck (Darmstadt,
AC

152 Germany). Milli-Q ultrapure water (18.2 M cm) was used to prepare all solutions. The

153 single element standards CertiPUR of Eu, U and Ho (1 g L-1) were also obtained from

154 Merck. Ho was diluted 1 : 100 in Milli-Q water and used as spike and internal standard

155 solution for all experiments. For adjusting the pH value perchloric acid (70 %, p.a.),

156 hydrochloric acid (35 %, suprapure) and NaOH (p.a.) from Merck (Darmstadt,

157 Germany) were applied. Argon 5.0 (99.999 %, Praxair Deutschland GmbH, Dsseldorf,

7
ACCEPTED MANUSCRIPT
158 Germany) was used as plasma gas for ICP-MS measurements. For sample

159 preparation, solutions of sodium chloride (Merck, emsure) with an ionic strength of 0.1

160 and 1 moL L-1 were used. Additionally, to the sodium chloride solution different

161 concentrations (0.5 and 5 moL L-1) of CaCl2 were added. For ICP-MS measurement,

162 3.33 mL of the sample were diluted with 6.36 mL Milli-Q water. Additionally, 300 L of

PT
163 conc. HNO3 (69 %, suprapure) were added to adjust a pH value <1 and prevent wall

164 adsorption onto the tube and vial walls. Finally, 10 L of a 10 mg L-1 solution of Ho

RI
165 were added as internal standard to compensate variations of the instrument sensitivity

SC
166 over time.

167

168 2.2 Natural organic matter (NOM)


U
AN
169 Most NOM used in the experiments are commercially available humic acids (HA). One
M

170 humic acid is available from Aldrich (St. Louis, USA; AHA sodium salt), has lignite as

171 source and was purified as described in the literature (Kim et al., 1990). The other
D

172 humic acids are available from the International Humic Substances Society (IHSS, St.
TE

173 Paul, USA) or from a site in Germany. These other humic acids have partly similar

174 sources. The Elliott soil humic acid (EHA) is an extract of fertile prairie soils of Illinois
EP

175 and the Suwannee River natural organic matter (SNOM) is obtained out of the

176 blackwater river with a DOC content in the range of 25-75 mg L-1. Additionally, to the
C

177 commercially available humic acids, an isolated humic acid from Gorleben aquifer
AC

178 system (borehole 851) was used (Artinger et al., 2000; Artinger et al., 1999; Buckau et

179 al., 2000). The Gorleben humic acid (GoHy851) was kindly provided by our project

180 partner KIT-INE (Karlsruhe Institute of Technology - Institute for Nuclear Waste

181 Disposal). For sample preparation, 1 g L-1 solutions of all NOM were prepared. 10 mg

182 of NOM were dissolved in 2 mL of 0.1 moL L-1 NaCl for 24 h. After this time, the

8
ACCEPTED MANUSCRIPT
183 solution was restocked to 10 mL. More information is given in the supplementary

184 material.

185

186 2.3 Ultrafiltration (UF)

187 The determination of log values and levels of binding to NOM is based on the

PT
188 complete separation of uncomplexed and complexed metal ions. For the separation of

RI
189 different species, the method of ultrafiltration was used followed by inductively coupled

190 plasma mass spectrometry to measure the uncomplexed metal ion concentration. The

SC
191 driving force for the separation process by ultrafiltration is a transmembrane pressure

192 differential. The small hydrated ions are able to pass through the membrane with the

193
U
liquid flow whereas the bulky complex composed of metal ion and HA is retained. The
AN
194 concentration of uncomplexed metal ions is measured by ICP-MS assuming that the
M

195 remaining metal formed a metal-HA-complex. In a preliminary study we tested different

196 UF systems: Amicon Ultra 4 mL filter (Merck Millipore, Darmstadt, Germany),


D

197 MicrosepTM Advance Centrifugal Devices 3K Omega Filter (Pall Corporation, New
TE

198 York, USA) and Vivaspin 2 Filter (Sartorius AG, Gttingen, Germany) concerning the

199 recovery rates of the used metal ions Eu and U (data not shown here). Finally, as best
EP

200 UF-material Amicon Ultra 4 Centrifugal Filter Devices with a molecular weight cut-off

201 of 3 kDa from Merck Millipore were used. A copolymer build by styrene and butadiene
C

202 is used as filter device. The membrane consists of a low binding regenerated Ultracel-
AC

203 cellulose. Phase separation takes place by centrifugation with a force effect of 7500 g

204 for 40 minutes using the 5804 R centrifuge from Eppendorf (Hamburg, Germany).

205 For the complexation studies via UF we used a concentration of 0.5 mgL-1 Eu

206 (3.3 mol L-1) or 0.5 mgL-1 U (2.1 mol L-1) and 25 mg L-1 NOM in a final volume of

207 10 mL aqueous solution of 0.1 and 1.0 moL L-1 NaCl. Additionally, 0.5 and 5 moL L-1

208 CaCl2 were added to a 0.1 moL L-1 NaCl solution. The pH values were adjusted with

9
ACCEPTED MANUSCRIPT
209 1 moL L NaOH and 1 moL L-1 HClO4 to pH 5. Before ultrafiltration, the samples were
-1

210 equilibrated for 48 h at 25C. After equilibration, no precipitation of NOM or Me-NOM

211 complexes could be observed (checked by centrifugation and TOC measurements). In

212 order to prove permeability of metal ions through the ultrafiltration membrane and to

213 verify recovery rates of uncomplexed Eu3+ and UO22+ experiments with pure metal ion

PT
214 solutions without HA were performed beforehand.

215 Different types of HA were added to calculate complex stability constants log . For

RI
216 the UF experiments Aldrich-HA (AHA), Elliot soil-HA (EHA), Gorleben 851-HA

SC
217 (GoHy851) and Suwannee River-NOM (SNOM) were used because of the

218 considerable differences in their complexation behavior. Each UF experiment was

U
219 performed as duplicate (two independent experiments under identical conditions).
AN
220
M

221 2.3 Capillary electrophoresis hyphenated with ICP-MS (CE-ICP-MS)

222 For selected complexation experiments at low ionic strength (0.1 moL L-1 NaCl), CE-
D

223 ICP-MS was used as an additional speciation technique to separate NOM complexed
TE

224 and uncomplexed Eu and U species, respectively. For the complexation studies we

225 used a concentration of 0.5 mgL-1 Eu (3.3 moL L-1) or 0.5 mgL-1 U (2.1 moL L-1) and
EP

226 25 mg L-1 NOM in a final volume of 10 mL aqueous solution of 0.1 M NaCl. The pH-
C

227 values of the samples were adjusted with 1 moL L-1 NaOH and 1 moL L-1 HClO4 to pH 5
AC

228 (5.0 0.05). Capillary electrophoresis (CE, Beckman Coulter P/ACE MDQ) was

229 hyphenated by a homemade interface to inductively coupled plasma mass

230 spectrometry (ICP-MS, Agilent 7500cx, Santa Clara, USA). A CE electrolyte buffer

231 consisting of 0.1 moL L-1 acetic acid and 0.01 moL L-1 sodium acetate was used.

232 Detailed analytical conditions are described in previous studies (Kautenburger et al.,

233 2006; Mser et al., 2012).

10
ACCEPTED MANUSCRIPT
234

235

PT
RI
U SC
AN
M
D
TE
EP
C
AC

11
ACCEPTED MANUSCRIPT
236 2.4 Estimation of conservative complex stability constants

237 In this study, complex stability constants of Eu and U with NOM are conservatively

238 described by a simple approach sometimes referred to as single site model (Glaus et

239 al., 2000; Hummel et al., 2000; Kautenburger et al., 2014) or conditional stability

240 constant (Kautenburger and Beck, 2007; Kim and Czerwinski, 1996; Lippold and

PT
241 Lippmann-Pipke, 2009; Neubecker and Allen, 1983). In this approach, the underlying

242 distribution coefficient (Kd = CNOM-complexed metal / Cuncomplexed metal in solution) allows both to

RI
243 estimate the maximum influence of NOM on the complexation of higher valent metal

SC
244 ions and to compare the influence of the different employed NOM on Eu3+ and UO22+

245 complexation at pH 5. In contrast to well-defined organic molecules such as acetate,

U
246 lactate or oxalate, the molar concentration of the used NOM remains unknown.
AN
247 Consequently, we convert the mass of NOM added in our experiments (25 mg L-1) to

248 the available complexing sites per mass unit by correcting the different density of
M

249 binding sites of the NOM by the proton exchange capacity (PEC).
D

250 According to the simple 1:1 model M + L ML complex stability constants are
TE

251 estimated with Ltot = L + ML. Furthermore, the interaction of Eu and U with NOM can be

252 considered as neutralisation of the ionic charge by the NOM. At low metal ion
EP

253 concentrations, the formation of mononuclear complexes is assumed to be the only

254 reaction. Thereby, the group of complexing sites needed to neutralise Me is considered
C

255 as one NOM complexation unit. The reaction is described as


AC

256

257 Me + NOM MeNOM (1)

258

259 Thus, the conditional complex stability constant is expressed as

260

12
ACCEPTED MANUSCRIPT
[MeNOM]
261 = (2)
([Me ]free [NOM]free )

262

263 The concentrations of MeNOM and Mefree (corresponds to metal amount in filtrate after
153 238
264 UF) can be measured in our study via the Eu or U signal by ICP-MS. However,

265 [NOM]free cannot be measured directly and is calculated by

PT
266

267 [NOM]free = [NOM]effective [MeNOM] (3)

RI
268

SC
269 where [NOM]effective is estimated according to

270

U
NOM PEC
271 [NOM]effective = (4)
AN
n
272
M

273 with NOM being the NOM amount in g L-1 and PEC the proton exchange capacity

274 which is determined by potentiometric titration to be 5.0 meq g-1 for the AHA,
D

275 5.9 meq g-1 for EHA, 5.0 meq g-1 for GoHy851 and 7.24 meq g-1 for SNOM at
TE

276 0.1 moL L-1 NaCl, respectively. Molecular association (sometimes abbreviated as LC

277 for loading capacity) at pH 5 is assumed to be one and therefore not included in Eq.
EP

278 (4). Applying this simple binding model, the log values for Eu3+ or UO22+ with NOM

279 from different sites have been conservatively estimated and compared.
C

280
AC

281 2.5 Mass spectrometry with inductively coupled plasma (ICP-MS)

282 An Agilent 7500cx ICP-MS (Santa Clara, USA) with ORS collision cell was used for

283 the isotope measurements. Detailed analytical conditions are given in Table 1.

284

13
ACCEPTED MANUSCRIPT
285 Table 1: Operating parameters of the ICP-MS in spectrum and transient mode,

286 and for the CE-ICP-MS hyphenation

287

ICP-MS Agilent 7500cx (spectrum and transient mode)

RF-power 1550 W

PT
Cooling / auxiliary gas 15.0 / 1.05 L min-1

Dwell times / Repetition 100 ms per mass / 3 times

RI
Sample uptake parameters for transient mode

SC
Total analysis / scan time 800 s

Amounts of samples 6 per run

Sample uptake 10 s
U
AN
Washing time 90 s

CE Beckman Coulter P/ACE MDQ


M

Capillary fused silica (Polymicro Technologies)


D

Capillary dimensions / 74 m ID, 362 mm OD, 80 cm length / 25C


TE

Temperature

CE electrolyte buffer 100 mmoL L-1 acetic acid, 10 mmoL L-1 Na-acetate,
EP

pH 3.7

Internal Standard (IS) CE 200 ppb Cs


C

buffer
AC

DC-voltage / current + 30 kV / 1618 A

Interface homemade

Spray-chamber Cinnabar cyclonic, chilled at 4C, 20 mL volume

Nebuliser MicroMist 50 L, Type AR35-1-FM005E, 2.8 bar

Make-up fluid 2 % HNO3, 24 nmol L-1 (4 ppb) Ho (IS), 112 L min-1

14
ACCEPTED MANUSCRIPT
Samples

Eu, Ho, U ICP-standards CertiPUR (Merck), diluted in Milli-Q

153
Analysed isotopes Eu, 238U (analytes), 165Ho (internal standard),

Natural organic matter (NOM) AHA (Aldrich H1,675-2)


EHA (IHSS)
GoHy851 (Gorleben site)

PT
SNOM (IHSS)

Ion strength in all samples 0.1 moL L-1 NaCl

RI
0.1 moL L-1 NaCl + 0.5 mmoL L-1 CaCl2
0.1 moL L-1 NaCl + 50 mmoL L-1 CaCl2

SC
1.0 moL L-1 NaCl

pH value 5.0 0.05

U
AN
288 The helium collision mode is used for reliable, predictable removal of unknown
M

289 matrix interferences. ICP-MS measurements of real samples with higher ionic strength

290 are a general problem. Measurements of samples with high salt content, like sea water,
D

291 are generally possible only with a high matrix introduction system (HMI) for the 7500
TE

292 and 7700 series. The HMI is an online dilution system of the sample aerosol before

293 sample ionisation in the plasma torch. The maximum of the total dissolved solids (TDS)
EP

294 should be at 3 %, which increases the matrix tolerance about ten times.
C

295 In this work we used samples with 1 moL L-1 ionic strength. A 1 moL L-1 solution
AC

296 shows a TDS of nearly 6 % which is too much for the HMI. As an alternative to the

297 dilution procedure (which results in a lower detection sensitivity) described above a

298 matrix separation step during the sample preparation might lead to success. But

299 preliminary tests showed recovery problems with uranium as analyte element.

300 Additionally, for every analyte an optimisation step is necessary. Therefore, a transient

301 method of ICP-MS analysis for high saline samples up to 5 moL L-1 NaCl has been

15
ACCEPTED MANUSCRIPT
302 developed using the existing ICP-MS instrumentation (Hein et al.) and is used here for

303 the analysis of the samples containing 1 moL L-1 NaCl.

304

305 3. Results and discussion

PT
306 3.1 Europium-NOM complexation

307 Depending on the ionic strength in the sample solution the recovery of dissolved

RI
308 Eu3+(aq) ions in the filtrate ranged between 97.8 % (0.1 moL L-1 NaCl) and 106.3 % (1.0

SC
309 moL L-1 NaCl) as shown in Table 2. Compared to CE-ICP-MS where the recovery of Eu

310 is determined in the presence of NOM UF revealed a higher recovery rate for

U
311 uncomplexed Eu ions at identical conditions used.
AN
312

313 Table 2: Eu-recovery (initial concentration: 0.5 mg L-1 Eu) in the UF filtrate ( SD)
M

314 in the presence of different ionic strengths at pH 5 compared to CE-ICP-MS (0.3


D

315 mg L-1 Eu at pH 5, mean of all four NOM)


TE

316

317 Eu recovery [%]


EP

318 Conditions UF CE-ICP-MS


C

0.1 moL L-1 NaCl 97.8 0.8 82.9 8.5


AC

0.1 moL L-1 NaCl nd*


-1 2+
98.5 0.1
+0.5 mmoL L Ca

0.1 moL L-1 NaCl nd


98.3 0.2
+5.0 mmoL L-1 Ca2+

1.0 moL L-1 NaCl 106.3 4.0 nd

319 *nd: not determined

16
ACCEPTED MANUSCRIPT
320 The permeability of uncomplexed metal ions through the UF filter membrane is

321 thereby ensured for the subsequent complexation experiments. Table 3 shows the

322 calculated log values for the Eu3+-complexation with the different NOM (25 mg L-1,

323 respectively) revealed by the UF experiments followed by ICP-MS determination of the

324 uncomplexed Eu in the filtrate (ICP-MS measurement results can be found in the

PT
325 supplementary material). EHA obtained from a prairie soil extract forms very stable

326 complexes with Eu (mean log value of 6.77 0.09 revealed by UF at 0.1 moL L-

RI
1
327 NaCl ionic strength) followed by AHA (6.36 0.02) and GoHy851 (5.96 0.03).

SC
328

329 Table 3: Calculated log values ( SD) for Eu3+ complexation by UF (cEu in the UF

330
U
filtrate is corrected by the Eu recovery rates from Table 2) in comparison to CE-
AN
331 ICP-MS (0.1 moL L-1 NaCl); influence of ionic strength on complexation rate in %
M

332

Conditions AHA EHA GoHy851 SNOM


D

UF 6.36 0.02 6.77 0.09 5.96 0.03 4.53 0.01


TE

CE-ICP-MS 6.1 0.1 6.5 6.7 4.6


EP

Levels of binding
AHA EHA GoHy851 SNOM
by UF [%]
C

0.1 moL L-1 NaCl 98.9 99.6 97.2 66.2


AC

0.1 moL L-1 NaCl


96.7 99.6 97.1 65.7
+0.5 mmoL L-1 Ca2+

0.1 moL L-1 NaCl


96.9 99.4 95.6 61.7
+5.0 mmoL L-1 Ca2+

1.0 moL L-1 NaCl 92.0 99.1 82.8 43.3

17
ACCEPTED MANUSCRIPT
333 The complex stability for SNOM (which is extracted from a blackwater river in

334 contrast to soil extracts in the case of the other HA) is over two log units lower (4.53

335 0.01) compared to the log value evaluated for Eu and EHA. Apart from the complex

336 stability constant for Eu and GoHy851 being considerable higher determined by CE-

337 ICP-MS (6.7) compared to the results from UF (6.0). An explanation can be found

PT
338 assuming an overestimation of uncomplexed Eu by the UF method resulting in lower

339 log values. This expectation can be caused by a breakthrough of lower molecular Eu-

RI
340 NOM species of GoHy851 through the UF membrane, which also have been observed

SC
341 in a previous study (Kautenburger and Beck, 2007). The values for Eu and the other

342 three NOM are in a good agreement with the values derived by UF experiments as

U
343 shown in Table 3.
AN
344 As a consequence, a release of trivalent metal ions from precipitated NOM

345 complexes seems to be rather possible. At increasing ionic strength, the levels of
M

346 binding for Eu3+ decrease in the case of all used NOM. The presence of 1 moL L-1 ionic
D

347 strength (NaCl) in the sample solution leads for all NOM to a significant decrease of the
TE

348 complexation. Especially the levels of binding for both NOM with the lowest complex

349 stability constants at low ionic strength (Eu-GoHy851 and Eu-SNOM) show a relevant
EP

350 influence of their binding levels depending on the ionic strength (Table 3). For Eu-

351 GoHy851 we could determine a decrease in the complexation level of about 15 %


C

352 whereas the highest decrease for Eu-SNOM in the range of about 23 % can be
AC

353 observed. In summary, a ranking of the complex stability of the formed Eu-NOM

354 complexes rises in the following order:

355 SNOM < GoHy851 < AHA < EHA.

356 The same ranking has been found in a study for Suwannee river fulvic acid

357 compared to AHA at ionic strength up to 0.1 moL L-1 NaClO4 (Glaus et al., 2000). The

358 lowest complex stability constants are found for SNOM which represents a river extract

18
ACCEPTED MANUSCRIPT
359 containing the complete fraction of natural organic matter including fulvic and humic

360 acids as well as other NOM. Similar to the low log values, the calculated levels of

361 binding for Eu and SNOM are on the lowest level of all used NOM ranging from 66.2 %

362 (0.1 moL L-1 NaCl) to 43.3 % (1.0 moL L-1 NaCl). Compared to this finding, EHA with

363 the highest log values of all tested NOM in the UF experiments (6.77) only shows a

PT
364 slight decrease of the complexation level for Eu ranging from 99.6 % to 99.1 %

365 (between 0.1 and 1.0 moL L-1 NaCl) which is in the same order of magnitude. In

RI
366 contrast to these findings, the influence of salinity on complexation levels for the two

SC
367 other examined humic acids is clearly higher. Particularly the decrease of the binding

368 level for the europium-AHA complex from 99 % (0.1 moL L-1 NaCl) to 92 %

U
369 (1.0 moL L-1 NaCl) is obvious. Furthermore, the presence of relatively low
AN
370 concentrations of bivalent calcium ions (0.005 moL L-1 Ca2+) at 0.1 moL L-1 NaCl ionic

371 strength leads to a decrease of the binding level of about 3%. These results show the
M

372 negative influence of higher valent competing cations on the humate complexation of

373 trivalent metal ions like europium leading to a higher remobilisation of these metals out
D

374 of the humate complex in the aquifer of a potential HLW site. Subsequently, these
TE

375 released metal ions are exposed to a competition between sorption onto the host rock
EP

376 and complexation by other NOM present. In this context, a former study dealing with

377 the ternary system Eu AHA kaolinite showed that the presence of HA leads to a
C

378 reduced sorption affinity of Eu towards kaolinite as a model clay for the host rock.
AC

379 Thus, an increased mobility of Eu in the aquifer can be deduced (Kautenburger and

380 Beck, 2008). Similarly, competition between HA complexation and Eu sorption onto

381 clay colloids has been examined showing that humate complexation of Eu dominates

382 over sorption onto bentonite colloids (Bouby et al., 2011). An important finding that

383 underlines that the suitability of a potential waste site has to be judged also by the

384 organic substances present in the host rock.

19
ACCEPTED MANUSCRIPT
385

386 3.2 Uranyl-NOM complexation

387 In comparison to Eu3+ (mean Eu recovery of 100.2 % 4.1 % for UF) the recovery

388 UO22+ is lower (Table 4). Depending on the ionic strength in the solution the recovery of

389 dissolved uranyl in the UF filtrate ranged between 85.1 to 93.4 % (mean U recovery of

PT
390 89.8 % 3.4 % for UF).

RI
391

Table 4: U-recovery (initial concentration: 0.5 mg L-1 U) in the UF filtrate ( SD) in

SC
392

393 the presence of different ionic strengths at pH 5 compared to CE-ICP-MS

U
394 (0.3 mg L-1 U at pH 5, mean of all four NOM)
AN
395

U recovery [%]
M

Conditions UF CE-ICP-MS
D

0.1 moL L-1 NaCl 85.1 3.9 83.3 6.2


TE

0.1 moL L-1 NaCl nd*


90.6 0.7
+0.5 mmoL L-1 Ca2+
EP

0.1 moL L-1 NaCl nd


-1 2+
89.9 0.9
+5.0 mmoL L Ca
C

1.0 moL L-1 NaCl 93.4 9.9 nd


AC

396 *nd: not determined

397

398 Regarding the influence of salinity on the binding level, the uranyl complexation is

399 not relevantly influenced by an increasing ionic strength. The calculated log values as

400 well as the levels of binding are given in Table 5 (ICP-MS results can be found in the

401 supplementary material). Comparable to the results for Eu likewise UO22+ forms very

20
ACCEPTED MANUSCRIPT
402 stable complexes with EHA (log value of 6.08 0.01 for the UF experiments and

403 6.3 0.2 for CE-ICP-MS) which represents a humic acid soil extract. In contrast to Eu

404 the UO22+ complex with GoHy851 (5.93 0.06) seems to have a similar complex

405 stability than the uranyl-AHA complex (5.99 0.1) determined by UF. However, using

406 CE-ICP-MS the UO22+ complex with GoHy851 seems more stable (6.1) than the uranyl-

PT
407 AHA complex (5.3 0.3, n=3).

408

RI
409 Table 5: Calculated log values ( SD) for the UO22+ complexation applying UF

SC
410 (cU in the UF filtrate is corrected by the U recovery rates from Table 4) in

411 comparison to CE-ICP-MS (0.1 moL L-1 NaCl); influence of ionic strength on

U
412 levels of binding in %
AN
413
M

Conditions AHA EHA GoHy851 SNOM

UF 5.99 0.01 6.08 0.01 5.93 0.06 4.67 0.02


D

CE-ICP-MS 5.3 0.3 6.3 0.2 6.1 4.5


TE

Levels of binding
AHA EHA GoHy851 SNOM
by UF [%]
EP

0.1 moL L-1 NaCl 98.4 98.9 98.1 80.5


C

0.1 moL L-1 NaCl


96.7 99.0 98.4 82.6
AC

+0.5 mmoL L-1 Ca2+

0.1 moL L-1 NaCl


97.8 99.2 98.9 83.2
+5.0 mmoL L-1 Ca2+

1.0 moL L-1 NaCl 97.9 99.2 98.4 78.6

21
ACCEPTED MANUSCRIPT
414 All other results are in the same order of magnitude when comparing UF and CE-

415 ICP-MS results. Nevertheless, the complex stability constants for uranyl and AHA at

416 low ionic strength (0.1 M NaCl) derived by both methods are in a good agreement with

417 other studies using different geochemical conditions, experimental techniques, and log

418 evaluations (Pashalidis and Buckau, 2007; Sachs et al., 2007; Zeh et al., 1997) as

PT
419 well as the results for uranyl and SNOM (Glaus et al., 2000; Lubal et al., 2000; Mser

420 et al., 2012; Zhu and Ryan, 2016).

RI
421 The complex stability for SNOM (extracted from a blackwater river) is nearly 1.5 log

SC
422 units lower (4.67 0.02) compared to the log value evaluated for the other three used

423 HA with uranyl. Interestingly and in contrast to the findings for Eu no relevant influence

U
424 on the levels of binding for uranyl and NOM (except for AHA) in the presence of 0.1
AN
425 moL L-1 Na+ and 0.005 or 0.05 moL L-1 Ca2+ as well as for 1.0 moL L-1 NaCl can be

426 observed. Quite contrary to Eu no moderate decrease (more probably a slight


M

427 increase) of the complexation level for U in the presence of Ca can be observed.
D

428 Summarising the results for uranyl, the stability of the formed U(VI)-NOM complexes
TE

429 rises in the following order:

430 SNOM < AHA < GoHy851 < EHA.


EP

431 Equally to the europium complexation the complex stability constants for the uranyl

432 complexation with SNOM (4.7) are the lowest, while the log values for uranyl and HA
C

433 with values between 5.3 and 6.1 are in a slightly lower range compared to Eu (6.0
AC

434 6.8).

435

436 3.3 Comparison of NOM complexation between europium and uranyl

437 Differences in the complexation of Eu and U exist regarding the strength of the

438 complexing ligands. In the presence of strongly complexing ligands such as EHA, the

22
ACCEPTED MANUSCRIPT
439 europium humate formation shows higher log values (log Eu-EHA: 6.5 - 6.8) in

440 comparison with the uranyl complexation (log U-EHA: 6.1 - 6.3). Minor differences

441 between Eu3+ and UO22+ humate formation are given by complexation with weakly

442 binding ligands such as SNOM (Fig. 1).

443 A significant influence of the ionic strength on the levels of binding can only be

PT
444 observed for the Eu-NOM complexes where the complexation rates decrease about

445 15 % down to 83 % at higher ionic strength (Table 3). Increasing ionic strength (0.1 up

RI
446 to 1 M NaCl) reduces the complexation ability of the used HA especially for the trivalent

SC
447 Eu. For the uranyl complexation the influence of high NaCl salinity can be considered

448 as negligible. The presence of calcium (0.5 to 5 mM CaCl2 added to 0.1 M NaCl) as

U
449 competing cation leads to lower levels of binding particularly for europium (mainly for
AN
450 AHA with a decrease of 7 % and for GoHy851 with a decrease of the binding level of

451 nearly 15 %). For uranyl the results are contrary to europium. Higher amounts of Ca
M

452 (5 mmoL L-1) in the solution could increase the complexation rates of uranyl possibly by
D

453 the formation of stable ternary calcium uranyl humate complexes conceivably
TE

454 comparable to the very stable ternary calcium uranyl carbonate complexes (Dong and

455 Brooks, 2006; Gartman et al., 2015). A possible explanation for the deviance between
EP

456 Eu and UO22+ complexation could hypothetically be caused by the formation of micelle-

457 like aggregates especially for Eu and the examined humic acids as considered below.
C

458 Probably the calcium uranyl species forms a more stable complex with HA than the
AC

459 divalent uranyl cation alone. Additionally, the formation of predominantly inner-sphere

460 like complexes in the presence of higher ionic strength shows no relevant influence on

461 the complexation of uranyl by HA.

462 It is known from literature (Kautenburger, 2009; Kautenburger et al., 2014; Kouhail

463 et al.; Marsac et al., 2010; Pourret and Martinez, 2009; Tan et al., 2014) that higher

464 molecular NOM like the complex humic acids possess strong as well as weak binding

23
ACCEPTED MANUSCRIPT
465 sites for higher valent metal ions. As a consequence, metal ions at first occupy the HA

466 strong binding sites followed by structural changes of the humic acids, mainly resulting

467 in the formation of micelles in the solution (Gamboa and Olea, 2006; Gledhiir and Buck,

468 2012; Terashima et al., 2004; Timerbaev, 2007).

469

PT
RI
U SC
AN
M
D
TE

470
EP

Figure 1: Influence of ionic strength on the levels of binding to a) AHA, b) EHA,

3+ 2+
c) GoHy851 d) SNOM for Eu (blue) and UO2 (red)
C
AC

471 Assuming such a formation of micelles takes place under the parameters chosen in

472 this study the non-polar parts of the HA are directed outward of the aggregates along

473 with many (presumably weaker) negatively loaded binding sites. In our study, we used

474 uranyl at lower concentrations (2.1 moL L-1) compared to europium (3.3 moL L-1).

475 The metal ions complexed by the HA strong binding sites are predominantly inside

476 these micelles resembling some kind of an inner-sphere complex thus being well

24
ACCEPTED MANUSCRIPT
477 protected against competing cations like Na+ or Ca2+ from the high saline solutions

478 present in the matrix. Due to the threefold positive charge in combination with the lower

479 steric hindrance of the Eu3+ cation compared to UO22+ a structural change of the

480 humate complex as described above should be more likely. Consequently, also due to

481 the higher Eu concentration used in the experiments a higher amount of the added

PT
482 Eu3+ cations are predominantly outside these micelles. However, this complexed Eu

483 could be easier influenced by competing cations especially by higher valent cations like

RI
484 Ca2+. This might be a possible explanation why the influence of the ionic strength is

SC
485 obviously higher for Eu in comparison to uranyl, which should be verified in further

486 experiments.

U
487
AN
488 4 Conclusions
489 Objective of the research project was to analyse the interaction of the actinide uranium
M

490 and the lanthanide europium as chemical homologue of trivalent actinides like
D

491 americium with NOM from different sources. Special attention was given to the
TE

492 influence of groundwater and clay porewater with high salinity. In this study,

493 complexation experiments between metal ions (Eu3+ and UO22+) and NOM from
EP

494 different sources (AHA, EHA, GoHy851 and SNOM) at ionic strengths up to 1 M NaCl

495 have been performed by UF in combination with ICP-MS, and in selected cases also by
C

496 CE-ICP-MS. Increasing ionic strength reduces the complexation ability of the used HA
AC

497 especially for the trivalent Eu. For the uranyl complexation the influence of high NaCl

498 salinity can be considered as negligible. The presence of calcium as competing cation

499 leads to lower levels of binding particularly for europium but not for uranyl. Higher

500 amounts of Ca in the solution lead to an increase of the complexation rates for uranyl

501 possibly by the formation of stable ternary calcium uranyl humate complexes.

502 Furthermore, the different behavior between Eu and UO22+ complexation for the

25
ACCEPTED MANUSCRIPT
503 examined HA could be caused by the formation of micelle-like aggregates especially

504 for Eu. In contrast to the results for HA the Suwannee river NOM reveals for both metal

505 ions log values in the range of nearly two orders of magnitude lower (4.5 4.6 for

506 Eu3+ and 4.5 4.7 for UO22+). Additionally, the influence of higher ionic strength on the

507 levels of binding to SNOM for both metals is obvious. Apart from the influence of

PT
508 salinity, also the source of the three examined HA (soil and river HA extracts) shows

509 significant impact on their complexation behavior under the geochemical conditions

RI
510 applied in this study. In a nutshell, site specific NOM is expected to have a

SC
511 considerable influence on the safety case of a potential HLW disposal site.

512

U
513 Acknowledgement
AN
514 This work was carried out within the BMWi (German Federal Ministry for Economic

515 Affairs and Energy) project Retardation of radionuclides relevant to waste disposal in
M

516 clay stone and saline systems. The authors therefore thank the BMWi for financial
D

517 support through grant no. 02E10991 and our project partners for the kind collaboration.

518
TE
C EP
AC

26
ACCEPTED MANUSCRIPT
519 References

520 Altmann, S., 2008. Geochemical research: A key building block for nuclear waste
521 disposal safety cases. J. Contam. Hydrol. 102, 174-179.

522 Andersson, P.S., Porcelli, D., Gustafsson, O., Ingri, J., Wasserburg, G.J., 2001. The
523 importance of colloids for the behavior of uranium isotopes in the low-salinity
524 zone of a stable estuary. Geochim. Cosmochim. Acta Acta 65, 13-25.

PT
525 Artinger, R., Buckau, G., Geyer, S., Fritz, P., Wolf, M., Kim, J.I., 2000. Characterization
526 of groundwater humic substances: Influence of sedimentary organic carbon.

RI
527 Appl. Geochem. 15, 97-116.

528 Artinger, R., Buckau, G., Kim, I.J., Geyer, S., 1999. Characterization of groundwater

SC
529 humic and fulvic acids of different origin by GPC with UV/Vis and fluorescence
530 detection. Fresen. J. Anal. Chem. 364, 737-745.

531
U
Bouby, M., Geckeis, H., Ltzenkirchen, J., Mihai, S., Schfer, T., 2011. Interaction of
AN
532 bentonite colloids with Cs, Eu, Th and U in presence of humic acid: A flow field-
533 flow fractionation study. Geochim. Cosmochim. Acta Acta 75, 3866-3880.
M

534 Brewitz, W., 1982. Eignungsprfung der Schachtanlage Konrad fr die Endlagerung
535 radioaktiver Abflle: GSF-Abschlussbericht T136.
D

536 Bryan, N.D., Abrahamsen, L., Evans, N., Warwick, P., Buckau, G., Weng, L., Van
TE

537 Riemsdijk, W.H., 2012. The effects of humic substances on the transport of
538 radionuclides: Recent improvements in the prediction of behaviour and the
539 understanding of mechanisms. Appl. Geochem. 27, 378-389.
EP

540 Buckau, G., Artinger, R., Fritz, P., Geyer, S., I. Kim, J., Wolf, M., 2000. Origin and
541 mobility of humic colloids in the Gorleben aquifer system. Appl. Geochem. 15,
C

542 171-179.
AC

543 Dong, W., Brooks, S.C., 2006. Determination of the formation constants of ternary
544 complexes of uranyl and carbonate with alkaline earth metals (Mg2+, Ca2+, Sr 2+,
545 and Ba2+) using anion exchange method. Environ. Sci. Technol. 40, 4689-4695.

546 Drozdzak, J., Leermakers, M., Gao, Y., Elskens, M., Phrommavanh, V., Descostes, M.,
547 2016. Uranium aqueous speciation in the vicinity of the former uranium mining
548 sites using the diffusive gradients in thin films and ultrafiltration techniques. Anal.
549 Chim. Acta 913, 94-103.

27
ACCEPTED MANUSCRIPT
550 Dzombak, D.A., Fish, W., Morel, F.M.M., 1986. Metal-humate interactions. I. Discrete
551 ligand and continuous distribution models. Environ. Sci. Technol. 20, 669-675.

552 Ewing, R.C., 2015. Long-term storage of spent nuclear fuel. Nat. Mater. 14, 252-257.

553 Frhlich, D.R., Skerencak-Frech, A., Bauer, N., Rossberg, A., Panak, P.J., 2015. The
554 pH dependence of Am(III) complexation with acetate: An EXAFS study. J.
555 Synchrotron Radiat. 22, 99-104.

PT
556 Gamboa, C., Olea, A.F., 2006. Association of cationic surfactants to humic acid: Effect
557 on the surface activity. Colloids Surf. A Physicochem. Eng. Asp. 278, 241-245.

RI
558 Gartman, B.N., Qafoku, N.P., Szecsody, J.E., Kukkadapu, R.K., Wang, Z., Wellman,
559 D.M., Truex, M.J., 2015. Uranium fate in Hanford sediment altered by simulated

SC
560 acid waste solutions. Appl. Geochem. 63, 1-9.

561 Glaus, M.A., Hummel, W., Van Loon, L.R., 2000. Trace metal-humate interactions. I.

U
562 Experimental determination of conditional stability constants. Appl. Geochem. 15,
AN
563 953-973.

564 Gledhiir, M., Buck, K.N., 2012. The organic complexation of iron in the marine
M

565 environment: A review. Front. Microbiol. 3.

566 Guo, L., Warnken, K.W., Santschi, P.H., 2007. Retention behavior of dissolved uranium
D

567 during ultrafiltration: Implications for colloidal U in surface waters. Mar. Chem.
TE

568 107, 156-166.

569 Hahn, R., 2014. Untersuchungen zur Metall-Komplexierung von Europium und Uran
EP

570 mit organischen Komplexliganden mittels Dialyse und Ultrafiltration. Masterarbeit,


571 Naturwissenschaftlich-Technische Fakultt III, Universitt des Saarlandes,
572 Saarbrcken.
C

573 Harrison, R.M., Rapsomanikis, S., 1989. Environmental Analysis Using


AC

574 Chromatography Interfaced with Atomic Spectroscopy. John Wiley & Sons, New
575 York.

576 Hein, C., Sander, J.M., Kautenburger, R., 2014. Speciation via Hyphenation Metal
577 Speciation in Geological and Environmental Samples by CE-ICP-MS. J. Anal.
578 Bioanal. Tech. 5, 225.

28
ACCEPTED MANUSCRIPT
579 Hein, C., Sander, J.M., Kautenburger, R., New Approach of a transient ICP-MS

580 measurement method for samples with high salinity. Talanta,

581 http://dx.doi.org/10.1016/j.talanta.2016.06.059.

582 Herbert, H.-J., Schwandt, A., 2007. Salzlsungszuflsse im Salzbergbau


583 Mitteldeutschlands. Erfassung und Bewertung der chemischen und

PT
584 physikalischen Analyseergebnisse. GRS-Berichte 226.

585 Hummel, W., 1997. Binding models for humic substances. Modelling in Aquatic

RI
586 Chemistry, 153-206.

587 Hummel, W., Glaus, M.A., Van Loon, L.R., 2000. Trace metal-humate interactions. II.

SC
588 The 'conservative roof' model and its application. Appl. Geochem. 15, 975-1001.

589 Kautenburger, R., 2009. Influence of metal concentration and the presence of

U
590 competing cations on europium and gadolinium speciation with humic acid
AN
591 analysed by CE-ICP-MS. J. Anal. At. Spectrom. 24, 934-938.

592 Kautenburger, R., Beck, H.P., 2007. Complexation studies with lanthanides and humic
M

593 acid analyzed by ultrafiltration and capillary electrophoresis-inductively coupled


594 plasma mass spectrometry. J. Chromatogr. A 1159, 75-80.
D

595 Kautenburger, R., Beck, H.P., 2008. Waste disposal in clay formations: Influence of
TE

596 humic acidon the migration of heavy-metal pollutants. ChemSusChem 1, 295-


597 297.

598 Kautenburger, R., Beck, H.P., 2010. Influence of geochemical parameters on the
EP

599 sorption and desorption behaviour of europium and gadolinium onto kaolinite. J.
600 Environ. Monit. 12, 1295-1301.
C

601 Kautenburger, R., Hein, C., Sander, J.M., Beck, H.P., 2014. Influence of metal loading
AC

602 and humic acid functional groups on the complexation behavior of trivalent
603 lanthanides analyzed by CE-ICP-MS. Anal. Chim. Acta 816, 50-59.

604 Kautenburger, R., Nowotka, K., Beck, H.P., 2006. Online analysis of europium and
605 gadolinium species complexed or uncomplexed with humic acid by capillary
606 electrophoresis-inductively coupled plasma mass spectrometry. Anal. Bioanal.
607 Chem. 384, 1416-1422.

608 Kim, H.J., Baek, K., Kim, B.K., Yang, J.W., 2005. Humic substance-enhanced
609 ultrafiltration for removal of cobalt. J. Hazard. Mater. 122, 31-36.
29
ACCEPTED MANUSCRIPT
610 Kim, J.I., Buckau, G., Li, G.H., Duschner, H., Psarros, N., 1990. Characterization of
611 humic and fulvic acids from Gorleben groundwater. Fresen. J. Anal. Chem. 338,
612 245-252.

613 Kim, J.I., Czerwinski, K.R., 1996. Complexation of Metal Ions with Humic Acid: Metal
614 Ion Charge Neutralization Model. Radiochim. Acta 73, 5-10.

615 Kinniburgh, D.G., Milne, C.J., Benedetti, M.F., Pinheiro, J.P., Filius, J., Koopal, L.K.,

PT
616 Van Riemsdijk, W.H., 1996. Metal ion binding by humic acid: Application of the
617 NICA-Donnan model. Environ. Sci. Technol. 30, 1687-1698.

RI
618 Kinniburgh, D.G., Van Riemsdijk, W.H., Koopal, L.K., Borkovec, M., Benedetti, M.F.,
619 Avena, M.J., 1999. Ion binding to natural organic matter: Competition,

SC
620 heterogeneity, stoichiometry and thermodynamic consistency. Colloids Surf. A
621 Physicochem. Eng. Asp. 151, 147-166.

U
622 Kirishima, A., Ohnishi, T., Sato, N., Tochiyama, O., 2010. Simplified modelling of the
AN
623 complexation of humic substance for equilibrium calculations. J. Nucl. Sci.
624 Technol. 47, 1044-1054.

625 Kouhail, Y.Z., Benedetti, M.F., Reiller, P.E., 2016. Eu(III)-Fulvic Acid Complexation:
M

626 Evidence of Fulvic Acid Concentration Dependent Interactions by Time-Resolved


627 Luminescence Spectroscopy. Environ. Sci. Technol. 50, 3706-3713.
D

628 Kremleva, A., Krger, S., Rsch, N., 2013. Assigning EXAFS results for uranyl
TE

629 adsorption on minerals via formal charges of bonding oxygen centers. Surf. Sci.
630 615, 21-25.
EP

631 Lippold, H., Lippmann-Pipke, J., 2009. Effect of humic matter on metal adsorption onto
632 clay materials: Testing the linear additive model. J. Contam. Hydrol. 109, 40-48.
C

633 Lubal, P., Fetsch, D., irok, D., Lubalov, M., enkr, J., Havel, J., 2000.
AC

634 Potentiometric and spectroscopic study of uranyl complexation with humic acids.
635 Talanta 51, 977-991.

636 Marquardt, C.M., 2008. Migration of actinides in the system Clay, Humic substances,
637 Aquifer. Wissenschaftliche Berichte, FZKA 7407.

638 Marquardt, C.M., 2012. Interaction and Transport of Actinides in Natural Clay Rock with
639 Consideration of Humic Substances and Clay Organic Compounds. KIT Scientific
640 Reports 7633.

30
ACCEPTED MANUSCRIPT
641 Marsac, R., Davranche, M., Gruau, G., Bouhnik-Le Coz, M., Dia, A., 2011. An
642 improved description of the interactions between rare earth elements and humic
643 acids by modeling: PHREEQC-Model VI coupling. Geochim. Cosmochim. Acta
644 Acta 75, 5625-5637.

645 Marsac, R., Davranche, M., Gruau, G., Dia, A., 2010. Metal loading effect on rare earth
646 element binding to humic acid: Experimental and modelling evidence. Geochim.

PT
647 Cosmochim. Acta Acta 74, 1749-1761.

648 McKinley, I.G., Neall, F.B., Kawamura, H., Umeki, H., 2006. Geochemical optimisation

RI
649 of a disposal system for high-level radioactive waste. J. Geochem. Explor. 90, 1-
650 8.

SC
651 Melin, T., Rautenbach, R., 2007. Membranverfahren. Springer-Verlag, Berlin-
652 Heidelberg.

U
653 Michalke, B., 1999. Potential and limitations of capillary electrophoresis inductively
AN
654 coupled plasma mass spectrometry. Proceedings of the 1999 European Winter
655 Conference on Plasma Spectrochemistry 14, 1297-1302.

656 Milne, C.J., Kinniburgh, D.G., Tipping, E., 2001. Generic NICA-Donnan model
M

657 parameters for proton binding by humic substances. Environ. Sci. Technol. 35,
658 2049-2059.
D

659 Milne, C.J., Kinniburgh, D.G., Van Riemsdijk, W.H., Tipping, E., 2003. Generic NICA -
TE

660 Donnan model parameters for metal-ion binding by humic substances. Environ.
661 Sci. Technol. 37, 958-971.
EP

662 Mizera, J., Mizerov, G., Bene, P., 2005. Radiotracer study of europium interaction
663 with humic acid using electrophoresis, ultrafiltration, and dialysis. J. Radioanal.
C

664 Nucl. Chem. 263, 75-80.


AC

665 Mser, C., Kautenburger, R., Philipp Beck, H., 2012. Complexation of europium and
666 uranium by humic acids analyzed by capillary electrophoresis-inductively coupled
667 plasma mass spectrometry. Electrophoresis 33, 1482-1487.

668 Neubecker, T.A., Allen, H.E., 1983. The measurement of complexation capacity and
669 conditional stability constants for ligands in natural waters. Water Res. 17, 1-14.

670 Olesik, J.W., Kinzer, J.A., Olesik, S.V., 1995. Capillary electrophoresis inductively
671 coupled plasma spectrometry for rapid elemental speciation. Anal. Chem. 67, 1-
672 12.

31
ACCEPTED MANUSCRIPT
673 Pashalidis, I., Buckau, G., 2007. U(VI) mono-hydroxo humate complexation. J.
674 Radioanal. Nucl. Chem. 273, 315-322.

675 Perdue, E.M., Lytle, C.R., 1983. Distribution model for binding of protons and metal
676 ions by humic substances. Environ. Sci. Technol. 17, 654-660.

677 Poinssot, C., Gin, S., 2012. Long-term Behavior Science: The cornerstone approach
678 for reliably assessing the long-term performance of nuclear waste. J. Nucl. Mater.

PT
679 420, 182-192.

680 Pourret, O., Davranche, M., Gruau, G., Dia, A., 2007. Competition between humic acid

RI
681 and carbonates for rare earth elements complexation. J. Colloid Interface Sci.
682 305, 25-31.

SC
683 Pourret, O., Martinez, R.E., 2009. Modeling lanthanide series binding sites on humic
684 acid. J. Colloid Interface Sci. 330, 45-50.

U
685 Prange, A., Schaumlffel, D., 1999. Determination of element species at trace levels
AN
686 using capillary electrophoresis-inductively coupled plasma sector field mass
687 spectrometry. Proceedings of the 1999 European Winter Conference on Plasma
688 Spectrochemistry 14, 1329-1332.
M

689 Rana, D., Matsuura, T., Kassim, M.A., Ismail, A.F., 2013. Radioactive decontamination
D

690 of water by membrane processes A review. Desalination 321, 77-92.


TE

691 Sachs, S., Brendler, V., Geipel, G., 2007. Uranium(VI) complexation by humic acid
692 under neutral pH conditions studied by laser-induced fluorescence spectroscopy.
693 Radiochim. Acta 95, 103-110.
EP

694 Santos, J.S., Teixeira, L.S.G., dos Santos, W.N.L., Lemos, V.A., Godoy, J.M., Ferreira,
695 S.L.C., 2010. Uranium determination using atomic spectrometric techniques: An
C

696 overview. Anal. Chim. Acta 674, 143-156.


AC

697 Shanbhag, P.M., Choppin, G.R., 1981. Binding of uranyl by humic acid. J. Inorg. Nucl.
698 Chem. 43, 3369-3372.

699 Singhal, R.K., Joshi, S.N., Hegde, A.G., 2004. Association of uranium with colloidal and
700 suspended particulate matter in Arabian sea near the west coast of Maharashtra
701 (India). J. Radioanal. Nucl. Chem. 261, 263-267.

702 Stevenson, F.J., 1994 Humus Chemistry. Wiley-VCH, New York.

32
ACCEPTED MANUSCRIPT
703 Stockdale, A., Bryan, N.D., 2013. The influence of natural organic matter on
704 radionuclide mobility under conditions relevant to cementitious disposal of
705 radioactive wastes: A review of direct evidence. Earth-Sci. Rev. 121, 1-17.

706 Sutton, R., Sposito, G., 2005. Molecular structure in soil humic substances: The new
707 view. Environ. Sci. Technol. 39, 9009-9015.

708 Tan, X., Ren, X., Chen, C., Wang, X., 2014. Analytical approaches to the speciation of

PT
709 lanthanides at solid-water interfaces. TrAC Trends Anal. Chem. 61, 107-132.

710 Terashima, M., Fukushima, M., Tanaka, S., 2004. Influence of pH on the surface

RI
711 activity of humic acid: micelle-like aggregate formation and interfacial adsorption.
712 Colloids Surf. A Physicochem. Eng. Asp. 247, 77-83.

SC
713 Timerbaev, A.R., 2007. Recent trends in CE of inorganic ions: From individual to
714 multiple elemental species analysis. Electrophoresis 28, 3420-3435.

U
715 Tipping, E., 1998. Humic ion-binding model VI: An improved description of the
AN
716 interactions of protons and metal ions with humic substances. Aquat. Geochem.
717 4, 3-48.
M

718 Tipping, E., Hurley, M.A., 1992. A unifying model of cation binding by humic
719 substances. Geochim. Cosmochim. Acta Acta 56, 3627-3641.
D

720 Tipping, E., Lofts, S., Sonke, J.E., 2011. Humic Ion-Binding Model VII: A revised
TE

721 parameterisation of cation-binding by humic substances. Environ. Chem. 8, 225-


722 235.
EP

723 Van Loon, L.R., Granacher, S., Harduf, H., 1992. Equilibrium dialysis-ligand exchange:
724 a novel method for determining conditional stability constants of radionuclide-
725 humic acid complexes. Anal. Chim. Acta 268, 235-246.
C

726 Wrobel, K., Sadi, B.B.M., Wrobel, K., Castillo, J.R., Caruso, J.A., 2003. Effect of metal
AC

727 ions on the molecular weight distribution of humic substances derived from
728 municipal compost: Ultrafiltration and size exclusion chromatography with
729 spectrophotometric and inductively coupled plasma-MS detection. Anal. Chem.
730 75, 761-767.

731 Zeh, P., Czerwinski, K.R., Kim, J.I., 1997. Speciation of Uranium in Gorleben
732 Groundwaters. Radiochim. Acta 76, 37-44.

33
ACCEPTED MANUSCRIPT
733 Zhu, B., Ryan, D.K., 2016. Characterizing the interaction between uranyl ion and fulvic
734 acid using regional integration analysis (RIA) and fluorescence quenching. J.
735 Environ. Radioact. 153, 97-103.

736
737
738

PT
RI
U SC
AN
M
D
TE
C EP
AC

34
ACCEPTED MANUSCRIPT
739 Highlights

740 Complexation of Eu and U with NOM is analysed by UF and CE-ICP-MS

741 Influence of high salinity and source of NOM on complexation is determined

742 Source of NOM has a significant influence on complex stability for both metals

PT
743 In contrast to U Eu shows a decreased level of binding to NOM at high salinity

RI
744 Differences are caused by metal speciation or formation of micelle-like aggregates

SC
745

746

747 Keywords
U
AN
748 Ultrafiltration; ICP-MS; Europium; Uranium; NOM; Complex stability constants; Salinity
M

749
D

750
TE

751 Graphical abstract

752
C EP
AC

753

754

35

You might also like