Uranyl retention on quartz—New experimental data and blind prediction using an existing surface...

14
ORIGINAL PAPER Uranyl Retention on Quartz—New Experimental Data and Blind Prediction Using an Existing Surface Complexation Model Florian Huber Johannes Lu ¨ tzenkirchen Received: 3 June 2008 / Accepted: 1 January 2009 / Published online: 27 January 2009 Ó Springer Science+Business Media B.V. 2009 Abstract The adsorption behaviour of uranyl onto seven different samples of quartz was studied in batch experiments. Sea-sand (0.1–0.3 mm), Fil-Pro 12/20 (1–2 mm) and five Min-U-Sil samples with smaller particle sizes (5, 10, 15, 30 and 40 lm) were used. The uptake curves show ‘‘pH adsorption edges’’ in the range ofpH 4–5. A good agreement of the new data with literature data was found when plotting surface-normalised distribution coefficients versus pH. Differences in the adsorption behaviour for pre-treated and untreated sea-sand samples were detectable resulting in a shift of the pH edge to higher pH values after treatment. A literature surface complexation model was applied for blind predictions of the experimental results. The simulations described the experimental observations quite well for the Min-U-Sil samples. For the two coarser quartz samples, the calculated over-predictions were explained by the larger-than-expected measured specific surface area and measurable amounts of associated minerals, for Fil-Pro 12/20 and sea- sand, respectively. Dissolution of the samples was studied as a function of pH. After 5 days, the measured Si concentrations were all higher than equilibrium quartz solubilities, but lower than those of amorphous silica. With increasing pH, dissolved silica increased. This strongly suggests that formation of dissolved uranyl–silicato complexes have to be considered based on measured silica concentrations. Keywords Surface complexation model Adsorption Batch experiments Uranium (VI) 1 Introduction The release of uranium into the environment through anthropogenic activities (e.g. migration from a nuclear waste repository or from mine waste) is a potential threat to eco- and biosphere due to uranium chemo- and radiotoxicity. Various processes occurring in the F. Huber (&) J. Lu ¨tzenkirchen Forschungszentrum Karlsruhe (FZK), Institut fu ¨r Nukleare Entsorgung (INE), Karlsruhe Institute of Technology (KIT), P.O. Box 3640, 76021 Karlsruhe, Germany e-mail: [email protected] 123 Aquat Geochem (2009) 15:443–456 DOI 10.1007/s10498-009-9060-9

Transcript of Uranyl retention on quartz—New experimental data and blind prediction using an existing surface...

ORI GIN AL PA PER

Uranyl Retention on Quartz—New ExperimentalData and Blind Prediction Using an Existing SurfaceComplexation Model

Florian Huber Æ Johannes Lutzenkirchen

Received: 3 June 2008 / Accepted: 1 January 2009 / Published online: 27 January 2009� Springer Science+Business Media B.V. 2009

Abstract The adsorption behaviour of uranyl onto seven different samples of quartz was

studied in batch experiments. Sea-sand (0.1–0.3 mm), Fil-Pro 12/20 (1–2 mm) and five

Min-U-Sil samples with smaller particle sizes (5, 10, 15, 30 and 40 lm) were used. The

uptake curves show ‘‘pH adsorption edges’’ in the range of pH 4–5. A good agreement of

the new data with literature data was found when plotting surface-normalised distribution

coefficients versus pH. Differences in the adsorption behaviour for pre-treated and

untreated sea-sand samples were detectable resulting in a shift of the pH edge to higher pH

values after treatment. A literature surface complexation model was applied for blind

predictions of the experimental results. The simulations described the experimental

observations quite well for the Min-U-Sil samples. For the two coarser quartz samples, the

calculated over-predictions were explained by the larger-than-expected measured specific

surface area and measurable amounts of associated minerals, for Fil-Pro 12/20 and sea-

sand, respectively. Dissolution of the samples was studied as a function of pH. After

5 days, the measured Si concentrations were all higher than equilibrium quartz solubilities,

but lower than those of amorphous silica. With increasing pH, dissolved silica increased.

This strongly suggests that formation of dissolved uranyl–silicato complexes have to be

considered based on measured silica concentrations.

Keywords Surface complexation model � Adsorption � Batch experiments �Uranium (VI)

1 Introduction

The release of uranium into the environment through anthropogenic activities (e.g.

migration from a nuclear waste repository or from mine waste) is a potential threat to eco-

and biosphere due to uranium chemo- and radiotoxicity. Various processes occurring in the

F. Huber (&) � J. LutzenkirchenForschungszentrum Karlsruhe (FZK), Institut fur Nukleare Entsorgung (INE), Karlsruhe Instituteof Technology (KIT), P.O. Box 3640, 76021 Karlsruhe, Germanye-mail: [email protected]

123

Aquat Geochem (2009) 15:443–456DOI 10.1007/s10498-009-9060-9

subsurface environment influence the migration of uranium leading to retardation e.g.

through the adsorption onto matrix material or the precipitation of UO2 under anoxic

conditions caused by abiotically or microbially induced redox processes. Attachment of

dissolved uranium to colloids, which can be transported rapidly over relatively long dis-

tances, may enhance the mobility, although to a minor extent compared to other

radionuclides like e.g. Pu. Hence, it is of particular importance to gain insight into the

processes controlling the mobility and migration of potential contaminants such as uranium

for the development of sound risk-assessment strategies or high-level nuclear waste dis-

posal concepts.

In this study, we are interested in the behaviour of U(VI), i.e. oxidising conditions. The

linear uranyl (UO22?) is the thermodynamically stable form of uranium under oxidising

conditions (Grenthe et al. 1992), which forms strong aqueous complexes with different

ligands. Several studies about the adsorption of uranium(VI) from aqueous solutions onto a

wide range of solid materials (e.g. Fe- and Al-oxides/hydroxides, clay and quartz) under

different chemical conditions (pH, ionic strength, presence of different complexing

ligands, total quantity of uranium, etc) are available. All the studies involving silica or

quartz revealed a more or less strong impact of the solution composition on the adsorption

of uranyl (Kohler et al. 1996; Jung et al. 1999; Artinger et al. 2002; Hongxia and Zuyi

2002; Davis 2001; Pathak and Choppin 2006; Mibus et al. 2007; Gabriel et al. 2001; Zuyi

et al. 2000; Pathak and Choppin 2007a, b; Zuyi et al. 2004). Because of the relatively easy

procedure, batch experiments are the common way to study the adsorption behaviour of

e.g. heavy metals. Upscaling of batch results to contaminant transport in the environment

should be done carefully due to the static nature of the batch technique and the dynamic

nature of the migration process. This concerns, for example, the particle size: in laboratory

experiments most frequently colloidal materials are used, whereas matrix material such as

quartz in the environment usually has much larger grain sizes. In this respect, more realistic

approaches to study the behaviour of contaminants in contact with potential sorbents are

(dynamic) column experiments. These, however, require more effort than batch experi-

ments and are more prone to experimental as well as interpretational problems. Results of

batch experiments might be considered as first information about a given system which in a

second step should be transferred to more realistic conditions by carrying out the respective

column experiments.

Another frequently encountered problem with batch experiments is that in most cases a

full analysis of the aqueous ‘‘equilibrium’’ solution is not available. Thus, for the quartz

system studied here, it is important to take into account sorbent dissolution in order to be

able to evaluate the potential formation of aqueous uranyl–silicato complexes. This phe-

nomenon has previously been largely disregarded. Whenever considered in calculations,

the assumption of thermodynamic equilibrium is made, implying that quartz or silica

would show equilibrium solubilities.

The aims of this study were:

(1) experimental determination of adsorption data from batch experiments for the system

uranyl–quartz in CO2 free systems: such data are missing for the Min-U-Sil samples,

(2) experimental determination of adsorption data from batch experiments for a range of

quartz samples: samples were to span a range of physical properties and the results

were to be compared to available literature data,

(3) comparison of the experimental results with blind predictions using a model

calibrated on batch and column data (in the presence of carbonate, but under those

conditions where carbonate is not expected to affect results significantly): this is a test

444 Aquat Geochem (2009) 15:443–456

123

for an existing model on independent data and different samples of a nominally

identical sorbent,

(4) evaluation of the impact of sorbent dissolution on the speciation of uranyl in solution.

2 Materials and Methods

2.1 Materials

Seven different quartz samples were used: Min-U-Sil 5(lm), 10, 15, 30, 40 (referred to as

MUS hereafter), Fil-Pro 12/20 (*1 mm grain size; referred to as FP) and sea-sand (*0.1–

0.3 mm; referred to as SS). SS was obtained from Merck, while all other samples were

kindly provided by US Silica Company. To guarantee reproducible quartz samples, the

materials were pre-treated to remove impurities. First the quartz samples were heated to

600�C for 48 h to remove organic material, afterwards the samples were washed with 4 M

HNO3 to remove Fe- and Al-oxides/hydroxides and finally washed at least five times with

ultrapure deionised water (Milli-Q) to remove the acid.

Solutions and suspensions were prepared with ultrapure deionised water. Solutions

and suspensions were either acidified or bubbled with purified Argon (see below) before

use, whereby the impact of CO32- is minimised. As background electrolyte, NaNO3

(Merck suprapur) was chosen and adjusted to an ionic strength of 0.01 M to be com-

patible with the experimental work of Kohler et al. (1996). The original uranyl solution

was an acidic (HNO3) 1,000 ppm ICP-MS standard solution, from which U(VI) stock

solutions were prepared. Aliquots were taken from these stock solutions to check the

U(VI) concentrations in these stock solutions via ICP-MS at regular intervals in the

course of this study.

2.2 Solid Characterisation

The solid samples were characterised by X-ray photoelectron spectroscopy (XPS) before

the pre-treatment, and by energy disperse X-ray microanalysis (EDX) and scanning

electron microscopy (SEM) after pre-treatment.

According to XPS analysis, all the MUS samples can be considered as relatively pure

containing [99% quartz with some minor amounts of C, Al and Fe.

XPS results for FP showed amounts of C (*6 at.%), Fe (*0.4 at.%), Al (0.7 at.%) and

Ti (0.1 at.%). Regarding SS specimens, relatively high values of Al (*5.7 at.%) and K

(*1 at.%) were detected, which was later corroborated through EDX analysis and inter-

preted as small amounts of feldspar. The relatively high amount of impurities on the sea-

sand motivated a separate study of uranyl adsorption before and after treatment. EDX

analysis results showed that with the exception of SS the different samples after pre-

treatment are very pure consisting only of Si and O with a ratio of 1:2. The pre-treated SS

still contained minor amounts of Al and K.

Specific surface areas of the different particles were determined by BET N2-adsorption.

The samples were heated to 300�C and degassed. The determination of the surface area

was done via the multiple-point-method at different pressures and a subsequent fit with the

BET isotherm. Table 1 lists the specific surface areas for all the used quartz samples

together with their grain size, brand name and abbreviations. Kohler et al. (1996) reported

a lower specific surface area of 0.33 m2/g for MUS 30. Vuceta (1976) reported a specific

Aquat Geochem (2009) 15:443–456 445

123

surface area of 4.2 m2/g for MUS 5. The comparatively high surface areas for our MUS

samples are likely due to the presence of finer particles which were observed by SEM. In

general, all the MUS samples are very poorly sorted and contained fines. Note that Kohler

et al. (1996) removed the fines prior to their experimental work.

The two coarser quartz samples also exhibit relatively high surface areas based on the

nominal grain size. In contrast to the MUS samples, both SS and FP have no noticeable

amounts of fines. In the case of FP (0.24 m2/g), the ‘‘high’’ surface area is likely due to

pores and heterogeneities on the surface of the particles which can be seen in the SEM

pictures (Lutzenkirchen and Huber 2007). To characterise the micro-porosity of the FP

sample, a complete gas adsorption isotherm was measured. The observed hysteresis is

typical for micro-porous materials. From the full adsorption isotherm, an additional value

for the specific surface area was calculated resulting in a slightly higher value compared

to the standard-BET measurement (0.33 vs. 0.24 m2/g). Specific surface area measure-

ment for pre-treated SS yielded a value of 0.93 m2/g and a slightly lower value of

0.69 m2/g for untreated SS. In all the BET-measurements, care was taken to have suf-

ficient total surface for significant measurements. Pokrovsky et al. (2006) have shown that

it is actually possible to measure very low surface areas of nearly 1-mm grain size quartz

samples.

2.3 Experimental Procedure

All the experiments were done at ambient pressure and room temperature (21�C ± 1). To

minimise the impact of CO2, the samples were bubbled with Ar which was previously

flushed two washing bottles (NaOH and 0.01 M NaNO3, respectively).

To obtain comparable results for the different batch samples, defined amounts of the

different solids were weighed depending on their measured specific surface area, and a

defined amount of uranyl was added to obtain identical uranyl to surface ratios

(0.63 ± 0.1 nm/m2). For Min-U-Sil 40 a value half of that was used. For the sea-sand

sample a variation of the ratio was studied. In all cases, high solid-solution ratios were

employed to minimise wall effects. In addition, blanks were analysed for wall effects but

no measureable U(VI) retention in the test tubes was detected over the whole pH range

studied for the relevant total U(VI) concentrations.

All the experiments were carried out in the following way: known masses of quartz

samples were delivered into test tubes (Sarstedt, 15 ml tubes), and required volumes of

Table 1 Sample names, abbreviations used in the text, grain sizes and specific surface areas of the usedquartz samples

Brand name Abbreviation Grain size(lm)

Specific surfacearea (m2/g)

Min-U-Sil 5 MUS5 5 6.5

Min-U-Sil10 MUS10 10 3.6

Min-U-Sil15 MUS15 15 2.4

Min-U-Sil30 MUS30 30 2.1

Min-U-Sil40 MUS40 40 1.8

Sea-sand (pre-treated) SS 100–300 0.93

Sea-sand (un-treated) SS 100–300 0.69

Fil-Pro 12/20 FP 1,000–2,000 0.24

446 Aquat Geochem (2009) 15:443–456

123

Milli-Q water (bubbled with purified Ar), 1 M NaNO3 solution (to obtain 0.01 M), 0.01 M

HNO3 solution and defined volumes of the U(VI) stock solution were added. This resulted

in a suspension at low pH where U(VI) was not adsorbed. The last step was the adjustment

of the pH under constant flushing with Ar with NaOH (CO2 free) and final flushing with

purified Ar on top before closing the test tubes and sealing them with parafilm. The test

tubes were automatically shaken for 5 days.

After centrifugation (Hettich EBA 8) for 30 min at 7,000 g, the pH was measured in the

supernatant (Orion Ross semi-micro combination electrode). In some cases, the pH was

also measured before centrifugation, but no significant differences were found. The pH

measurement set-up was calibrated using at least five commercial buffers (Merck Titrisol,

kept in refrigerator prior use). All the measurements were done on the potential scale and

under constant flow of Ar over the samples.

After centrifugation a volume of 3 ml of the supernatant was pipetted into polyethylene

tubes (Sarstedt, 5 ml) and analysed by ICP-MS (ICP-MS Elan 6100) for U(VI) and Si. The

amount of adsorbed U(VI) was calculated from the difference between measured U(VI)

and known total initial U(VI).

2.4 Thermodynamic Simulations

To predict the results of batch experiments, a Three-Site Non-Electrostatic surface

complexation Model (TSNEM, Kohler et al. 1996) was applied. This model had been

calibrated on results of column experiments (at pH 4–4.9) and a series of batch results

both with a sample of MUS 30 in 0.01 M NaNO3. The smaller surface area reported by

Kohler et al. (1996) is due to their effort to separate the fines since they used the sample

for column experiments and had to avoid losses of small particles from the column. The

development of the surface complexation model by Kohler et al. was a step-wise process.

To begin with, a one-site approach was tested followed by a two-site approach, which

both failed to describe the experimental results. The use of an additional site resulted in

good agreement with experiment. Table 2 summarises the complete set of model

parameters for the aqueous speciation and the surface complexation reactions. The set of

constants has to be used in a self-consistent way, so that it should be limited to the same

ionic medium used by Kohler et al. (1996). In this model, uranyl–silica complexes are not

considered. We performed separate speciation calculations on this issue with a complete

set of stability constants from Nuclear Energy Agency Thermodynamic Database,

including associated auxiliary data (NEA 2003). It should be noted that the applied

model, which was developed by Kohler et al. (1996) is purely based on best-fit param-

eters. In this study, no attempt was made to adjust any of these parameters. We also stress

that we do not attempt to associate any mechanistic meaning to the adsorption model,

neither with respect to the nature of the sites nor with respect to the postulated surface

species. We refer to Goyne et al. (2002) for a discussion of the issue of site densities.

Given that the full model by Kohler et al. (1996) is accepted as such, one crucial particle

property is the specific surface area. As discussed above, there is some uncertainty about

the measured surface areas. For the Fil-Pro sample this has been discussed elsewhere

(Lutzenkirchen and Huber 2007). Since great care had been taken to measure the areas

with sufficient surface, we accept the measured values. Making use of these values, the

respective site concentrations can be calculated for the modelling using the site densities

in Table 2.

The multi-component speciation code ECOSAT 4.8 (Keizer and van Riemsdijk 1998)

was applied for all the modelling calculations.

Aquat Geochem (2009) 15:443–456 447

123

3 Results and Discussion

3.1 Batch Adsorption Experiments

Figure 1 displays the results of the batch experiments for the different samples as per-

centage adsorbed U(VI) versus pH. At low pH (\4, quartz is uncharged under this

condition), adsorption is weak but increases in the range of pH 4–5 sharply (from *10% to

80–90%). From pH 5 to 10, adsorption is constant at almost 100% showing that there is no

influence of CO2. The presence of CO2 would cause a decrease of adsorption with

increasing pH because of the formation of dissolved uranyl–carbonato complexes.

There is no significant influence of the ‘‘nominal’’ grain size detectable for the MUS

samples, which is not expected due to the high amount of fines controlling the behaviour or

due to identical adsorption properties of the different particles. Also for the two other pre-

treated samples the pH-edge is between pH 4 and 5. Thus, the results for all the pre-treated

samples agree (pH edge in the same pH range and no influence of CO2).

The results of the batch experiments for the untreated sea-sand differ from those

described above. Adsorption starts at lower pH values. Independent results of a column

experiment with the untreated sea-sand sample at pH 4 support this observation since no

breakthrough of uranyl was detected during the sampling period (9 h) while column

Table 2 Thermodynamic data for aqueous phase reactions and surface complexation reactions (Kohleret al. 1996)

Species Reaction Log K (I = 0)

[UO2OH]? H2O(l) ? UO22? $ [UO2OH]? ? H? -5.20

UO2(OH)2(aq) 2H2O(l) ? UO22? $ UO2(OH)2 ? 2H? -12.00

[UO2(OH)3]- 3H2O(l) ? UO22? $ [UO2(OH)3]- ? 3H? -20.00

[UO2(OH)4]2- 4H2O(l) ? UO22? $ [UO2(OH)4]2- ? 4H? -32.00

[(UO2)2(OH)]3? H2O(l) ? 2UO22? $ [(UO2)2OH]3? ? H? -2.80

[(UO2)2(OH)2]2? 2H2O(l) ? 2UO22? $ [(UO2)2(OH)2]2? ? 2H? -5.63

[(UO2)3(OH)4]2? 4H2O(l) ? 3UO22? $ [(UO2)3(OH)4]2? ? 4H? -11.90

[(UO2)3(OH)5]? 5H2O(l) ? 3UO22? $ [(UO2)3(OH)5]? ? 5H? -15.56

[(UO2)3(OH)7]- 7H2O(l) ? 3UO22? $ [(UO2)3(OH)7]- ? 7H? -31.20

[(UO2)4(OH)7]? 7H2O(l) ? 4UO22? $ [(UO2)4(OH)7]? ? 7H? -21.90

[UO2NO3]? NO3- ? UO2

2? $ [UO2NO3]? 0.26

Surface complexation reactions Log K (I = 0.01) fa

S1OH ? UO22? ? H2O $ S1OUO2OH ? 2H? -5.04 0.9748

S2OH ? UO22? ? H2O $ S2OUO2OH ? 2H? -3.56 0.0250

S2OH ? UO22? $ S2OUO2

? ? H? 0.642 0.0250

S3OH ? UO22? ? H2O $ S3OUO2OH ? 2H? -1.15 0.0002

Total site density = 0.35 lmol sites/g of quartz (1.06 lmol sites/m2)a f = fraction of the total surface site density for each site type

Fig. 1 Results of the batch experiments (open circles) and corresponding calculations with ECOSAT (solidlines) for a Min-U-Sil 5, b Min-U-Sil 10, c Min-U-Sil 15 and d Min-U-Sil 30, e Min-U-Sil 40, f Fil-Pro 12/20,g sea-sand untreated and h sea-sand pre-treated

c

448 Aquat Geochem (2009) 15:443–456

123

Aquat Geochem (2009) 15:443–456 449

123

experiments under the same conditions with the pre-treated FP sample showed a break-

through. Preliminary streaming potential measurements on the two samples suggest that

the treated sample is behaving ‘‘quartz’’-like, since no isoelectric point was found for

pH [ 4 and a typical levelling off found for some silica samples occurred at pH \ 4 at

near-zero zeta potential, whereas the treated sample had a sharp isoelectric point of about

4.8, which is typical for coated silica samples. These results would suggest that the treated

sample macroscopically behaves like quartz/silica.

3.2 Dissolved Silica Concentrations

Measured silica concentrations for 5-day equilibration periods are shown in Fig. 2 for all

samples. In general, the lowest solubilities were obtained for acidic conditions at

*0.1 mM Si rising to *1 mM Si for high pH values. As expected and in agreement with

the literature (Iler 1979), finer quartz samples (MUS) showed higher aqueous Si concen-

trations than the coarser samples. Figure 2b also includes measured dissolved Si

concentrations for MUS 10 in the presence of U(VI). They are as expected in the range of

data in the absence of U(VI). The measured concentrations can be safely used to evaluate

the impact of the dissolution on the speciation of U(VI) in solution.

It is surprising that aqueous silica concentrations are higher than quartz equilibrium

concentrations. This is discussed in more detail in the subsequent section.

3.3 Results of the Thermodynamic Simulations

Figure 1 includes the results of blind predictions by the TSNEM. The model predicts the

new experimental data for the MUS quartz samples quite well, which may have been

expected due to the calibration of the model with MUS 30 by Kohler et al. (1996) although

in a sample without fines. Predictions of the behaviour of the pre-treated coarser quartz

samples were less successful. Surprisingly, the data for the untreated sea-sand sample are

nicely predicted. This indicates the importance of checking the surface composition of

such samples by spectroscopic methods. Without the XPS or EDX results, the untreated

sea-sand would have been handled as a quartz sample, a model would have been found to

be quite successful (Fig. 1g), and the underlying problems would not have been discov-

ered. For the pre-treated SS sample, the tendency with increasing solid content is captured,

Fig. 2 a Experimentally determined solubilities for all the quartz materials used in the batch experiments(symbols; I = 0.01 M) and calculated solubilities for amorphous (solid line) and crystalline (dashed line)quartz; b Comparison of the quartz solubility for all the Min-U-Sil sorts in the presence of uranyl (crosses),and for Min-U-Sil 10 (black circles) in the absence of uranyl

450 Aquat Geochem (2009) 15:443–456

123

but the pH-range of the predicted pH-edge is too low. The reason for the failure in the

predictions for the pre-treated coarse samples might be uncertainties in the surface area

measurements. For the FP sample, the lower value for the surface area is used. A slight

increase in the value (to that found from the full isotherm) shifts the curve to lower pH. The

variation in the available values for the specific surface areas of these samples indicates

that the uncertainty in this important property is high. Moreover, for the sea-sand samples,

the increase in specific surface area after pre-treatment was somewhat unexpected. The low

values involve some uncertainty although the amount of surface in the measurement set-up

was deemed sufficient. The data can be exactly described using specific surface areas

between calculated geometrical and measured values. The predictions using the measured

values lie within one pH unit of the experimental results.

Calculated equilibrium solubilities for amorphous and crystalline quartz are compared

with measured concentrations in Fig. 2a. The experimentally determined concentrations

are intermediate between the two theoretical curves. All the data are consistently higher

than the equilibrium quartz solubility, but dissolution has not reached equilibrium with

amorphous silica. The results suggest that the surface of quartz transformed to a surface

corresponding to amorphous silica in agreement with earlier suggestions that the high

charging of some quartz sample indicates a gel-like layer (Riese 1982). This experimental

results suggest that it is necessary to measure silica concentrations in solutions instead of

relying on equilibrium solubilities when estimating the effect of dissolved silica (here, this

is important for uranyl–silicato complexes). Figure 3 shows the results of speciation cal-

culations for the aqueous uranyl–silica system for equilibrium with respect to amorphous

silica and quartz. Experimental results are between quartz and amorphous silica. The fact

that aqueous silica is above the quartz equilibrium concentrations may have different

reasons. One would be that pre-treatment like grinding causes an amorphisation of the

surface, which then results in interfacial behaviour of amorphous silica. Another possibility

to explain this effect would be the often-suggested formation of the gel-like layer when

quartz is in contact with water. It should be noted that there is no general agreement in the

scientific literature about the existence of such a gel layer or an amorphous surface (see

Ball et al. 1997 and references therein). The calculations indicate that a refined model

should consider the formation of uranyl–silicato complexes, which appear or dominate in

the mildly acidic pH-range.

Fig. 3 Calculated speciation of uranyl in the presence of dissolved silica for a equilibrium with quartz,b equilibrium with amorphous silica. Only major species are shown. All the thermodynamic data areadopted from the most recent NEA compilation (Vol. 5)

Aquat Geochem (2009) 15:443–456 451

123

3.4 Comparison of the Experimental Results with Literature Data

Figure 4 shows a comparison of our adsorption data for the seven pre-treated samples with

literature data for quartz and amorphous silica. Data are shown as distribution coefficients

calculated with the surface area of the respective sorbent to include the effects of specific

surface area. It can be expressed as follows:

Kd ¼CU

½U� ¼ðTOTU� ½U�Þ

½U� =/m ¼ TOTU

½U� � 1

� �=/m;

where

Kd = Distribution coefficient in l/m2,

CU = Surface excess of U(VI) in mol/m2,

[U] = Dissolved U(VI) concentration in mol/l,

TOTU = Total U(VI) concentration in mol/l,

/ = Surface area in m2/g,

m = Solid concentration in g/l.

It is important to note that the uncertainty in the distribution coefficient close to 100%

adsorption is typically getting high. Moreover, at low pH where most of the uranyl is

remaining in solution, the relative error (difference between two big numbers) becomes

important.

Literature data in Fig. 4 may be in the presence or absence of carbonate. For the low pH

range, carbonate complexation can be neglected. Since carbonate does not adsorb to silica

in significant amounts as was shown experimentally by Vuceta (1976), the surface charge

of pre-treated samples (i.e. largely or entirely Fe-free based on XPS results) is not affected

by carbonate. The results of our study agree fairly well with published results. Deviations

for data at high pH values ([8) are due to the presence of carbonate in some literature data

Fig. 4 Comparison of literature data for uranyl adsorption with silica polymorphs. Data are represented as asurface specific distribution coefficient (Kd/[l/m2]). Only data for ionic strength of 0.01 M are used

452 Aquat Geochem (2009) 15:443–456

123

resulting in a decrease of log Kd due to uranyl-carbonato complexation. Interestingly, in

this plot the SS (treated) and FP samples fit well. Instead, the data for untreated SS sample

are somewhat lower than the bulk of the data. This would suggest that the success of the

model for the untreated SS sample is accidental and that the model should be considered

with care when transferring it from one solid to another. Note that the literature Kd values

span about one pH unit. So if the experimental data were reliable, then a deviation of one

pH unit can be tolerated. However, such a deviation can, in particular within a model

prediction, result in significant over- or underprediction of uranyl retention. From this point

of view such a deviation is undesirable.

3.5 Surface Speciation

Figure 5 shows the surface speciation predicted by the applied surface complexation model

(TSNEM). The figures also include the amount of uranyl relative to the surface area in the

system. The inherent site densities of the three surface sites of the model are a fraction of

1.06 lmol/m2, namely 97.48%, 2.5% and 0.02% for sites 1, 2 and 3, respectively. For none

of the conditions, the high-affinity site is present in sufficient amount to fix all of the uranyl

present. For the experiments with the high uranyl-to-surface area ratios, a common

sequence in the surface speciation occurs: at low pH values (\4) where the overall

adsorption is low (max. *20%) and the free uranyl ion is the major aqueous species,

S2OUO2? is the dominating surface species. This is somewhat unexpected since S3 is the

high-affinity site, but since only a hydrolysed uranyl surface species was postulated on that

site by Kohler et al. (1996), the predominant adsorption on S2 is caused by the pH-

dependence. S2OUO2? rarely reaches up to more than about 10%, and is negligible at

pH [ 5. Above pH 4, S1OUO2OH becomes the major surface species over the whole

remaining pH range ([4–10) at about 40% except for pre-treated SS at the lowest uranyl-

to-surface ratio where S3OUO2OH dominates (Fig. 5h). S3OUO2OH reaches values up to

*20% except for the uranyl-to-surface area ratios below 0.3 nmol m-2 can be understood

from comparing this ratio to the site density. S2OUO2OH usually plays an intermediate

role (contributing about *30%) and is almost dominated by S1OUO2OH. This can be

understood by considering the product of site density and adsorption constant, which is

higher for S1 than for S2 for the hydrolysed surface species by a factor of about 1.3, which

reflects fairly well the calculated differences. In summary, saturation of the strong sites

occurs even for the low uranyl concentrations and the sites with ‘‘weaker’’ affinity dominate,

whereas at higher pH site 1 dominates site 2 although site 2 is not saturated. The role of site 2

as a ‘‘more’’ reactive site is only in the postulated S2OUO2? species, which always domi-

nates at low pH. For most of our batch experiment conditions, the model predicts similar

contributions from all the four surface species at the onset of adsorption (around pH 4) where

most column experiments were done, but for the latter the solid-to-liquid ratio is even higher,

so that it is expected that the high-affinity site will be more important.

4 Summary and Conclusions

The adsorption of uranyl onto seven different samples of quartz was studied by static batch

experiments. Thermodynamic predictions were carried out to simulate the experimental

results using the complete surface complexation model calibrated on adsorption of U(VI)

onto Min-U-Sil 30 as published by Kohler et al. (1996).

Aquat Geochem (2009) 15:443–456 453

123

454 Aquat Geochem (2009) 15:443–456

123

The applied SCM describes well the results of the batch experiments for the MUS

samples given that the experimental data were obtained in a different laboratory and that

the samples used in this study included significant amounts of fine particles.

Comparison of our results with literature data gives confidence in the measurements

under various experimental conditions (in terms of procedures such as solid liquid sepa-

ration or uranyl analysis). In contrast to the model predictions, the measured Kd values for

the pre-treated samples are consistent with literature data and with each other while the

data for the untreated sample deviates.

Prediction of the data for the coarser samples is satisfactory but off by a pH unit. This

may be due to uncertainty in the measured specific surface area which is higher than

expected based on grain size. For the untreated sea-sand sample, a perfect prediction was

obtained although this should not be the case due to the impurities on this sample.

Experimentally determined silica concentrations for the seven samples are in the range

from 0.1 mM to about 1 mM. No solubility equilibrium was attained within 5 days as

inferred from comparison of the measured values with calculated equilibrium solubilities

for amorphous silica and quartz. Measured values of dissolved silica concentrations are

higher than equilibrium quartz solubilities. This implies some amorphisation of the sur-

faces either due to grinding of the raw material or due to formation of gel-like layers in

contact with water. The solubility measurements along with speciation calculations for

uranyl indicate that uranyl–silicato complexes need to be considered and that silica dis-

solution needs to be measured if ambiguity with respect to the total dissolved silica in the

aqueous speciation scheme is to be avoided.

Surface speciation calculations within the model used show that depending on the

uranyl-to-surface ratio, the high-affinity site is often saturated in batch experiments. The

situation is expected to be different in column experiments, where that ratio is very low.

The use of the unhydrolysed surface species with an intermediate site density causes

adsorption on that site only at the low pH range.

Acknowledgements We thank A. Kaufmann and F. Geyer for ICP-MS analysis, D. Schild for XPSmeasurements, T. Kisely for BET analysis (all of them from INE) and T. Wiss and H. Thiele (both JRC-ITU) for EDX and SEM investigations. A thorough reading of the manuscript by Dr. B. Kienzler (INE) isgratefully acknowledged.

References

Artinger R, Rabung T, Kim JI, Sachs S, Schmeide K, Heise KH, Bernhard G, Nitsche H (2002) Humiccolloid-borne migration of uranium in sand columns. J Contam Hydrol 58:1–12. doi:10.1016/S0169-7722(02)00032-3

Ball MC, Brown DS, Perkins LJ (1997) Diffuse layers on the surface of mineral quartz. J Mater Chem7:365–368. doi:10.1039/a603570a

Davis JA (ed) (2001) Surface complexation modeling of uranium(VI) adsorption on natural mineralassemblages, Report NUREG/CR-6708, US Nuclear Regulatory Commission W D.C.

Gabriel U, Charlet L, Schlapfer CW (2001) In situ speciation of uranium (VI) at the silica–water interface: acombined TRLFS and surface complexation study. In: Hellmann R, Wood SA (eds) Water–rockinteractions, ore deposits, and environmental geochemistry: a tribute to David A. Crerar. GeochemicalSociety, St. Louis, Spec. Publ. No. 7

Fig. 5 Surface species distribution of the surface complexation model (TSNEM) for the correspondingbatch experiments. a MUS 5, b MUS 10, c MUS 15, d MUS 30, e MUS 40, f Fil-Pro 12/20, g sea-sand(treated) at 0.18 kg/l, and h sea-sand (treated) at 0.6 kg/l

b

Aquat Geochem (2009) 15:443–456 455

123

Goyne KW, Zimmerman AR, Newalkar BL, Komarneni S, Brantley SL, Chorover J (2002) Surface chargeof variable porosity Al2O3(s) and SiO2(s) adsorbents. J Porous Mater 9:243–256. doi:10.1023/A:1021631827398

Grenthe I, Fuger J, Konings RJM, Lemire RJ, Muller AB, Nguyen-Trung C, Wanner H (1992) Chemicalthermodynamics of uranium. Elsevier Science, Amsterdam

Hongxia Z, Zuyi T (2002) Sorption of uranyl ions on silica: effects of contact time, pH, ionic strength,concentration and phosphate. J Radioanal Nucl Chem 254:103–107. doi:10.1023/A:1020801816866

Iler K (1979) The chemistry of silica: solubility, polymerization, colloid and surface properties and bio-chemistry. Wiley, New York

Jung J, Hyun SP, Lee JK, Cho YH, Hahn PS (1999) Adsorption of UO22? on natural composite materials.

J Radioanal Nucl Chem 242:405–412. doi:10.1007/BF02345570Keizer MG, van Riemsdijk WH (1998) ECOSAT: technical report of the department of soil science and

plant nutrition, Wageningen agricultural. University of Wageningen, The NetherlandsKohler M, Curtis GP, Kent DB, Davis JA (1996) Experimental investigations and modelling of uranium(VI)

transport under variable chemical conditions. Water Resour Res 32:3539–3551Lutzenkirchen L, Huber F (2007) Heterogeneities in adsorption of aqueous solutions—an example of the

effect of surface coverage. Adsorpt Sci Technol 25:503–516Mibus J, Sachs S, Pfingsten W, Nebelung C, Bernhard G (2007) Migration of uranium (IV)/(VI) in the

presence of humic acids in quartz sand: a laboratory column study. J Contam Hydrol 89:199–217. doi:10.1016/j.jconhyd.2006.08.005

NEA (2003) Vol. 5. Update on the chemical thermodynamics of uranium, neptunium, plutonium, americiumand technetium. OECD Nuclear Energy Agency, Paris

Pathak NP, Choppin GR (2006) Kinetics and thermodynamics of uranium(VI) sorption on hydrous silica.Radiochim Acta 95:507–512. doi:10.1524/ract.2007.95.9.507

Pathak NP, Choppin GR (2007a) Sorption of uranyl on hydrous silica: effects of ionic strength and eth-ylenediaminetetraacetic acid (EDTA). J Radioanal Nucl Chem 272:37–43. doi:10.1007/s10967-006-6783-6

Pathak NP, Choppin GR (2007b) Sorption of uranyl(VI) cations on suspended silicate: effects of N-donorligands, carboxylic acids, organic solvents, and metal ions. Radiochim Acta 95:245–250. doi:10.1524/ract.2007.95.5.245

Pokrovsky OS, Golubev SV, Mielczarski JA (2006) Kinetic evidences of the existence of positively chargedspecies at the quartz-aqueous solution interface. J Colloid Interface Sci 296:189–194. doi:10.1016/j.jcis.2005.09.001

Riese AC (1982) Adsorption of radium and thorium onto quartz and kaolinite: A comparison of solution/surface equilibria models. Ph. D. dissertation, Colorado School of Mines

Vuceta J (1976) Adsorption of Pb(II) and Cu(II) on a-quartz from aqueous solutions: influence of pH, ionicstrength and complexing ligands. Ph. D. dissertation, Calif. Inst. of Technol., Pasadena

Zuyi T, Taiwei C, Jinzhou D, XiongXin D, Yingjie G (2000) Effect of fulvic acids on sorption of U(VI), Zn,Yb, I and Se(IV) onto oxides of aluminium, iron and silicon. Appl Geochem 15:133–139. doi:10.1016/S0883-2927(99)00042-6

Zuyi T, Taiwei C, Weijuan L, Jinzhou D, Xiongxin D, Yingjie G (2004) Cation adsorption of NpO2?,UO2

2?, Zn2?, Sr2?, Yb3?, and Am3? onto Oxides of Al, Si and Fe from aqueous solution: ionicstrength effects. Colloid Surf A 242:39–45. doi:10.1016/j.colsurfa.2004.04.044

456 Aquat Geochem (2009) 15:443–456

123