You are on page 1of 10

Science of the Total Environment 610–611 (2018) 511–520

Contents lists available at ScienceDirect

Science of the Total Environment

journal homepage: www.elsevier.com/locate/scitotenv

Cesium sorption reversibility and kinetics on illite, montmorillonite,


and kaolinite
Chad B. Durrant a,b,⁎, James D. Begg a, Annie B. Kersting a, Mavrik Zavarin a
a
Glenn T. Seaborg Institute, Physical & Life Sciences, Lawrence Livermore National Laboratory, Livermore, CA 94550, United States
b
The Pennsylvania State University, 101 Brezeale Nuclear Reactor, University Park, PA 16802, United States

H I G H L I G H T S G R A P H I C A L A B S T R A C T

• Permeable membrane used to examine


desorption of 137Cs in the presence of
two minerals.
• Ternary experiments carried out of over
N 500 days.
• Cs is reversibly sorbed to kaolinite and
montmorillonite.
• Ternary experiments indicate a slow ki-
netic effect on Cs desorption from illite.

a r t i c l e i n f o a b s t r a c t

Article history: Understanding sorption and desorption processes is essential to predicting the mobility of radionuclides in the en-
Received 19 June 2017 vironment. We investigate adsorption/desorption of cesium in both binary (Cs + one mineral) and ternary (Cs
Received in revised form 12 August 2017 + two minerals) experiments to study component additivity and sorption reversibility over long time periods
Accepted 12 August 2017
(500 days). Binary Cs sorption experiments were performed with illite, montmorillonite, and kaolinite in a
Available online 17 August 2017
5 mM NaCl/0.7 mM NaHCO3 solution (pH 8) and Cs concentration range of 10–3 to 10–11 M. The binary sorption
Editor: Jay Gan experiments were followed by batch desorption experiments. The sorption behavior was modeled with the
FIT4FD code and the results used to predict desorption behavior. Sorption to montmorillonite and kaolinite was
Keywords: linear over the entire concentration range but sorption to illite was non-linear, indicating the presence of multiple
Ternary system sorption sites. Based on the 14 day batch desorption data, cesium sorption appeared irreversible at high surface
Desorption loadings in the case of illite but reversible at all concentrations for montmorillonite and kaolinite. A novel experi-
Irreversibility mental approach, using a dialysis membrane, was adopted in the ternary experiments, allowing investigation of
Permeable membrane the effect of a second mineral on Cs desorption from the original mineral. Cs was first sorbed to illite, montmoril-
lonite or kaolinite, then a 3.5–5 kDalton Float-A-Lyzer® dialysis bag with 0.3 g of illite was introduced to each ex-
periment inducing desorption. Nearly complete Cs desorption from kaolinite and montmorillonite was observed
over the experiment, consistent with our equilibrium model, indicating complete Cs desorption from these min-
erals. Results from the long-term ternary experiments show significantly greater Cs desorption compared to the
binary desorption experiments. Approximately ~45% of Cs desorbed from illite. However, our equilibrium model
predicted ~65% desorption. Importantly, the data imply that in some cases, slow desorption kinetics rather than
permanent fixation may play an important role in apparent irreversible Cs sorption.
© 2017 Elsevier B.V. All rights reserved.

⁎ Corresponding author at: Glenn T. Seaborg Institute, Physical & Life Sciences, Lawrence Livermore National Laboratory, Livermore, CA 94550, United States.
E-mail address: durrant1@llnl.gov (C.B. Durrant).

http://dx.doi.org/10.1016/j.scitotenv.2017.08.122
0048-9697/© 2017 Elsevier B.V. All rights reserved.
512 C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520

1. Introduction is readily desorbed from the surface of kaolinite (Shahwan et al., 1999;
Erten et al., 1988). However, the behavior with illite is more complex
Significant quantities of radioactive cesium (Cs) have been released with apparent irreversible behavior observed in some cases. This irre-
to the environment. These releases have occurred as a result of nuclear versibility has been attributed to sorption to high affinity FES, and
weapons testing (Klement, 1965), nuclear accidents at Chernobyl, their subsequent collapse that prevents access to the sites (Sawhney,
Russia and Fukushima, Japan (Castrillejo et al., 2016; Steinhauser et al., 1972). High concentrations of Cs have been shown to exacerbate inter-
2015), and leaks from high level waste storage sites such as Hanford, layer collapse leading to apparent irreversible sorption behavior. A time
USA (Zachara et al., 2002) and Mayak, Russia (Novikov et al., 1998). dependence on the amount of reversibly sorbed Cs has been attributed
The persistence of 137Cs (t1/2 30.1 y) in the environment can pose sub- to diffusion of Cs from FES to the interlayer sites of illite (Comans and
stantial health concerns. The U.S. EPA regulates the amount of beta Hockley, 1992). However, more recently a time dependent increase in
emitting radionuclides in drinking water such that the equivalent dose sorption irreversibility has been observed at timescales too short to be
is b4 mrem per year (United States Environment Protection Agency, explained by diffusion processes (De Koning and Comans, 2004).
2009). This dose corresponds to a maximum 137Cs concentration of Most sorption/desorption studies are conducted as single mineral
7.4 Bq/L or 1.7 × 10−14 M. Thus, 137Cs is considered a drinking water binary batch experiments over short time periods (days to weeks).
hazard even at trace concentrations. The environmental contamination However, in the environment, Cs will interact with multiple minerals si-
and potential health risks associated with 137Cs necessitate understand- multaneously and over long time scales. This presence of multiple min-
ing its behavior in the environment. Understanding this behavior may erals is especially important to consider in the case of colloid-facilitated
also guide remediation strategies for Cs clean-up in the environment transport of Cs and previous column experiments have demonstrated
and help to design strategies for the safe storage of radionuclides pres- that Cs can be desorbed from mobile colloids and adsorbed on the im-
ent in spent nuclear fuel. mobile host minerals (Zhuang et al., 2003; Turner et al., 2006). Previous-
Sorption and desorption processes play a large role in controlling the ly, De Koning and Comans (2004) examined Cs sorption reversibility on
fate of Cs in the environment. Sorption to immobile soils or sediments illite with ammonium hexacyanoferrate (HCF) which has a high selec-
will limit the transport of Cs in solution, while sorption to colloids (de- tivity and capacity for Cs. The high sorption affinity of HCF ensures
fined as b1 μm particulates) may enhance mobility. If Cs is sorbed to that solution Cs concentrations are low and drives desorption of Cs
minerals, then its desorption behavior will affect its bioavailability and from illite. In our current work, we employ a similar approach but use
in the case of mineral-colloid association, the extent of transport. illite to drive desorption. This novel approach has the advantage of
Due to their high sorption capacity and their ubiquity in nature, alu- being more representative of natural environments.
minosilicate clays are particularly important in controlling Cs behavior In this work, we evaluate Cs sorption to illite, montmorillonite, and
in natural systems. Cs sorption reversibility on clay minerals with kaolinite over a wide range of initial Cs concentration that extends to
high, intermediate, and low Cs affinities (illite, montmorillonite, and ka- low, environmentally relevant concentrations. We perform binary Cs
olinite, respectively) was the focus of this study. Illite is a non- sorption and desorption experiments with illite, montmorillonite, and
expanding 2:1 (2 tetrahedral silica layers sandwiching an octahedral kaolinite across an initial Cs concentration range of 10−3 to 10−11 M.
alumina layer, TOT) aluminosilicate clay. Cesium sorption to illite has We then examine long-term Cs sorption/desorption behavior in two-
been described as a cation exchange process on either the basal/planar, mineral ternary experiments. By keeping the two mineral phases phys-
edge, or interlayer sites (Staunton and Roubaud, 1997). Weathering ically separated using a Specta/Por® Float-A-Lyzer® dialysis bag, we are
near the edges of the illite expands the interlayer producing wedge or able to examine the contribution of each mineral to the overall sorption
frayed edge sites (FES) (Jackson, 1963). These FES sites are highly selec- and evaluate the competitive sorption and desorption of Cs in ternary
tive for Cs (Rich and Black, 1964; Bolt et al., 1963; Brouwer et al., 1983). experiments over a period of nearly 17 months. This method provides
As a result, while illite has much lower cation exchange capacity (CEC) a means to evaluate component additivity (Davis et al., 1998), revers-
than montmorillonite, at Cs concentrations below 10− 7 M, illite will ibility, and desorption kinetics in the presence of clay minerals over en-
sorb more Cs than montmorillonite per unit mass (Sawhney, 1970; vironmentally relevant timescales.
Komarneni and Roy, 1980).
Montmorillonite is a 2:1 smectite clay with an expanding interlayer 2. Materials and methods
(Jackson, 1963) that allows for increased interaction between cations in
solution and those in the clay interlayer. The high CEC of montmorillon- 2.1. Materials
ite leads to its high affinity for Cs (Sawhney, 1970; Sawhney, 1972; Hsu
and Chang, 1995; Komarneni and Roy, 1979). All solutions were prepared using ACS grade chemicals without fur-
Kaolinite is a 1:1 aluminosilicate non-expanding clay. Sorption of Cs ther purification and ultrapure water (Milli-Q Gradient System,
to kaolinite occurs by cation exchange at permanently charged sites on N18 MΩ·cm). Fithian illite from Rochester, NY (Ward's Science), SWy-
the basal surfaces of the mineral (Swartzen-Allen and Matijevic, 1974; 1 montmorillonite (Clay Minerals Repository), and KGa-1b kaolinite
Kim et al., 1996; Kim and Cygan, 1996; Ohnuki and Kozai, 2013; (Clay Minerals Repository) were used in this study. Preparation of
Reinoso-Maset and Ly, 2014). Due to its low CEC and absence of FES homoionic SWy-1 Na-montmorillonite has been previously reported
sites, kaolinite sorbs Cs less effectively than montmorillonite or illite. (Zavarin et al., 2012) and the same method was also used for the prep-
In terms of desorption behavior, research has focused on whether Cs aration of illite and kaolinite. Briefly, the minerals were prepared by
is irreversibly sorbed by clay minerals. In this context we mean that mixing 100 g/L of the mineral in 0.001 M HCl for 30 min in order to re-
sorption is considered to be reversible if Kdsorption ≈ Kddesorption move soluble salts and impurities. Next, 0.03 M H2O2 was added for
(Missana et al., 2004). However, determining true irreversibility can 30 min to minimize the oxidative and reductive potential of any remain-
be difficult because sorption and desorption behavior can be affected ing impurities as well as removing trace organic material. The clay sus-
by the solution composition, kinetics, Cs concentration; and the pres- pensions were then centrifuged for 6 h at 2500g and the supernatant
ence of multiple sorption sites with differing Cs affinities (Comans (calculated to contain particles b50 nm diameter) was discarded. The
et al., 1991; Bostick et al., 2002; Missana et al., 2004; Iijima et al., 2010). wet solids were transferred to 6000–8000 Da molecular weight cut off
In the case of montmorillonite and montmorillonite dominated (MWCO) dialysis bags and suspended in a 0.01 M NaCl electrolyte solu-
clays, sorption is believed to be reversible (Missana et al., 2004; Iijima tion and dialyzed for seven days to homoionize the minerals. It is possi-
et al., 2010), however recent work has shown that some fixation of in- ble that there still exists some intrinsic Cs and this may have an effect on
trinsic Cs may be achieved by the addition of divalent cations before de- the sorption processes. However, the effect is likely to be small and no
sorption (Fukushi et al., 2014; Fukushi and Fukiage, 2015). Similarly, Cs further consideration is given in this work. The minerals were extruded
C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520 513

from the dialysis bags into Milli-Q H2O. The suspension was then centri- orbital shaker in the dark and allowed to equilibrate for two weeks.
fuged (sedimentation for 1 h was used for illite and kaolinite) at 100g After the equilibration period, 5 mL of a 60 g/L illite suspension (0.3 g
for 5 min and the supernatant containing particles b 2 μm was discarded. total) was loaded into 3500–5000 MWCO Da Float-A-Lyzers (see graph-
The remaining solids (50 nm b − b 2 μm) were dried at 60 °C. A portion ical abstract). The Float-A-Lyzers with the illite suspensions were then
of the dried minerals was resuspended in a 5 mM NaCl and 0.7 mM inserted into each of the centrifuge tubes. The centrifuge tubes were
NaHCO3 solution (pH 8) with an ionic strength of 0.0057 M as a synthet- then returned to the dark on the orbital shaker. The suspensions outside
ic groundwater. The solid:solution ratio of the mineral suspension was the Float-A-Lyzers were periodically sampled over the duration of the
1 g/L unless otherwise noted. The mineral suspensions were allowed experiment, a period of nearly 17 months. Samples were taken as fol-
to equilibrate over several days. lows: a 1 mL aliquot of the suspension was taken and counted by LSC.
Then a 1.2 mL aliquot of the suspension was removed and centrifuged
2.2. Binary sorption experiments for 2 h at 3600 g to achieve a particle size cutoff of 50 nm. A 1 mL aliquot
of the supernatant was removed and counted by LSC to determine the
Ten mL of each mineral suspension was added to 15 mL polyethyl- aqueous Cs concentration. Separate experiments were performed to
ene centrifuge tubes. A total of 17 different initial Cs concentrations verify that the suspensions were sufficiently stable to allow for quanti-
ranging from 10−3 to 10−11 M were prepared. A combination of stable tative and reproducible measurement of the total Cs activity in the sus-
CsCl and 137Cs was added to samples with Cs concentrations of 10−9 M pensions. Characterization of sample blanks indicated that Cs sorption
and higher and only 137Cs for concentrations lower than 10−9 M. The to the membrane or walls of the centrifuge tube was negligible and
centrifuge tubes were then placed on an orbital shaker at 0.4 g at the Float-A-Lyzer pore size allowed for exchange of aqueous Cs (SI Fig.
room temperature and kept in the dark for the duration of the experi- A1). Therefore, it can be assumed that the decrease in total Cs outside
ment. The samples were shaken continuously for at least two weeks the Float-A-Lyzer corresponds to a proportional increase of the total
to allow sufficient time for Cs to sorb to the mineral (Comans et al., Cs sorbed to illite inside the Float-A-Lyzer. The amount of Cs sorbed to
1991). Mineral blanks were also performed to test for any sorption to the illite in the Float-A-Lyzer was calculated as follows:
container walls. At the end of the sorption period, each tube was centri-
fuged for two hours at 3600g in order to obtain a b 50 nm particle size Csid;n ¼ Cs0 −Csbulk;n ð2Þ
cut off in the supernatant. An aliquot of the supernatant was taken
from each sample and the 137Cs measured by liquid scintillation where Csid,n is the amount of Cs (moles) on the illite inside the Float-A-
counting (LSC, Packard Tri-Carb TR2900 LSA and Ultima Gold cocktail). Lyzer after n days, Cs0 is the amount of Cs in the suspension before the
The distribution coefficient, Kd, for each sample was calculated as fol- Float-A-Lyzer was added and Csbulk,n is the amount of Cs in the bulk sus-
lows: pension outside the Float-A-Lyzer after n days. After the last sampling
time point, the Float-A-Lyzer was removed from the centrifuge tube
½Csi −½Cseq 1 and rinsed several times with Milli-Q water to remove any residue
Kd ¼  ð1Þ
½Cseq ½mineral from the outside of the Float-A-Lyzer. The Cs activity in the Float-A-
Lyzer was then measured by gamma spectroscopy to determine the
where [Cs]i is the initial concentration of Cs, [Cs]eq is the equilibrium total amount of Cs sorbed to the illite and compared to the values calcu-
concentration of Cs in the aqueous phase, and [mineral] is the lated from Eq. (2).
solid:solution ratio in g/mL. The units of Kd are mL/g.
2.5. Modeling
2.3. Binary desorption experiments
The ion exchange reaction between Cs and each of the Na-
Binary desorption experiments were performed at the conclusion of homoionized clay minerals was written as follows:
the sorption experiments. At the end of the 14-day sorption period and
after centrifugation, a total of 9.5 mL of the supernatant was quantita- Na−Clay þ Csþ ↔Cs−Clay þ Naþ ð3Þ
tively removed and replaced with 9.5 mL of Cs-free buffer solution.
The centrifuge tubes were placed in the dark on the orbital shaker and Based on this reaction an ion exchange model was developed to de-
shaken continuously for at least two weeks. The centrifuge tubes were scribe Cs sorption to each mineral using the FIT4FD code (Zavarin et al.,
then spun for 2 h at 3600 g (b50 nm particle size cut off). An aliquot 2005). The FIT4FD code was allowed to adjust the selectivity coefficient
of the supernatant was removed and 137Cs measured by LSC. The mea- and CEC parameters to reproduce Cs sorption for each of the Cs mineral
sured 137Cs was used to determine the total fraction of Cs desorbed sorption isotherms. The Vanselow convention, which is used in the
from each of the samples. At least one sequential batch exchange was model for this work, defines the selectivity coefficient (KSEL) for the re-
carried out for each mineral by repeating the above steps. action as follows:

2.4. Ternary sorption/desorption experiments Cs ðM Cs ÞZ Na ðaNa ÞZ Cs


Na K SEL ¼ ð4Þ
ðM Na ÞZ Cs ðaCs ÞZ Na
In our two-mineral ternary Cs sorption/desorption experiments,
150 mL of each clay suspension (dry clay mass of 0.15 g) was added to where MCs and MNa are the mole fractions of Cs and Na defined as the
250 mL centrifuge tubes. 137Cs was added as a tracer to stable Cs before moles of Cs or Na sorbed divided by the total number of moles sorbed,
being added to each of the mineral suspensions. Cesium concentrations Z is the cation charge (+ 1 in the case of Na and Cs), and aCs and aNa
were chosen such that Cs sorption to illite is dominated by the high af- are the activities of Cs and Na is solution, respectively (Vanselow, 1932).
finity FES sites (low Cs) and edge or planar site begin to contribute to
total Cs sorption (high Cs). Three illite suspensions with Cs concentra- 3. Results and discussion
tions of 10−6, 5 × 10−7 and 10−7 M were prepared. Three montmoril-
lonite suspensions with Cs concentrations of 5 × 10−7, 10− 7, and 5 3.1. Binary sorption
× 10−8 M were prepared and two kaolinite suspensions with Cs con-
centrations of 5 × 10−7 and 10−7 M were prepared. Lastly, solutions Binary Cs sorption experiments were conducted with illite, kaolinite,
with Cs concentrations of 10−6, 5 × 10−7, 10−7, and 5 × 10−8 M (and and montmorillonite across an initial Cs concentration range of 10−3 to
no minerals) were used as controls. All samples were placed on an 10−11 M. Fig. 1 presents these sorption data as the logarithm of the
514 C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520

Fig. 1. Cs sorption Kd values (mL/g) for ( ) Na-illite, ( ) Na-kaolinite, and ( ) Na-montmorillonite as a function of Cs concentration in solution (mol/L). All experiments performed in
0.7 mM NaHCO3, 5 mM NaCl buffer solution at pH 8. The solid solution ratio is 1 g/L for each mineral and 2 σ error bars are propagated from 2% LSC uncertainties. The solid lines
represent the sorption models with parameters developed using the FIT4FD code.

distribution coefficient (log(Kd)) versus the logarithm of the equilibri- At Cs concentrations below 5 × 10−9 M, there is a pronounced differ-
um aqueous Cs concentration. Cs sorption to illite is linear for Cs concen- ence in the extent of Cs sorption between illite, montmorillonite, and
trations below 5 × 10−9 M with an average log(Kd) value of 4.7 ± 0.4 kaolinite. The equilibrium Kd for illite is approximately two orders of
(uncertainty reported at 2 standard deviations throughout the text). magnitude greater than the Kd values for montmorillonite or kaolinite.
At Cs concentrations N 5 × 10− 9 M, the Kd decreases with increasing This highlights the importance of Cs concentration and the mineralogy
concentration. At a Cs concentration of 4 × 10− 4 M, the log(Kd) was present when predicting Cs environmental behavior. The Kd values for
2.4. The transition from linear sorption to non-linear sorption has previ- each of the three minerals are consistent with Kd values reported in re-
ously been explained by the presence of multiple sorption sites on illite cent literature. However, the specific Kd values are affected by the ionic
(Brouwer et al., 1983; Comans et al., 1991; Poinssot et al., 1999). The Cs strength (SI Fig. A2), electrolyte composition, and the mineral source
initially sorbs to the high affinity FES sites until filled and then the Cs be- used by the various authors (Ohnuki and Kozai, 2013; Benedicto et al.,
gins to sorb to the edge and planar sites. Hence, at low Cs concentrations 2014; Fuller et al., 2014; Missana et al., 2014; Reinoso-Maset and Ly,
where the FES sites are not saturated, the Kd values are high and con- 2014).
stant. As the FES sites become saturated and the Cs begins to sorb to The FIT4FD code was used to model the sorption of Cs to each of the
the less selective edge and planar sites, the Kd begins to decrease. Our minerals. Selectivity coefficients and cation exchange capacities, includ-
experiments are consistent with behavior reported in the literature ing the site concentrations which comprise the CEC of each mineral,
where the high affinity FES sites are saturated around 5 × 10−9 M were initially based on values from the literature for each clay
(Missana et al., 2014). (Missana et al., 2014). However, the selectivity coefficients were adjust-
In contrast to illite, Cs sorption to montmorillonite is linear across ed to improve the fits to the data. Model fits to the data are shown as
the entire concentration range investigated with an average log(Kd) solid lines in Fig. 1 and the fitted selectivity coefficients and cation ex-
value of 2.9 ± 0.2. Montmorillonite has previously been described as change capacities are listed in Table 1. Also listed are the selectivity co-
having only a single sorption site (Staunton and Roubaud, 1997) and efficients and cation exchange capacities from Missana et al., 2014 for
the linearity of our sorption results are consistent with that conceptual comparison. Although the source of the minerals used by Missana
model. In the case of kaolinite, sorption is linear for equilibrium Cs con- et al. (2014) is different from those used here, e.g. FEBEX bentonite vs.
centration of 10−11–10−5 M with an average log(Kd) value of 2.5 ± 0.2. SWy-1 montmorillonite, the total site concentrations presented by
Above 10−5 M Cs, sorption to kaolinite becomes non-linear.
Both 1-site and 2-site models have previously been used to describe
Table 1
the sorption of Cs to kaolinite (Erten et al., 1988; Cornell, 1993; Ohnuki,
Selectivity coefficient and site concentration parameters used in model development of Cs
1994; Missana et al., 2014). The 2-site models usually do not represent sorption onto Na-clay minerals. The uncertainties represent two standard deviations pro-
2-sites on kaolinite but rather reflect the presence of high affinity, low duced by the model.
capacity sites attributed to trace amounts of other clay minerals
log (Cs
NaKSEL) Sites (mol/g)
such as illite and vermiculite (Reinoso-Maset and Ly, 2014; Ohnuki
and Kozai, 2013; Ohnuki, 1994; Missana et al., 2014). In our work, Missana This Work Missana This Work
et al.a et al.a
X-ray diffraction (XRD) analysis showed no evidence of additional
trace minerals in montmorillonite or kaolinite (SI Figs. A3, A4). Fur- Illite FES sites 6.90 ± 0.15 6.39 ± 0.1 4.62 × 10−7 5.92 × 10−7 ± 0.08
Illite site 2 3.10 ± 0.15 3.02 ± 0.04 3.88 × 10−5 1.76 × 10−5 ± 0.13
thermore, the presence of high affinity, low capacity sites attributed
Illite site 3 1.75 ± 0.15 1.31 ± 0.04 1.55 × 10−4 1.81 × 10−4 ± 0.16
to trace phases should lead to non-linearities at low Cs concentration Smectite site 1 7.59 ± 0.15 NA 2.31 × 10−8 NA
that were not observed here. Therefore, Cs sorption to kaolinite ob- Smectite site 2 1.68 ± 0.15 0.77 ± 0.1 1.02 × 10−3 1.03 × 10−3 ± 0.15
served here can be attributed to a single cation exchange site. Kaolinite site 1 5.50 ± 0.15 NA 5.00 × 10−9 NA
Above 4 × 10− 5 M Cs, the decrease in Kd is likely caused by a satura- Kaolinite site 2 2.10 ± 0.15 2.05 ± 0.1 1.99 × 10−5 2.01 × 10−5 ± 0.07

tion of planar sorption sites. a


(Missana et al., 2014).
C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520 515

Missana et al. (2014) are similar to our fitted CEC values. Moreover, our mineral impurities in the KGa-1 kaolinite used here has been discussed
values are consistent with previously published values for the minerals in previous work (Pruett and Webb, 1993) and a 1-site model was suf-
used in this work: 19.1, 94.0, and 2.0 meq/100 g for Fithian illite ficient to simulate the sorption behavior of Cs. The Cs-kaolinite sorption
(Comans et al., 1991), SWy-1 montmorillonite (Staunton and data (Fig. 1) suggest site saturation at Cs concentrations N 10−4 M. Sim-
Roubaud, 1997) and KGa-1 kaolinite (van Olphen and Fripiat, 1979) ilar behavior was also observed by Missana et al. (2014). A second low
respectively. affinity site could be used to better fit the 2 highest concentration data
Previous evaluations of Cs sorption to illite over a wide range of Cs points. However, the CEC of this additional site would need to be
concentrations have used both 2-site and 3-site models (Poinssot more than an order of magnitude higher and unreasonable when com-
et al., 1999; Bradbury and Baeyens, 2000; Zachara et al., 2002; pared to CEC values reported in the literature for kaolinite. Furthermore,
Benedicto et al., 2014; Missana et al., 2014). The third site has been given the large errors associated with the high Cs concentration kaolin-
shown to contribute to sorption only at Cs concentrations N10−4 M ite data, the additional modeling complexity was not warranted in our
(Bradbury and Baeyens, 2000; Benedicto et al., 2014). Since the highest case.
initial Cs concentration in our binary isotherm experiments was
10−3 M, a 3-site model was used here. The fitted parameters are in 3.2. Binary desorption
very good agreement with the recently published values of Missana
et al. (2014). Batch desorption experiments were conducted to determine Cs
Cs sorption to montmorillonite has been simulated with both 1-site sorption reversibility on illite, montmorillonite, and kaolinite. In Fig. 2,
and 2-site models. Staunton and Roubaud (1997) and Iijima et al. the fraction of sorbed Cs that had desorbed from each mineral after
(2010) showed that a 1-site model can describe the sorption of Cs to two weeks is plotted against the logarithm of the equilibrium aqueous
montmorillonite (Staunton and Roubaud, 1997; Iijima et al., 2010). Cs concentration after desorption. In the case of illite, at equilibrium
When 2-site models have been used, the second site is often attributed Cs concentrations below 10− 7 M, b2% of the Cs desorbed after two
to minerals such illite or vermiculite that are interstratified in the mont- weeks. In contrast, at higher Cs concentrations, the percent Cs desorbed
morillonite (Zachara et al., 2002; Missana et al., 2004; Missana et al., increased until it reached a maximum of 33% for equilibrium Cs concen-
2014). In our case, the presence of a second site was not inferred from trations of 4 × 10−4 to 6 × 10−5 M. It is apparent that both the sorption
the data nor warranted in our modeling (see SI Fig. A5). The fitted selec- and desorption behavior of Cs is non-linear. The minimal desorption ob-
tivity coefficient for our data is an order of magnitude smaller than that served at low Cs concentrations may be suggestive of an irreversible
reported by Missana et al. (2014). This is likely the result of significant sorption phenomenon. However, the reversibility can only be evaluated
illite inclusions in the smectite used by Missana et al. (2014), however, by comparing the observed desorption behavior to predicted desorption
our fitted selectivity coefficient is in good agreement with that reported using the equilibrium sorption constants determined in the sorption
by Gast et al. (1969) for the same montmorillonite used in this work. experiments.
Gast et al. (1969) reported a log(Cs NaKSEL) of 0.79 compared to a log(Na
- At trace concentrations, irreversible sorption to illite has been attrib-
Cs
KSEL) of 0.69 reported in this work (Gast et al., 1969). uted to binding of Cs to FES sites that are susceptible to interlayer col-
For Cs sorption to kaolinite, the presence of a very low concentration lapse causing the fixation of Cs (Hird and Rimmer, 1995). Schulz et al.
but high affinity site has been speculated (Cornell, 1993) and 2-site (1960) showed that Cs sorbed on illite at low initial Cs concentrations
models have been used to describe Cs sorption (Missana et al., 2014; (b10− 7 M) exhibited “irreversible” sorption and that 10% or less of
Reinoso-Maset and Ly, 2014). However, this second site may also result the Cs was desorbed from illite even with multiple extractions (Schulz
from a trace amount of high affinity Cs sorbing mineral interstratified in et al., 1960). In other studies, b 5% of the Cs was desorbed from illite at
the kaolinite (Cornell, 1993; Kim and Cygan, 1996). The absence of concentrations below 10−7 M or less (Ohnuki, 1994; Hsu and Chang,

Fig. 2. Cs batch desorption experiments showing the fraction desorbed for ( ) Na-illite, ( ) Na-kaolinite, and ( ) Na-montmorillonite as a function of Cs concentration in solution (mol/L).
All experiments performed in 0.7 mM NaHCO3, 5 mM NaCl buffer solution at pH 8. The solid solution ratio is 1 g/L for each mineral and 2 σ error bars are propagated from 2% LSC
uncertainties. The solid lines represent the predicted fraction of Cs to desorb by the sorption models.
516 C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520

1995; McKinley et al., 2001). However, these observations of low de- Approximately 40% of Cs desorbs from montmorillonite under these
sorption should not be attributed to irreversible sorption since the conditions. The amount of Cs desorbed from montmorillonite as pre-
high affinity of Cs for FES sites would limit the overall quantities of dicted by the 1-site sorption model (Fig. 2) is in good agreement with
desorbed Cs even if these reactions are completely reversible. To test these desorption data. Thus, sorption reversibility is clearly demonstrat-
sorption reversibility, the amount of Cs to desorb from illite was simu- ed under these conditions. Approximately 50% of Cs desorbs from kao-
lated using the same model parameters developed to describe sorption linite across the equilibrium concentration range 10−12–10−5 M and
(Fig. 2). With the exception of the highest Cs concentration samples, the the 1-site sorption model effectively predicts this behavior. Above
predicted Cs desorption is in agreement with the data. The ability of the 10−5 M, the experimental data and predicted desorption behavior di-
adsorption parameters to describe the desorption behavior is indicative verge with the experimental fraction desorbed decreasing while the
of a reversible sorption process at these timescales. predicted fraction desorbed increases. However, given the large errors
As the equilibrium Cs concentration increases and the FES sites are in the experimental fraction desorbed, potential irreversibility at high
filled, subsequent adsorption occurs to the edge and planar sites for Cs concentrations cannot be confirmed with these data.
which Cs has a lower affinity. Previous work has shown that at increas- Sequential batch desorption experiments were conducted to further
ing Cs equilibrium concentrations, there is a concomitant increase in the examine reversibility of Cs sorption to each mineral. Fig. 3 shows the
amount of Cs desorbed from illite ranging from 10 to 50% depending on distributions coefficients for illite, montmorillonite and kaolinite in
factors including; Cs equilibrium concentration, ionic strength and solid our sequential batch desorption experiments. In Fig. 3-A the Kd values
solution ratio (Comans et al., 1991; Bostick et al., 2002). Our sorption for Cs desorption from illite are constant at low Cs concentrations
model also predicts that the fraction of Cs to desorb will increase with (b10−8 M). The Kd values then decrease until ~10−5 M where they ap-
equilibrium Cs concentration. pear to plateau with increasing Cs concentration. For Cs concentrations
Interestingly, our model prediction diverges from the experimental below ~10−5 M the Kd predicted by the model follow the measured de-
data at concentrations N10−5 M. This difference between the model, sorption Kd values. Above ~10−5 M the predicted desorption Kd values
which assumes reversible sorption, and experimental data is indicative underestimate the observed values. Furthermore, the second extraction
of some irreversible sorption behavior at high Cs concentrations. These leads to a larger deviation between predicted and measured Kd values.
results are consistent with the observations of De Koning and Comans The results suggest that Cs sorption to illite at high Cs concentrations
(2004) in which high concentrations of Cs were shown to exacerbate in- may be partially irreversible, consistent with our Fig. 2 observations
terlayer collapse and, thus, lead to irreversible sorption behavior at high and the findings of De Koning and Comans (2004) who attributed irre-
Cs concentrations. However, it should be noted that our desorption ex- versible sorption to the collapse of FES sites caused by the presence of
periments were performed over two weeks. Thus, slow desorption ki- excess Cs cations. However, we cannot exclude kinetics effects in
netics could also play a role in the apparent irreversible behavior at these relative short-term batch desorption experiments.
high Cs concentrations. Fig. 3-B and C show the Kd values for montmorillonite and kaolinite
In contrast to illite, Cs desorption from montmorillonite is relatively after each batch desorption, respectively (aqueous desorbed fraction re-
constant over the entire equilibrium concentration range (Fig. 2). ported in SI Figs. A6 and A7). The solid lines represent the corresponding

Fig. 3. Cs sequential batch desorption Kd values for (A) Na-illite after ( ) 14 days, ( ) 197 days, (B) Na-montmorillonite after ( ) 14 days, ( ) 142 days, ( ) 209 days, and (C) Na-kaolinite
after ( ) 14 days, ( ) 307 days. All experiments performed in 0.7 mM NaHCO3, 5 mM NaCl buffer solution at pH 8. The solid solution ratio is 1 g/L for each mineral and error bars
propagated from 2% LSC uncertainties are removed for clarity. The sold lines represent predicted Kd values from sorption models.
C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520 517

Kd values predicted by the sorption models for montmorillonite and ka- montmorillonite, which is consistent with its weaker affinity for Cs. Im-
olinite. The montmorillonite desorption Kd values are linear across the portantly, the near complete desorption of Cs from both montmorillon-
range of Cs equilibrium concentrations, are the same for each batch de- ite and kaolinite in these ternary experiments suggests that Cs sorption
sorption period, and correspond well with both the model and experi- to these minerals is not only reversible but that complete desorption
mental sorption behavior from Fig. 1. The similarity in Kd values over will occur in the presence of a competing sorbent and given sufficiently
the three extractions is consistent with reversible sorption of Cs to long timescales (months to years).
montmorillonite at all Cs concentrations. Fig. 3-C shows that kaolinite For the illite experiments (Fig. 4A), the three Cs concentrations used
desorption Kd values are also linear across the range of equilibrium con- were 1 × 10−6, 5 × 10−7, and 1 × 10−7 M and correspond to equilibri-
centrations studied. Moreover, the similarity in Kd values obtained for um Cs concentrations around the non-linear inflection point of the Kd
the two extractions also supports reversible Cs sorption to kaolinite values from Fig. 1. Overall a smaller fraction of the initially sorbed Cs
across all Cs concentrations. is desorbed from illite than from montmorillonite and kaolinite and
the desorption kinetics are exceedingly slow. Also, the amount of Cs
3.3. Ternary desorption desorbed from illite appears to be dependent on the initial Cs concentra-
tion. 52, 44 and 40% of the Cs desorbed from illite by 500 days in the ex-
Long-term two-mineral “ternary” Cs desorption experiments were periments where 1 × 10−6, 5 × 10−7, and 1 × 10−7 M Cs respectively,
carried out to investigate the reversibility and kinetics of Cs desorption was initially added. While equilibrium appears to have been reached
in the presence of an additional clay mineral To ensure that Cs sorption after one year at the highest Cs concentration, it is not clear whether
to illite inside a Float-A-Lyzer was not hindered, four control experi- equilibrium was achieved after 500 days at the two lower
ments were performed. In the control experiments, Cs was added to concentrations.
the background electrolyte solution inside the centrifuge tube and illite Compared to the batch desorption data presented in Fig. 2, the frac-
suspensions were placed in the Float-A-Lyzers. Nearly all of the Cs in so- tion of Cs desorbed is much greater in the ternary experiments. After a
lution was sorbed to the illite in the Float-A-Lyzers (SI Fig. A8) consis- single batch desorption, 50% of the Cs desorbed from the kaolinite sur-
tent with results from batch adsorption experiments. face whereas in the ternary experiments near complete desorption is
For the ternary experiments, mineral suspensions in the centrifuge observed (Fig. 4C). For montmorillonite, 40% of the Cs was desorbed in
tubes were sampled prior to adding the Float-A-Lyzers so that the initial batch experiments compared with 85% in the ternary experiments. Se-
Cs partitioning between the aqueous and mineral phase was known. quential batch desorption can approach the amounts desorbed in the
These initial conditions are reported in Table 2. Of the initial Cs added ternary experiments for low Kd minerals such as montmorillonite and
to each of the samples, 60, 30, and 5% was sorbed to illite, montmoril- kaolinite (70% total Cs desorbed, see appendix Figs. A6 and A7). Howev-
lonite, kaolinite, respectively, prior to the addition of the illite- er, that is not the case for high Kd minerals such as illite.
containing Float-A-Lyzers. Cesium desorption from illite in the ternary experiments was also
After adding the illite-containing Float-A-Lyzers, small aliquots, so as greater than that measured in binary batch desorption experiments.
not to significantly change the total amount of mineral, were taken pe- After a single binary desorption 6, 2 and b 1% of the Cs desorbed from
riodically from the outer mineral suspensions and used to measure the the illite surface for initial Cs concentrations of 10− 6, 5 × 10− 7 and
total amount of Cs in the suspension and the supernatant. The amount 10−7 M respectively. Sequential binary batch desorption resulted in
of Cs desorbed from the clays in the suspension was calculated from similarly low Cs desorption (total Cs desorbed of 10, 4 and 2%, see Ap-
the measured suspension and supernatant Cs concentrations. The pendix Figs. A6 and A7). For the same concentrations in the ternary ex-
amount of Cs sorbed to the illite in the Float-A-Lyzer was then calculated periments, 52, 44 and 40%, respectively, of the Cs desorbed.
according to Eq. (2) and assuming no loss of Cs to container walls. Fig. 4 A higher desorbed fraction of Cs in our ternary experiments com-
is a plot of the Cs fraction desorbed as a function of time for the suspen- pared to binary desorption experiments is consistent with previous re-
sions of illite, montmorillonite, and kaolinite (A, B and C, respectively). sults using ammonium hexacyanoferrate inside a dialysis membrane
For all experiments with montmorillonite and kaolinite (Fig. 4B and C, (De Koning and Comans, 2004). De Koning and Comans (2004) posit
respectively), the majority of the initially sorbed Cs was desorbed that Cs is reversibly sorbed to illite in specific conditions including low
from the mineral surface by the end of the experimental period regard- cation concentrations and a short equilibration time of 5 min. However,
less of the initial Cs concentration. However, equilibrium was achieved if cation concentrations surpass a threshold then even a 5-min equili-
only after several months (kaolinite) or over a year (montmorillonite). bration time will result in partial irreversibility of Cs (De Koning and
491 days after the Float-A-Lyzers were added, an average of 85% of the Comans, 2004). The most likely environmental scenario is a longer
Cs desorbed from the surface of montmorillonite and an average of equilibration time where some Cs diffuses into the interlayer and be-
97% of the Cs desorbed from the surface of kaolinite. Desorption of Cs comes fixed (Fuller et al., 2015).
appears to have occurred more quickly from kaolinite than from To quantitatively test for sorption reversibility in the ternary exper-
iments, Kd values for the ternary experiments were predicted from the
binary sorption models. The predicted Kd values for the ternary experi-
Table 2 ments were determined by combining the binary sorption models for il-
Initial conditions for each of the ternary experiments prior to Float-A-Lyzer addition.
lite and the mineral outside the Float-A-Lyzer. The model determined
Mineral Suspension [Cs]i [Cs]eq [Cs]solid Cs sorbed the equilibrium solution Cs concentration and the total Cs concentration
percent expected on each mineral surface at equilibrium, from which a total
mol/L
mineral Kd value was determined. The predicted amount of Cs desorbed
montmorillonite 5.0 × 10−7 3.5 × 10−7 1.4 × 10−7 28
montmorillonite 1.0 × 10−7 6.9 × 10−8 2.9 × 10−8 29
from the mineral outside the Float-A-Lyzer was determined by compar-
montmorillonite 5.0 × 10−8 3.5 × 10−8 1.4 × 10−8 28 ing the final predicted Cs concentration on that mineral to the initial
kaolinite 5.0 × 10−7 4.7 × 10−7 2.2 × 10−7 4.4 predicted Cs concentration on that mineral (i.e. Table 2). The predicted
kaolinite 1.0 × 10−7 9.4 × 10−8 5.0 × 10−9 5.0 log(Kd) values, percent desorbed and equilibrium aqueous Cs concen-
illite 1.0 × 10−6 4.6 × 10−7 5.5 × 10−7 55
trations for each case are presented in Table 3 along with the experi-
illite 5.0 × 10−7 1.9 × 10−7 3.0 × 10−7 60
illite 1.0 × 10−7 3.6 × 10−8 6.1 × 10−8 61 mental values.
none 1.0 × 10−6 9.8 × 10−7 0 0 For the ternary montmorillonite experiments, the predicted Kd and
none 5.0 × 10−7 5.0 × 10−7 0 0 percent desorption values were higher than the experimentally deter-
none 1.0 × 10−7 1.0 × 10−7 0 0 mined values. However, the differences are not statistically significant.
none 5.0 × 10−8 5.0 × 10−8 0 0
Nearly all of the Cs was predicted to desorb from the montmorillonite
518 C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520

Fig. 4. The fraction of Cs desorbed from the (A) illite, (B) montmorillonite and (C) kaolinite mineral suspensions outside the dialysis membrane over time is shown. Red, blue, orange and
purple colors represent initial [Cs] of 10−6, 5 × 10−7, 10−7 and 5 × 10−8 M, respectively. Error bars based on propagation of 2 s % LSC uncertainties are removed for clarity.

and the experimental and modeling results confirm sorption Cs concentration (10−6 M Cs) the predicted and experimental values
reversibility. are statistically equivalent. This is in contrast to the batch high concen-
For the ternary kaolinite experiments, the predicted Kd and percent tration sorption/desorption experiments and suggests that slow de-
desorption values were in agreement with the experimentally deter- sorption kinetics can lead to misidentification of irreversibility at these
mined values. The large errors associated with the kaolinite values re- concentrations.
sult from the small amount of Cs initially sorbed to kaolinite before As the initial Cs concentration decreases, the predicted and experi-
the Float-A-Lyzer was added. Thus, even small variations in the fraction mental percent Cs desorbed values diverge and the difference between
of Cs to desorb from the kaolinite can lead to large errors when propa- the predicted and experimental Kd becomes statistically significant.
gated to the calculated Kd and percent desorbed values. Despite the However, the data in Fig. 4 also suggest that equilibrium may not have
large variance, the predicted and experimental values indicate that Cs been reach even after 500 days. Thus, Cs desorption from illite is simply
is completely desorbed from the kaolinite and this is consistent with re- kinetically limited. This apparent irreversibility appears to contradict
versible sorption of Cs on kaolinite. the earlier results of the batch desorption experiments that suggested
Unlike montmorillonite and kaolinite, the predicted Kd and percent reversible sorption at these Cs concentration (Fig. 3). However, in the
desorption values for the illite ternary experiments do not agree with batch desorption experiments, a much smaller fraction of Cs desorbed
the experimental values for each concentration. At the highest initial from illite. Thus, batch desorption experiments only accessed a small

Table 3
The measured and predicted values for aqueous Cs, Kd, and percent Cs desorbed from the mineral outside the Float-A-Lyzer. The uncertainties represent two standard deviations. For the
experimental results this error is propagated from the counting error. For the predicted values, the uncertainty was determined by accounting for the uncertainty in the fitted selectivity
coefficients (Table 1).

Outside mineral Experimental Predicted

[Cs]i log[Cs]eq log(Kd) Desorbed log[Cs]eq log(Kd) Desorbed


mol/L mol/L mL/g percent mol/L mL/g percent

Montmorillonite 5.0 × 10−7 −7.5 ± 0.01 3.7 ± 0.1 84 ± 7 −8.3 ± 0.1 4.0 ± 0.1 98 ± 10
Montmorillonite 1.0 × 10−7 −8.3 ± 0.01 3.8 ± 0.1 86 ± 7 −9.1 ± 0.1 4.2 ± 0.1 99 ± 8
Montmorillonite 5.0 × 10−8 −8.8 ± 0.01 4.0 ± 0.1 88 ± 7 −9.4 ± 0.1 4.2 ± 0.1 99 ± 7
Kaolinite 5.0 × 10−7 −8.0 ± 0.01 4.2 ± 0.1 99 ± 23 −8.3 ± 0.1 4.0 ± 0.1 99 ± 10
Kaolinite 1.0 × 10−7 −8.5 ± 0.01 4.0 ± 0.1 94 ± 20 −9.1 ± 0.1 4.2 ± 0.1 99 ± 10
Illite 1.0 × 10−6 −7.5 ± 0.01 3.9 ± 0.1 52 ± 4 −8.0 ± 0.1 4.1 ± 0.1 63 ± 16
Illite 5.0 × 10−7 −7.9 ± 0.01 4.1 ± 0.1 44 ± 4 −8.5 ± 0.1 4.2 ± 0.1 65 ± 13
Illite 1.0 × 10−7 −8.7 ± 0.01 4.2 ± 0.1 40 ± 4 −9.3 ± 0.1 4.3 ± 0.1 66 ± 12
C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520 519

fraction of the most labile Cs component on the surface. These experi- Castrillejo, M., Casacuberta, N., Breier, C.F., Pike, S.M., Masque, P., Buesseler, K.O., 2016. Re-
assessment of Sr, Cs and Cs in the coast off Japan derived from the Fukushima Dai-ichi
mental results indicate that under these experimental conditions and nuclear accident. Environ. Sci. Technol. 50 (1), 173–180.
over long time periods significantly more Cs can desorb from illite Comans, R.N., Hockley, D.E., 1992. Kinetics of cesium sorption on illite. Geochim.
than the few percent that was initially determined from batch experi- Cosmochim. Acta 56 (3), 1157–1164.
Comans, R.N., Haller, M., DePreter, P., 1991. Sorption of cesium on Illite: non-equilibrium
ments over timeframes of weeks. While the ternary experiments sug- behaviour and reversibility. Geochim. Cosmochim. Acta 55, 433–440.
gest the possible existence of irreversible Cs sorption to illite, the more Cornell, R.M., 1993. Adsorption of cesium on minerals: a review. J. Radioanal. Nucl. Chem.
significant results of these experiments is that desorption kinetics on il- Artic. 171 (2), 483–500.
Davis, J.A., Coston, J.A., Kent, D.B., Fuller, C.C., 1998. Application of the surface complexa-
lite are extremely slow (months to years). tion concept to complex mineral assemblages. Environ. Sci. Technol. 32, 2820–2828.
De Koning, A., Comans, R.N., 2004. Reversibility of radiocaesium sorption on illite.
Geochim. Cosmochim. Acta 68 (13), 2815–2823.
4. Conclusions Erten, H.N., Aksoyoglu, S., Hatipoglu, S., 1988. Sorption of cesium and strontium on mont-
morillonite and kaolinite. Radiochim. Acta 44 (5), 147–151.
Batch sorption experiments demonstrate nonlinear sorption of Cs to Fukushi, K., Fukiage, T., 2015. Prediction of intrinsic cesium desorption from Na-smectite
in mixed cation solutions. Environ. Sci. Technol. 49, 10398–10404.
illite and linear sorption of Cs to montmorillonite and kaolinite. Fukushi, K., Sakai, H., Itono, T., Tamura, A., Arai, S., 2014. Desorption of intrinsic cesium
Fourteen-day batch desorption experiments, when combined with from smectite: inhibitive effects of clay partical organization on cesium desorption.
equilibrium sorption models, show that Cs desorption from these min- Environ. Sci. Technol. 48, 10743–10749.
Fuller, A.J., Shaw, S., Peacock, C.L., Trivedi, D., Small, J.S., Abrahamsen, L.G., Burke, I.T., 2014.
erals appears reversible except in the case of high concentration Cs on Ionic strength and pH dependent multi-site sorption of Cs onto a micaceous aquifer
illite. These batch experiments provided the foundation to analyze the sediment. Appl. Geochem. 40, 32–42.
results of ternary experiments. Fuller, A.J., Shaw, S., Ward, M.B., Haigh, S.J., Mosselmans, J.F., Peacock, C.L., ... Burke, I.T.,
2015. Caesium incorporation and retention in illite interlayers. Appl. Clay Sci. 108,
A methodology was developed using permeable dialysis bags to 128–134.
investigate the reversibility of Cs sorption to the minerals illite, mont- Gast, R.G., Van Bladel, R., Deshpande, K.B., 1969. Standard heats and entropies of exchange
morillonite and kaolinite. Predictions of the ternary experiment equilib- for alkali metal cations on Wyoming bentonite. America Proceedings]–>Soil Sci. Soc.
Am. Proc. 33 (5), 661–664.
rium were made using the sorption models developed from the binary
Hird, A.B., Rimmer, D.L., 1995. Total caesium-fixing potentials of acid organic soils.
experiments and then compared to the measured data from the ternary J. Environ. Radioact. 26, 103–118.
experiments. This comparison shows that Cs is reversibly sorbed to Hsu, C.N., Chang, K.P., 1995. Study of factors dominating sorption-desorption of Cs in soil.
Radiochim. Acta 68 (2), 129–133.
montmorillonite and kaolinite. The addition of a mineral with strong
Iijima, K., Tomura, T., Shoji, Y., 2010. Reversibility and modeling of adsorption behavior of
affinity for Cs such as illite was sufficient to drive the nearly complete cesium ions on colloidal montmorillonite particles. Appl. Clay Sci. 49, 262–268.
desorption of Cs from montmorillonite and kaolinite over a timescale Jackson, M.L., 1963. Interlayering of expansible layer silicates in soils by chemical
of months. However, the amount of Cs desorbed from illite was less weathering. Clay Clay Miner. 11, 29–46.
Kim, Y.K., Cygan, R.T., 1996. Cs-133 NMR study of cesium on the surfaces of kaolinite and
that predicted by our reversible sorption model. This appears to suggest illite. Geochim. Cosmochim. Acta 60 (21), 4059–4074.
some irreversible sorption of Cs to illite, likely caused by fixation of Cs in Kim, Y., Cygan, R.T., Kirkpatrick, R.J., 1996. 133Cs NMR and XPS investigation of cesium
the interlayer due to the collapse of the FES sites. However, in some ex- adsorbed on clay minerals and related phases. Geochim. Cosmochim. Acta 60 (6),
1041–1052.
periments, equilibrium was not achieved in the ternary experiments Klement, A.W., 1965. Radioactive fallout phenomena and mechanisms. Health Phys. 11,
even after 500 days. Thus, desorption kinetics on illite are extremely 1265–1274.
slow (months to years) and investigations into sorption reversibility Komarneni, S., Roy, D.M., 1979. Shale as a radioactive-waste respoitory-importance of
vermiculite. J. Inorg. Nucl. Chem. 41 (12), 1793–1796.
should be performed on the scale of years to distinguish between Komarneni, S., Roy, D.M., 1980. Hydrothermal effects of cesium sorption and fixation by
kinetically-limited and irreversible sorption behavior. clay-minerals and shales. Clay Clay Miner. 28 (2), 142–148.
McKinley, J.P., Zeissler, C.J., Zachara, J.M., Serne, R.J., Lindstrom, R.M., Schaef, H.T., Orr, R.D.,
2001. Distribution and retention of Cs-137 in sediments at the Hanford site,
Acknowledgements Washington. Environ. Sci. Technol. 35 (17), 3433–3441.
Missana, T., Garica-Gutierrez, M., Alonso, U., 2004. Kinetics and irreversibility of cesium
and uranium sorption onto bentonite colloids in a deep granitic environment. Appl.
This work was supported by the Subsurface Biogeochemical Re- Clay Sci. 26 (1–4), 137–150.
search Program of the U.S. Department of Energy's Office of Biological Missana, T., Garcia-Gutierrez, M., Benedicto, A., Ayora, C., De-Pourcq, K., 2014. Modelling
and Environmental Research. Prepared by LLNL under Contract DE- of Cs sorption in natural mixed-clays and the effects of ion competition. Appl.
Geochem. 49, 95–102.
AC52-07NA27344. This material is based upon work supported by the Novikov, A.P., Pavlotskaya, F.I., Goryachenkova, T.A., Posokhov, A.K., Kazinskaya, I.E.,
U.S. Department of Homeland Security under Grant Award Number, Emel'yanov, V.V., ... Myasoedov, B.F., 1998. Radionuclide content in underground wa-
2012-DN-130-NF0001. The views and conclusions contained in this ters and rock from observation wells around Lake Karachai. Radiochemistry 40 (5),
484–490.
document are those of the authors and should not be interpreted as Ohnuki, T., 1994. Sorption characteristics of cesium on sandy soils and their components.
representing the official policies, either expressed or implied, of the Radiochim. Acta 65, 75–80.
U.S. Department of Homeland Security. Ohnuki, T., Kozai, N., 2013. Adsorption behavior of radioactive cesium by non-mica min-
erals. J. Nucl. Sci. Technol. 50 (4), 369–375.
van Olphen, H., Fripiat, J.J. (Eds.), 1979. Data Handbook for Clay Minerals and Other Non-
Appendix A. Supplementary data metallic Minerals. Pergamon Press, New York.
Poinssot, C., Baeyens, B., Bradbury, M.H., 1999. Experimental and modelling studies of cae-
sium sorption on illite. Geochim. Cosmochim. Acta 63 (19–20), 3217–3227.
Supplementary data to this article can be found online at http://dx.
Pruett, R.J., Webb, H.L., 1993. Sampling and analysis of KGa-1B well-crystallized kaolin
doi.org/10.1016/j.scitotenv.2017.08.122. source clay. Clay Clay Miner. 41 (4), 514–519.
Reinoso-Maset, E., Ly, J., 2014. Study of major ions sorption equilibria to characterize the
ion exchange properties of kaolinite. J. Chem. Eng. Data 59, 4000–4009.
References Rich, C.I., Black, W.R., 1964. Potassium exchange as affected by cation size, pH, and min-
eral structure. Soil Sci. 97 (6), 384–390.
Benedicto, A., Missana, T., Fernandez, A.M., 2014. Inter layer collapse affects on caesium Sawhney, B.L., 1970. Potassium and cesium ion selectivity in relation to clay mineral
adsorption onto illite. Environ. Sci. Technol. 48 (9), 4909–4915. structure. Clay Clay Miner. 18, 47–52.
Bolt, G.H., Sumner, M.E., Kamphorst, A., 1963. A study of equilibria between three catego- Sawhney, B.L., 1972. Selective sorption and fixation of cations by clay minerals: a review.
ries of potassium in an illitic soil. Soil Sci. Soc. Am. Proc. 27 (3), 294–299. Clay Clay Miner. 20, 93–100.
Bostick, B.C., Vairavamurthy, M.A., Karthikeyan, K.G., Chorover, J., 2002. Cesium adsorp- Schulz, R.K., Overstreet, R., Barshad, I., 1960. On the soil chemistry of cesium 137. Soil Sci.
tion on clay minerals: an EXAFS spectroscopic investigation. Environ. Sci. Technol. 89 (1), 16–27.
36 (12), 2670–2676. Shahwan, T., Erten, H.N., Black, L., Allen, G.C., 1999. TO-SIMS study of Cs+ sorption on
Bradbury, M.H., Baeyens, B., 2000. A generalised sorption model for the concentration de- natural kaolinite. Sci. Total Environ. 226 (2–3), 255–260.
pendent uptake of caesium by argillaceous rocks. J. Contam. Hydrol. 42, 141–163. Staunton, S., Roubaud, M., 1997. Adsorption of Cs-137 on montmorillonite and illite: ef-
Brouwer, E., Baeyens, B., Maes, A., Cremers, A., 1983. Cesium and rubidium ion equilibria fect of charge compensating cation, ionic strength, concentration of Cs, K and fulvic
in Illite clay. J. Phys. Chem. 87, 1213–1219. acid. Clay Clay Miner. 45 (2), 251–260.
520 C.B. Durrant et al. / Science of the Total Environment 610–611 (2018) 511–520

Steinhauser, G., Niisoe, T., Harada, K.H., Shozugawa, K., Schnieder, S., Synal, H.-A., ... Zachara, J.M., Smith, S.C., Liu, C., Mckinley, J.P., Serne, R.J., Gassman, P.L., 2002. Sorption of
Koizumi, A., 2015. Post-accident sporadic releases of airborne radionuclides from Cs+ to micaceous subsurface sediments from the Hanford site, USA. Geochim.
the Fukushima Daiichi nuclear power plant site. Environ. Sci. Technol. 49 (24), Cosmochim. Acta 66 (2), 193–211.
14028–14035. Zavarin, M., Roberts, S.K., Hakem, N., Sawvel, A.M., Kersting, A.B., 2005. Eu(III), Sm(III),
Swartzen-Allen, S.L., Matijevic, E., 1974. Surface and colloid chemistry of clays. Chem. Rev. Np(V), Pu(V), and Pu(IV) sorption to calcite. Radiochim. Acta 93, 93–102.
74 (3), 385–400. Zavarin, M., Powell, B.A., Bourbin, M., Zhao, P.H., Kersting, A.B., 2012. Np(V) and Pu(V) ion
Turner, N.B., Ryan, J.N., Saiers, J.E., 2006. Effect of desorption kinetics on colloid-facilitated exchange and surface-mediated reduction mechanisms on montmorillonite. Environ.
transport of contaminants: cesium, strontium, and illite colloids. Water Resour. Res. Sci. Technol. 46 (5), 2692–2698.
42 (12), W12S09. Zhuang, J., Flury, M., Jin, Y., 2003. Colloid-facilitated Cs transport through water-saturated
United States Environment Protection Agency, 2009. National Primary Drinking Water Hanford sediment and Ottawa sand. Environ. Sci. Technol. 37 (21), 4905–4911.
Regulations EPA 816-F-09-004. US EPA.
Vanselow, A.P., 1932. Equilibria of the base-exchange reaction of bentonites, permutites,
soil coloids and zeolites. Soil Sci. 33 (2), 95–114.

You might also like