0% found this document useful (0 votes)
37 views10 pages

Modified Bentonite for Radioactive Iodine Removal

We present hexadecylpyridinium chloride monohydrate modified bentonite (HDPy-bent) for the efficient and selective removal of iodine anions from contaminated water. Batch experiments showed that HDPy-bent could remove more than 95% of iodide and iodate within 10 min, and had maximum iodine and iodate adsorption capacities of 80.0 and 50.2 mg/g, respectively. This work presents a promising adsorbent material for the decontamination of radioactive iodine anions from wastewater on a large scale.

Uploaded by

Pavle Mocilac
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
37 views10 pages

Modified Bentonite for Radioactive Iodine Removal

We present hexadecylpyridinium chloride monohydrate modified bentonite (HDPy-bent) for the efficient and selective removal of iodine anions from contaminated water. Batch experiments showed that HDPy-bent could remove more than 95% of iodide and iodate within 10 min, and had maximum iodine and iodate adsorption capacities of 80.0 and 50.2 mg/g, respectively. This work presents a promising adsorbent material for the decontamination of radioactive iodine anions from wastewater on a large scale.

Uploaded by

Pavle Mocilac
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 10

Chemosphere 292 (2022) 133401

Contents lists available at ScienceDirect

Chemosphere
journal homepage: www.elsevier.com/locate/chemosphere

Enhanced removal of radioactive iodine anions from wastewater using


modified bentonite: Experimental and theoretical study
Junqiang Yang a, b, Wenya Tai b, Fei Wu a, Keliang Shi a, b, c, *, Tianyi Jia b, Yin Su b,
Tonghuan Liu a, b, c, Pavle Mocilac a, b, Xiaolin Hou a, b, Ximeng Chen a, b, c
a
Frontier Science Center for Rare Isotopes, Lanzhou University, 730000, Lanzhou, PR China
b
School of Nuclear Science and Technology, Lanzhou University, 730000, Lanzhou, PR China
c
Key Laboratory of Special Function Materials and Structure Design, Ministry of Education, Lanzhou University, 730000, Lanzhou, PR China

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

• HDPy-bent showed excellent adsorption


abilities for iodide and iodate.
• Anions exchange mechanism was
confirmed by batch experiments and
DFT simulations.
• The pyridine N was the intrinsic active
site of the modified bentonite.
• HDPy-bent was efficient for removing
radioiodine anions from wastewater.

A R T I C L E I N F O A B S T R A C T

Handling Editor: Milena Horvat Efficient and cost-effective removal of radioactive iodine anions from contaminated water has become a crucial
task and a great challenge for waste treatment and environmental remediation. Herein, we present hex­
Keywords: adecylpyridinium chloride monohydrate modified bentonite (HDPy-bent) for the efficient and selective removal
Radioactive iodine of iodine anions (I− and IO3− ) from contaminated water. Batch experiments showed that HDPy-bent could
Modified bentonite
remove more than 95% of I− and IO3− within 10 min, and had maximum I− and IO3− adsorption capacities of
Adsorption
80.0 and 50.2 mg/g, respectively. Competitive experiments indicated that HDPy-bent exhibited excellent I− and
Density functional theory
Wastewater IO3− selectivity in the excessive presence of common concomitant anions including PO43− , SO42− , HCO3− , NO3− ,
Cl− (maximum mole ratio of anions vs iodine anions was ~50,000). An anion exchange mechanism was proposed
for the selective adsorption of iodine anions. Optimal adsorption structure of HDPy+/I− (IO3− ) at atomic level
and driving forces of the I− (IO3− ) adsorption were calculated by density functional theory (DFT) simulations.
Moreover, the good durability and reusability of the HDPy-bent has been demonstrated with 5 adsorption-
desorption cycles. Dynamic column experiment also demonstrated that HDPy-bent exhibited excellent removal
and fractional recovery capabilities towards I− and IO3− from simulated groundwater and environmental water
samples. In conclusion, this work presents a promising adsorbent material for the decontamination of radioactive
iodine anions from wastewater on a large scale.

* Corresponding author. Frontier Science Center for Rare Isotopes, Lanzhou University, 730000, Lanzhou, PR China.
E-mail address: [email protected] (K. Shi).

https://doi.org/10.1016/j.chemosphere.2021.133401
Received 28 October 2021; Received in revised form 12 December 2021; Accepted 21 December 2021
Available online 22 December 2021
0045-6535/© 2021 Elsevier Ltd. All rights reserved.
J. Yang et al. Chemosphere 292 (2022) 133401

1. Introduction materials (Ag+, Hg2+ and Pb2+), the toxicity, high cost, and subse­
quent desorption limit the application of these materials. In addition,
Due to increased energy demands and concern about global warm­ they are only effective for the removal of iodide (I− ) but not iodate
ing, inexpensive and carbon-free nuclear energy has been proposed as an (IO3− ) (Chapman et al., 2010; Sánchez-Polo et al., 2007). The practical
alternative to thermal power derived from fossil fuels (Apergis et al., applications of LDHs based materials in iodine anions adsorption have
2010). However, with increasing nuclear power plants operation, the been constrained due to their low surface activity and limited adsorption
radioactive waste generated in nuclear power plants presents a high risk sites, which lead to their poor selectivity and low adsorption capacities.
for public health and environment (Sawka et al., 2009; Li et al., 2021a). Porous MOF materials are a new family of the adsorbents for iodine
Radioactive iodine (129I), a product of plutonium-239 and uranium-235 capture, however, the relatively low stability and complex synthesis
fission is volatilized during the reprocessing of nuclear fuel, becoming process restrict their practical applications (Schlichte et al., 2004). In
an important environmental concern due to its long half-live of 1.57 × general, most of these adsorbents mentioned above possess high
107 years and easily mobility despite its formal low radiotoxicity. The adsorption capability mainly toward I− but not IO3− . Presently, only few
rate and amount of iodine transport is largely dependent on its chemical studies have been devoted on developing new adsorbents for the
speciation (Hansen et al., 2011; Zhang and Hou, 2021). Iodine mainly removal of IO3− (Liu et al., 2016; Theiss et al., 2016). However, a recent
exists as molecular iodine (I2), iodide (I− ) and iodate (IO3− ), depending report on iodine speciation shows that IO3− is the predominant species
on the pH and redox conditions. The iodine-containing wastewater will and accounted for up to 84% of the total iodine present in groundwater
lead to the contamination of public water sources and groundwater, samples recovered from the Hanford Site (Zhang et al., 2013). Therefore,
radioactive iodine ions dissolved in water can be ingested through food it is necessary to develop new adsorbent materials for the efficient
or water consumption, which may result in thyroiditis, disorders and removal of both radioactive I− and IO3− , with excellent adsorption
even thyroid cancer, threaten seriously the health of human being abilities, low cost, as well as environmentally friendly preparation
especially children. Therefore, remedial strategies are required to methods.
decrease 129I aqueous concentrations below the international standards Bentonite has been widely used for the removal of cationic radio­
for public consumption (WHO: 1Bq/L, DWS in USA 0.04 Bq/L) (Kaplan nuclides and pollutants (Sellaoui et al., 2018; Shi et al., 2013), owing to
et al., 2014; Wang et al., 2019). its large specific area, high cation exchange capacity and low cost.
To remove I− and IO3− from aqueous solution, various techniques However, due to the negative nature of bentonite’s charge, it is usually
have been conducted, such as adsorption (Zhang et al., 2017), ion ex­ ineffective for anionic contaminants such as I− and IO3− . Extensive
change (Lv et al., 2015), surface precipitation (Xu et al., 2016), solvent studies have shown the adsorption abilities of bentonite for anionic ra­
extraction (Fujiwara, 2016), and membrane separation (Gryta, 2013). dionuclides can be greatly enhanced by modification process, where
Among these, adsorption is notably an efficient and cleanly approach exchangeable cations (Fe3+, Al3+, HDTMA+ and CTMAB+) were inserted
because of its facile handling, high treatment efficiency, and suitability into the bentonite between layers and external surface, thus increasing
for industrial-scale treatment of large volumes of water especially under the specific surface area and/or changing the surface charge of bentonite
the continuous operation regime. Thus, much attention has been paid to (Gao et al., 2016; Li et al., 2013; Martinez et al., 2017; Li et al., 2016).
the development of various functional adsorbents. Up to now, a variety For example, in our previous work, the modified bentonites
of materials have been synthesized for the removal of iodine anions, (Fe–OOH-bent and HDPy-bent) showed excellent adsorption abilities for
such as anion exchange resins (Ye et al., 2019), clay minerals (Kaplan the removal of SeO32− , SeO42− and TcO4− from aqueous solution (Yang
et al., 2000; Li et al., 2018b), composite materials containing Ag(I) et al., 2020a, 2020b). Choung et al. researched the effects of radiation
(Chapman et al., 2010; Hoskins et al., 2002; Sánchez-Polo et al., 2007), and temperature on adsorption of I− by HDTMA modified bentonite
Hg(II) (Balsley et al., 1996), Pb(II) (Kodama, 1999), and Cu(I) (Mao considering of deep geological disposal of radioactive waste (Choung
et al., 2018; Seon and Hwang, 2021; Yadollahi et al., 2020), layered et al., 2014). Dultz and Riebe et al. have demonstrated that bentonite
double hydroxides (LDHs) (Li et al., 2018a; Yazdankish et al., 2020), and modified with certain cationic surfactants such as hexadecylpyridinium
metal-organic frameworks (MOFs) (He et al., 2021; Wang et al., 2021). (HDPy+), hexadecyltrimethylammonium (HDTMA+), benzethonium
However, most of them have their own drawbacks. For instance, some of (BE+) and dipyridinododecane (DPyDD2+) exhibit high adsorption ca­
anion exchange resins have low removal rates and high swelling which pabilities for 125I− (Dultz et al., 2005; Riebe et al., 2005). Despite of
seriously blocked quick column separation efficiency. For metal-based some investigations concerned modified bentonite as adsorbents for

Fig. 1. The schematic diagram of synthesis process for HDPy-bent.

2
J. Yang et al. Chemosphere 292 (2022) 133401

Fig. 2. Effects of contact time on adsorption of (a) I− and (b) IO3− ; Effect of initial concentration on adsorption of (c) I− and (d) IO3−

radioactive iodine anions, most of the current studies mainly focused on Phenom Prox, Holland). FT-IR spectroscopy (NEXUS 670, American)
iodide (I− ) transportation. Systemic adsorption behavior of I− and IO3− from 4000 cm− 1 to 400 cm− 1 was selected to verify the HDPy-bent
on modified bentonite are still in the maze. The understanding of the structure after I− and IO3− adsorbed. The chemical surface of samples
adsorption structure at the atomic level and the interactions between was identified by X-ray photoelectron spectroscopy (XPS) (Kratos AXIS
I− /IO3− and adsorbents at the electronic level are necessary to guide the UltraDLD, Kyoto, Japan) using Al-Kα radiation (hν = 1486.6 eV) with
design of continuous sequestration technologies based on modified operation conditions of 15 kV and 20 mA. Survey spectra were used for
bentonites. the examination of the elements on the surface. The zeta potential of
Accordingly, the objectives of this study are: 1) to quantify the samples was measured at ambient temperature (298 K) using Zetasizer
adsorption abilities of HDPy-bent for removing I− and IO3− from Nano ZS90 (Malvern Nano ZS, UK).
aqueous solution using batch adsorption experiments, with attention to
the effects on the presence of concomitant anions (SO42− , NO3− , CO32− , 2.3. Batch and column experiments
Cl− and Br− ); 2) to identify the adsorption mechanisms of I− and IO3−
onto HDPy-bent by characterization and density functional theoretical The adsorption of I− and IO3− onto HDPy-bent was carried out in
calculations; and 3) to investigate the behavior of chromatographic batch experiments by varying experimental conditions including solu­
separation for I− and IO3− using column filled with HDPy-bent. The tion pH, initial concentration of I− and IO3− , contact time and co-
work will provide a novel approach for the efficient removal of radio­ existing anions. In a typical batch experiment, the fixed adsorbent
active I− and IO3− from wastewater. dispersion (solid-liquid ratio was 2 g/L) was mixed with I− and IO3−
stock solution of fixed concentration, in advance adjusted to the
2. Materials and methods required pH using NaOH or HCl solution. After the suspension was
stirred at 298 ± 1 K for a desired time, the solid phase was separated by
2.1. Chemicals and materials centrifugation, the supernatant was analyzed by HR-ICP-MS.
Desorption of I− and IO3− was carried out as follows: the loaded
Sodium iodide (NaI, purity >99.9%) and potassium iodate (KIO3, adsorbent was dispersed in desired desorption reagent solution and
purity >99.8%) were purchased from Sigma Aldrich. Analysis-grade stirred for 12 h. After which the solid phase was separated by centrifu­
(HDPyCl⋅H2O; Fluka AG, Switzerland) was used for the preparation of gation. The recycled adsorbent was re-used for adsorption after being
the organoclay. Deionized water (18.2 MΩ cm) was used in all experi­ washed with both highly deionized water and ethanol, and final drying
ments. The HDPy-bent was synthesized as described in our previous under vacuum. The adsorption-desorption processes were repeated 5
work (Yang et al., 2020a). The schematic diagram of synthesis process times under the same condition as above described.
for HDPy-bent was shown in Fig. 1 and the details were described in Column experiments were conducted to investigate the dynamic
Supporting Information. process of I− and IO3− adsorption by HDPy-bent, which was packed into
a column (4.3 cm × 0.8 cm) and set to remove 100 μg/L I− and IO3− from
2.2. Characterization two simulated groundwater (SGW) and deionized water (DW) samples.
The chemical composition of the SGW samples is listed in Table S1. The
The morphology of HDPy-bent before and after adsorption of I− and tap water and Yellow River water (Lanzhou, China) were collected for
IO3− were characterized with scanning electron microscopy (SEM-EDS, further evaluating the practical application ability of HDPy-bent, the

3
J. Yang et al. Chemosphere 292 (2022) 133401

Fig. 3. (a) Effects of solution pH on adsorption of I− and IO3− ; (b) Zeta potential of HDPy-bent, HDPy-bent adsorption of I− and IO3−

water samples had been pretreated before using (Supporting Informa­


Table 1
tion). The feed flow rate was controlled by a vacuum chamber (As-
Adsorption capacities for I− and IO3− by different adsorbents.
EH20B, Brazil) and a peristaltic pump (ZK-26/25, MIULAB), after which
the eluted solution was collected and analyzed. Adsorbents Adsorbates Equilibrium Adsorption References
time capacities
The concentration of I− , IO3− and Cl− was measured by HR-ICP-MS
(Thermo Scientific, USA). These experiments were performed in dupli­ Ag@activated I− – 38.1 mg/g Hoskins
cate and the average values of two replicates ±5% uncertainties were carbon et al. (2002)
Surfactant I− 10 min 70.0 mg/g Choung
estimated in the whole adsorption experiment. The equilibrium modified et al. (2014)
adsorption amount (qe, mg/g) and adsorption ratio (R%) were calculated bentonite
as follows: Modified zeolite I− – 3.37 mg/g Warchoł
et al. (2006)
C0 − Ce Organoclays I− 30.0 mg/g Li et al.
qe = ×V (1) –
m (2018a)
Organoclays IO3− – 29.0 mg/g Li et al.
C0 − Ce (2018a)
R% = × 100% (2) Nanosized IO3− 50 h 3.9 mg/g Yu et al.
C0
tubular (2020)
halloysite
where C0 (mg/L) and Ce (mg/L) were I− and IO3− initial and equilibrium
HDTMA- I− 12 h 36.1 mg/g Chen et al.
concentrations, respectively; V (mL) is the volume of the aqueous phase; geopolymer (2019)
and m (mg) is the weight of HDPy-bent adsorbent. In addition, the data BCA@n235 I− 2h 110.5 mg/g Sun et al.
of adsorption kinetics and adsorption isotherms were fitted with (2019)
CTS/HMMT I− 80 min 47.9 mg/g Li et al.
different mathematical models, respectively, which were illustrated in
(2019)
Supporting information. HDPy-bent I−
5 min 80.0 mg/g this work
HDPy-bent IO3− 30 min 50.2 mg/g
2.4. Density functional theory calculation

Density functional theory (DFT) was employed to optimize the 3. Results and discussion
structure of HDPy+, HDPy+/IO3− , HDPy+/I− , HDPy+/Br− , HDPy+/Cl− ,
HDPy+/SO42− and HDPy+/PO43− in the gas phase at B3LYP-D3/ 3.1. Adsorption abilities of I− and IO3− onto HDPy-bent
6–311++G** level (Grimme et al., 2010; Lee et al., 1988). For iodine
Andreas Bergner’s quasi-relativistic energy-adjusted ab initio pseudo­ 3.1.1. Effect of contact time and initial concentration
potentials (Bergner et al., 1993) and Martin’s corresponding The effect of the contact time on I− and IO3− adsorption by HDPy-
(14s10p3d2f1g)/(4s4p3d2f1g) basis set (Martin and Sundermann, bent was investigated at the temperature of 298 ± 1 K, and the results
2001) were adopted. For each species, several initial geometries were were presented in Fig. 2a and b. The adsorption capacity for the iodine
considered in the optimization process to make sure that the lowest anions of the HDPy-bent absorbent increased rapidly at the first stage
energy structure can be captured. The optimized complex structures (0–10 min), more than 95% of the equilibrium adsorption capacity was
then undergo single point energy calculations in the liquid phase achieved within 10 min for both I− and IO3− , possibly ascribed to the
(M06–2X/6–311++G** level). To consider the effect of water solvent, high positive zeta potentials of HDPy-bent (Fig. 3b), which provides a
SMD model was used. Binding energies (Ebinding) were calculated based strong driving force to overcome all mass transfer resistances of I− and
on the following equation: IO3− between the solution and solid phases. Therefore, the mass transfer
through the solution was not rate-limiting. Significantly, the adsorption
Ebinding = Ecomplex − Eadsorbent − Eadsorbate (3) rate of HDPy-bent for I− was much faster than commercial resins D201
The Eadsorbent, Eadsorbate and Ecomplex are single point energies of each and IRA-900 (>3 h), silica-based ion exchange resin (30 min) (Ye et al.,
adsorbents, adsorbates and their complexes. For all the calculations in 2019), Fe3O4@Mg/Al LDH (5 h) (Jung et al., 2020), and Cu-Based Ad­
the current work Gaussian 16 software (Frisch, 2016) was used. sorbents (8 h) (Mao et al., 2018). The time dependent data was further
analyzed by employing the pseudo-second-order model (Supporting
information), and the corresponding kinetic model parameters were
given in Table S2, From these observations we could notice that the
correlation coefficient value fitted well with the pseudo-second-order
model, which indicates that the adsorption would be a chemical

4
J. Yang et al. Chemosphere 292 (2022) 133401

Fig. 4. Effects of co-existing anions on adsorption of (a) I− and (b) IO3−

process (Bagherifam et al., 2014). 3.1.3. Effect of co-existing anions


The adsorption capacities of HDPy-bent for I− and IO3− were In general, co-existing anions could compete with the target iodine
investigated by changing the initial concentration and the results were anions on the surface of adsorbents because of the higher charge density
shown in Fig. 2c and d. It suggests that the adsorption capacity enhanced as well as similarity of the molecular structure (Zhu et al., 2017; Shen
with the increase of I− and IO3− concentration, and the Langmuir and et al., 2020). These competing co-existing anions adversely interfere
Freundlich models were employed to fit the experimental results (Sup­ with the adsorption of I− and IO3− when dealing with the real waste­
porting information). The parameters are summarized in Table S3 and water. Therefore, adsorption experiments were performed in the pres­
the adsorption isotherms of I− and IO3− were described better by the ence of different concentrations of PO43− , SO42− , HCO3− , NO3− , Cl− and
Langmuir model (R2>0.99), indicating a monolayer adsorption as the Br− to investigate the effect of competitive co-existing anions. As shown
primarily dominant adsorption process. The maximum adsorption ca­ in Fig. 4a, the adsorption of I− is slightly inhibited in the presence of
pacities of I− and IO3− calculated by fitting are 80.0 and 50.2 mg/g, high concentration (>0.005 mol/L) concomitant PO43− , SO42− , HCO3− ,
respectively. Notably, the adsorption capacities of HDPy-bent are however the adsorption ratios of I− is still up to 90%. Unfortunately,
significantly better than for the most homogeneous materials (Table 1), when Br− was the competitive anion with high concentration (>0.01
such as HDTMA+ modified bentonite (Choung et al., 2014), modified mol/L), the I− adsorption efficiencies was reduced significantly. Clearly,
zeolite (Warchoł et al., 2006), organ-clay and other adsorbent materials this was due to the strong competitive adsorption of high concentration
(Li et al., 2018b; Yu et al., 2020), most of them are almost ineffective for Br− . It should be noted that in the real samples the concentration of Br−
the adsorption of IO3− . Considering its fast adsorption kinetics, high is usually relatively low (7.69 × 10− 8–5.61 × 10− 5 mol/L in fresh water,
adsorption capacities especially for IO3− , low-cost raw material and fresh groundwater and rainwater) (Vainikka and Hupa, 2012). Thus, the
simple synthesis method, HDPy-bent is particularly promising adsorbent decrease of the iodide decontamination efficiency of HDPy-bent due to
material for the treatment of radioiodine-contaminated wastewater. the bromide ions has no practical effect on real water samples.
Furthermore, it can be found that the adsorption efficiency for IO3− was
3.1.2. Effect of solution pH still more than 95% even when the concentration of co-existing anions
The pH of the aqueous solution is one of the most important factors was up to 0.5 mol/L (mole ratio of anions/IO3− was ~56,000) (Fig. 4b),
for the adsorption of iodine that could influence the surface charges and indicating that the adsorption of IO3− onto HDPy-bent has priority over
potential of the adsorbents. The effect of pH on I− and IO3− adsorption other competitive anions in the competitive adsorption process because
onto HDPy-bent were investigated and the results were shown in Fig. 3. of its lower binding energy than anions with HDPy+ (see DFT calcula­
As can be observed, the adsorption ratio of HDPy-bent both for I− and tion). The results showed that HDPy-bent have better selectivity for the
IO3− remains at a high level (>95%) within a wide pH range, spanning adsorption of I− and IO3− than most reported adsorbent materials, such
from 2 to 11 (Fig. 3a), indicating that HDPy-bent is a pH-stable material as organoclays, granular activated carbon (Li et al., 2018b), LDH
for decontamination of wastewater from radioiodine anions. It should be (Fe3O4@Mg/Al) (Jung et al., 2020), Ag@Cu2O nanoparticles (Mao
noted that many adsorbents for I− and IO3− adsorption (such as modi­ et al., 2016) and Bi-GO (Han et al., 2019).
fied bentonite and nano-Bi2O2.33) showed high capacity only in a narrow
pH range, beyond which the adsorption ratio of I− and IO3− significantly
declined due to the change of pH (Jang and Lee, 2018; Liu et al., 2016). 3.2. Adsorption mechanism
Measurement of HDPy-bent zeta potential indicated that it was posi­
tively charged across the entire tested pH range of 2.5–10.5 (Fig. 3b), The mechanisms of I− and IO3− adsorption onto HDPy-bent were
while the zeta potential decreased slightly with increasing solution pH, investigated by using various approaches. The relationship between the
from 48 mV (pH 4.0) to 33 mV (pH 10.0). Such stability of positive quantities of adsorbed I− /IO3− and released Cl− was investigated and
surface charge of HDPy-bent ensures strong affinity for anions, the the results are presented in Fig. 5a and b. The equilibrium concentration
surface charge of HDPy-bent remains positive even after adsorption of of adsorbed I− and IO3− and released Cl− in adsorption process was well
iodine anions (I− and IO3− ). fitted, with the slop of fitted line close to 1 (R2 > 0.99), which confirmed
the anion-exchange mechanism both for I− and IO3− . SEM-EDS spec­
trums showed that Cl content was significantly reduced and iodine

5
J. Yang et al. Chemosphere 292 (2022) 133401

Fig. 5. Cl− release content vs (a) I− and (b) IO3− adsorption in adsorption process; (c) The FT-IR spectrum of HDPy-bent before and after adsorption of I− and IO3−
samples; (d) High-resolution XPS spectrum of N 1s on HDPy-bent before and after adsorption of I− ; High-resolution XPS spectrum of (e) O 1s and (f) N 1s on HDPy-
bent before and after adsorption of IO3−

content increased after adsorption, which also support the postulated undertaken to clarify the nature of bonding in the adsorption process.
anion-exchange mechanism (Fig. S3 and Table S4). There were no new The phenomenon of I 3d appearance and Cl 2p disappearance of on the
FT-IR peaks observed after adsorption of I− and IO3− (Fig. 5c), probably survey scan spectra of HDPy-bent after I− and IO3− adsorption also
due to only small amount of iodine anions adsorbed onto HDPy-bent, confirmed the anion exchange mechanism (Fig. S4). Further analysis of
while the characteristic peaks of HDPy-bent at 2920.0 cm− 1, 2050.6 N 1s high resolution spectra of HDPy-bent after I− and IO3− adsorption
cm− 1, 1642.2 cm− 1, 1492.4 cm− 1 and 1472.0 cm− 1 got slightly redshift, revealed that (Fig. 5d and f), binding energies of N 1s shifted from 399.3
indicating that the functional groups (HDPy+) might participate in the to 399.75/399.8 eV for I− and IO3− , respectively, confirming the for­
adsorption of I− and IO3− (Qu et al., 2021; Yang et al., 2020a). mation of a chemical bond. Besides, the new peak at 402.5 eV was
To further investigate the adsorption mechanism details, XPS was assigned to N–O signal after adsorption of IO3− (Zhang and Chen, 2018),

6
J. Yang et al. Chemosphere 292 (2022) 133401

Fig. 8. Molecular orbital interaction diagrams of (a) HDPy+/IO3− (b)


Fig. 6. Electrostatic potential (ESP) of HDPy , IO3 and I
+ − −
HDPy+/I−

Table 2
Binding energies of HDPy+/anions from DFT simulation calculations.
Anion IO3− Br− I− Cl−

Binding energy (kJ/mol) − 34.57 − 22.68 − 16.82 − 2.97

carbon (Fig. 7b). HDPy+/I− exhibited a slightly different structure, with


I− adsorbed on the side of pyridinium ring, close to C2 carbon.
To identify the driving forces of IO3− and I− adsorption, electronic
structures of HDPy+/IO3− and HDPy+/I− adducts were investigated.
From Fig. 8a, we can clearly see p orbital of the IO3− oxygen hybridized
with π orbital of the pyridinium ring, with the p orbital located on the
top of the π orbital (see Fig. 8a). Such p-π interaction seems to be the
driving force of IO3− adsorption. This agrees with Wang’s work and our
previous work concerning absorption of ReO4− , which identified p-π
interaction as the driving force (Li et al., 2018c; Yang et al., 2022). Based
on our simulation (Fig. 8b), the driving force of I− adsorption should also
Fig. 7. (a) Optimal adsorption structure of HDPy+/IO3− and HDPy+/I− (b) be ascribed to a slightly different p-π interaction where p orbital of
label of C and N atoms in the pyridinium ring of HDPy+. iodine is being hybridized with π orbital of the pyridinium ring. Unlike
HDPy+/IO3− , in HDPy+/I− adsorption adduct the p orbital is not on the
the new O–N and O–I signals appeared in Fig. 5e, which demonstrated top of π orbital but on the corner of π orbital. This can be related to the
that N of pyridine plays an important role in the adsorption process, difference in the structure of HDPy+/IO3− vs. HDPy+/I− adsorption
especially for the IO3− adsorption (Zhong et al., 2021). Combined with adducts. In HDPy+/IO3− , the oxygen involved in the p-π interaction is
the above results, we propose that the adsorption mechanism of I− and directly located on the top of C6 carbon. For comparison, in HDPy+/I−
IO3− onto HDPy-bent is an anion exchange process, where I− and IO3− the iodine involved in the p-π interaction is on the side of C2. The current
are exchanged with Cl− in adsorption process without involving the work and previous work/literature highlighted the importance of p-π
redox reaction. interaction (Li et al., 2018c; Yang et al., 2022), which could lead to a
strategy for designing decontamination materials that could potentially
3.3. Density functional theory calculation form strong p-π interactions with contaminants.
Our experiments showed that I− adsorption efficiency was reduced
To shed light on the structure of HDPy+/I− , HDPy+/IO3− adsorption with high Br− concentration whereas Cl− has no obvious effect on I−
adducts and elucidate the driving forces for I− , IO3− adsorption, gas adsorption. This could be explained by calculating binding energies of
phase DFT calculations were performed. Electrostatic potential (ESP) HDPy+/I− , HDPy+/Br− , HDPy+/Cl− adsorption adducts from DFT sim­
results (Fig. 6) and our previous paper suggested that pyridinium ring of ulations. As our prior findings as well as previous literature suggest, the
HDPy+ has higher (bluer) ESP than alkyl chain part (Yang et al., 2022). quantity could be used to explain the adsorption preference of anion
Species that have higher positive ESP should have stronger electrostatic contaminants (Li et al., 2018c; Yang et al., 2022). Anions with lower
interactions with species that have negative ESP (Zhong et al., 2021; Li binding energies are preferred in the adsorption process. Because the
et al., 2021b). Therefore, iodine anions (IO3− , I− ) with negative ESP binding energy of HDPy+/I− is higher than of HDPy+/Br− (see Table 2),
(highlighted in orange and red) should be adsorbed by the pyridinium HDPy+ cations prefer to adsorb Br− instead of I− . Therefore, I− anions
ring of HDPy+ (highlighted in blue). A prediction that I− and IO3− would are less likely to be adsorbed by HDPy+ when high concentration of Br−
be adsorbed by the pyridinium ring of HDPy+ was further confirmed by is present. Since the binding energy of HDPy+/I− is lower than
the optimized structures of HDPy+/IO3− and HDPy+/I− adducts HDPy+/Cl− , Cl− is less preferred in the adsorption process. Conse­
(Fig. 7a), in which IO3− is adsorbed on the top of pyridinium ring, with quently, Cl− has no obvious effect on the adsorption of I− . Similarly,
one oxygen on the top of C6 carbon and two oxygen close to C2 and C3 HDPy-bent is more likely to adsorb I− and IO3− than SO42− and PO43− in

7
J. Yang et al. Chemosphere 292 (2022) 133401

SGW#1, SGW#2 and DW samples, respectively. With the same pro­


cedure, IO3− complete breakthrough points of 390, 420 and 450 mL bed
volume for SGW#1, SGW#2 and HDW samples were obtained (Fig. 10
and Table S5). Obviously, the treatment volume of SGW samples was
less than DW samples, the results can be explained by larger amounts of
concomitant ions in SGW samples, which declined the adsorption ca­
pacities both for I− and IO3− . Furtherly, two environmental water
samples (tap water and Yellow River water) were conducted to evaluate
the practical application of HDPy-bent with label recovery method, I−
and IO3− (10 μg/L) were artificial added to the environmental samples,
respectively. The results showed high recoveries (more than 90%) both
for I− and IO3− in the experiments. In considering of the low-cost and
simple synthesis of HDPy-bent, these results furtherly demonstrated that
HDPy-bent is a promising material to treat wastewater contained
radioactive iodine.

4. Conclusion

In summary, a low-cost modified bentonite adsorbent (HDPy-bent)


was synthesized by a facile surfactant modification approach under mild
Fig. 9. The adsorption cycles of I− and IO3− on HDPy-bent.
conditions. The HDPy-bent showed to be a highly efficiency for the
adsorption for iodine anions (I− and IO3− ). It possessed fast adsorption
kinetics and high adsorption capacities at a wide pH range (2–11) for
the co-existing anions system because of higher energies of the latter
both I− and IO3− . Its selectivity towards IO3− and I− is due to the lower
(Fig. S5) (HDPy+/SO42− (11.76 kJ/mol), HDPy+/PO43− (51.50
binding energies than in the case of other HDPy+/anion adducts. The
kJ/mol)).
underlying adsorption mechanism was thoroughly elucidated by batch
experiments, samples characterization and DFT calculations, showing
3.4. Desorption and adsorbent recycling that anions exchange process is responsible for I− and IO3− adsorption.
In addition, p-π interactions between oxygen of IO3− /I− and pyridinium
The regeneration of the HDPy-bent is an important consideration in ring of HDPy + are the main driving forces of IO3− /I− adsorption.
practical application. Herein, various reagent solutions (NaOH, NaCl, Further, HDPy-bent was packed into column for successfully sequential
NaNO3, C2H5OH, HCl and HNO3) were employed for the desorption of removal and recovery of I− and IO3− from environmental water samples
IO3− and I− adsorbed onto HDPy-bent, as shown in Fig S6. By optimizing contaminated with iodine anions. Our study demonstrated an important
the concentration and types of desorption reagents, the desorption ratio proof-of-concept for the modified bentonite as adsorbent for iodine
of I− can reach more than 90% using 2.5 M NaNO3, while nearly 100% of anions and indicated the importance of considering strong p-π interac­
IO3− can be desorbed using 8 M HNO3. In order to evaluate the reus­ tion with contaminants in designing similar adsorbent materials.
ability of HDPy-bent, the adsorption-desorption cycles were repeated 5
times using 2.5 M NaNO3 as an eluting reagent (Fig. S6). From Fig. 9, Author contributions statement
HDPy-bent presented an ignorable decrease in adsorption capacity after
5 adsorption-desorption cycles, testifying that HDPy-bent was stable and Junqiang Yang: Conceptualization, Methodology, Formal analysis,
recyclable adsorbent for I− and IO3− adsorption. Investigation, Writing - original draft, Visualization. Wenya Tai:
Methodology, Investigation. Fei Wu: Software, Formal analysis.
3.5. Column experiments Keliang Shi: Conceptualization, Writing - Review & Editing, Supervi­
sion, Funding acquisition. Tianyi Jia: Investigation. Yin Su: Investiga­
The series of column experiments were conducted to assess the tion. Tonghuan Liu: Investigation. Pavle Mocilac: Investigation,
suitability of HDPy-bent in a more practical column set-up. Fig. 10 Writing-review & editing. Xiaolin Hou: Writing-review & editing,
displayed the breakthrough profiles of I− and IO3− at 100 μg/L with Investigation. Ximeng Chen: Investigation.
relatively stable flow rate of 0.25–0.5 mL/min. The I− complete break­
through point took place at 480, 570 and 660 mL bed volume for

Fig. 10. The breakthrough curves of (a) I− and (b) IO3− in purified water and simulated groundwater 1# and 2#.

8
J. Yang et al. Chemosphere 292 (2022) 133401

Declaration of competing interest Jang, J., Lee, D.S., 2018. Magnetite nanoparticles supported on organically modified
montmorillonite for adsorptive removal of iodide from aqueous solution:
optimization using response surface methodology. Sci. Total Environ. 615, 549–557.
The authors declare that they have no known competing financial Jung, I.-K., Jo, Y., Han, S.-C., Yun, J.-I., 2020. Efficient removal of iodide anion from
interests or personal relationships that could have appeared to influence aqueous solution with recyclable core-shell magnetic Fe3O4@Mg/Al layered double
the work reported in this paper. hydroxide (LDH). Sci. Total Environ. 705, 135814.
Kaplan, D.I., Denham, M.E., Zhang, S., Yeager, C., Xu, C., Schwehr, K.A., Li, H.P., Ho, Y.
F., Wellman, D., Santschi, P.H., 2014. Radioiodine biogeochemistry and prevalence
Acknowledge in groundwater. Crit. Rev. Environ. Sci. Technol. 44 (20), 2287–2335.
Kaplan, D.I., Serne, R.J., Parker, K.E., Kutnyakov, I.V., 2000. Iodide sorption to
subsurface sediments and illitic minerals. Environ. Sci. Technol. 34 (3), 399–405.
The financial support from National Natural Science Foundation of Kodama, H., 1999. Removal of iodide ion from simulated radioactive liquid waste.
China (Nos. 22061132004; 21771093; 22106059), Fundamental Czech. J. Phys. 49 (S1), 971–977.
Lee, C., Yang, W., Parr, R.G., 1988. Development of the Colle-Salvetti correlation-energy
Research Funds for the Central Universities of China (lzujbky-2021-
formula into a functional of the electron density. Phys. Rev. B 37 (2), 785–789.
kb11; lzujbky-2021-sp41) and Gansu guiding program of Science and Li, C., Wei, Y., Wang, X., Yin, X., 2018a. Efficient and rapid adsorption of iodide ion from
Technology Innovation (No.20JR10RA610) are gratefully appreciated. aqueous solution by porous silica spheres loaded with calcined Mg-Al layered double
This work was supported by Supercomputing Center of Lanzhou hydroxide. J. Taiwan Inst. Chem. Eng. 85, 193–200.
Li, D., Kaplan, D.I., Sams, A., Powell, B.A., Knox, A.S., 2018b. Removal capacity and
University. chemical speciation of groundwater iodide (I(-)) and iodate (IO3(-)) sequestered by
organoclays and granular activated carbon. J. Environ. Radioact. 192, 505–512.
Li, Y., Bi, M., Wang, Z., Li, R., Shi, K., Wu, W., 2016. Organic modification of bentonite
Appendix A. Supplementary data
and its application for perrhenate (an analogue of pertechnetate) removal from
aqueous solution. J. Taiwan Inst. Chem. Eng. 62, 104–111.
Supplementary data to this article can be found online at https://doi. Li, J., Chen, L., Shen, N., Xie, R., Sheridan, M.V., Chen, X., Sheng, D., Zhang, D., Chai, Z.,
Wang, S., 2021a. Rational design of a cationic polymer network towards record high
org/10.1016/j.chemosphere.2021.133401.
uptake of 99TcO4− in nuclear waste. Sci. China Chem. 64 (7), 1251–1260.
Li, J., Dai, X., Zhu, L., Xu, C., Zhang, D., Silver, M.A., Li, P., Chen, L., Li, Y., Zuo, D.,
References Zhang, H., Xiao, C., Chen, J., Diwu, J., Farha, O.K., Albrecht-Schmitt, T.E., Chai, Z.,
Wang, S., 2018c. 99TcO4− remediation by a cationic polymeric network. Nat.
Commun. 9 (1), 3007.
Apergis, N., Payne, J.E., Menyah, K., Wolde-Rufael, Y., 2010. On the causal dynamics
Li, J., Li, B., Shen, N., Chen, L., Guo, Q., Chen, L., He, L., Dai, X., Chai, Z., Wang, S.,
between emissions, nuclear energy, renewable energy, and economic growth. Ecol.
2021c. Task-specific tailored cationic polymeric network with high base-resistance
Econ. 69 (11), 2255–2260.
for unprecedented 99TcO4– cleanup from alkaline nuclear waste. ACS Cent. Sci. 7 (8),
Bagherifam, S., Komarneni, S., Lakzian, A., Fotovat, A., Khorasani, R., Huang, W., Ma, J.,
1441–1450.
Hong, S., Cannon, F.S., Wang, Y., 2014. Highly selective removal of nitrate and
Li, Q., Mao, Q., Yang, C., Zhang, S., He, G., Zhang, X., Zhang, W., 2019. Hydrophobic-
perchlorate by organoclay. Appl. Clay Sci. 95, 126–132.
modified montmorillonite coating onto crosslinked chitosan as the core-shell micro-
Balsley, S.D., Brady, P.V., Krumhansl, J.L., Anderson, H.L., 1996. Iodide retention by
sorbent for iodide adsorptive removal via Pickering emulsion polymerization. Int. J.
metal sulfide surfaces: cinnabar and chalcocite. Environ. Sci. Technol. 30 (10),
Biol. Macromol. 141, 987–996.
3025–3027.
Li, Z., Yao, M., Lin, J., Yang, B., Zhang, X., Lei, L., 2013. Pentachlorophenol sorption in
Bergner, A., Dolg, M., Küchle, W., Stoll, H., Preuß, H., 1993. Ab initio energy-adjusted
the cetyltrimethylammonium bromide/bentonite one-step process in single and
pseudopotentials for elements of groups 13–17. Mol. Phys. 80 (6), 1431–1441.
multiple solute systems. J. Chem. Eng. Data 58 (9), 2610–2615.
Chapman, K.W., Chupas, P.J., Nenoff, T.M., 2010. Radioactive iodine capture in silver-
Liu, S., Kang, S., Wang, H., Wang, G., Zhao, H., Cai, W., 2016. Nanosheets-built
containing mordenites through nanoscale silver iodide formation. J. Am. Chem. Soc.
flowerlike micro/nanostructured Bi2O2.33 and its highly efficient iodine removal
132 (26), 8897–8899.
performances. Chem. Eng. J. 289, 219–230.
Chen, S., Qi, Y., Cossa, J.J., Deocleciano Salomao Dos, S.I., 2019. Efficient removal of
Lv, L., Yang, J., Zhang, H.-M., Liu, Y.-Y., Ma, J.-F., 2015. Metal-ion exchange, small-
radioactive iodide anions from simulated wastewater by HDTMA-geopolymer. Prog.
molecule sensing, selective dye adsorption, and reversible iodine uptake of three
Nucl. Energy 117, 103112.
coordination polymers constructed by a new resorcin[4]arene-based
Choung, S., Kim, M., Yang, J.S., Kim, M.G., Um, W., 2014. Effects of radiation and
tetracarboxylate. Inorg. Chem. 54 (4), 1744–1755.
temperature on iodide sorption by surfactant-modified bentonite. Environ. Sci.
Mao, P., Jiang, J., Pan, Y., Duanmu, C., Chen, S., Yang, Y., Zhang, S., Chen, Y., 2018.
Technol. 48 (16), 9684–9691.
Enhanced uptake of iodide from solutions by hollow Cu-based adsorbents. Materials
Dultz, S., Riebe, B., Bunnenberg, C., 2005. Temperature effects on iodine adsorption on
11 (5).
organo-clay minerals. Appl. Clay Sci. 28 (1–4), 17–30.
Mao, P., Liu, Y., Jiao, Y., Chen, S., Yang, Y., 2016. Enhanced uptake of iodide on Ag@
Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, J.R.,
Cu2O nanoparticles. Chemosphere 164, 396–403.
Scalmani, G., Barone, V., Petersson, G.A., Nakatsuji, H., Li, X., Caricato, M.,
Martin, J.M.L., Sundermann, A., 2001. Correlation consistent valence basis sets for use
Marenich, A.V., Bloino, J., Janesko, B.G., Gomperts, R., Mennucci, B., Hratchian, H.
with the Stuttgart-Dresden-Bonn relativistic effective core potentials: the atoms Ga-
P., Ortiz, J.V., Izmaylov, A.F., Sonnenberg, J.L., Williams, Ding, F., Lipparini, F.,
Kr and In-Xe. J. Chem. Phys. 114 (8), 3408–3420.
Egidi, F., Goings, J., Peng, B., Petrone, A., Henderson, T., Ranasinghe, D.,
Martinez, J.M., Volzone, C., Garrido, L.B., 2017. Evaluation of polymeric Al-modified
Zakrzewski, V.G., Gao, J., Rega, N., Zheng, G., Liang, W., Hada, M., Ehara, M.,
bentonite for its potential application as ceramic coating. Appl. Clay Sci. 149, 20–27.
Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O.,
Qu, G., Han, Y., Qi, J., Xing, X., Hou, M., Sun, Y., Wang, X., Sun, G., 2021. Rapid iodine
Nakai, H., Vreven, T., Throssell, K., Montgomery, J.A., Peralta, J.E., Ogliaro, F.,
capture from radioactive wastewater by green and low-cost biomass waste derived
Bearpark, M.J., Heyd, J.J., Brothers, E.N., Kudin, K.N., Staroverov, V.N., Keith, T.A.,
porous silicon–carbon composite. RSC Adv. 11 (9), 5268–5275.
Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A.P., Burant, J.C., Iyengar, S.
Riebe, B., Dultz, S., Bunnenberg, C., 2005. Temperature effects on iodine adsorption on
S., Tomasi, J., Cossi, M., Millam, J.M., Klene, M., Adamo, C., Cammi, R., Ochterski, J.
organo-clay minerals. Appl. Clay Sci. 28 (1–4), 9–16.
W., Martin, R.L., Morokuma, K., Farkas, O., Foresman, J.B., Fox, D.J., 2016.
Sánchez-Polo, M., Rivera-Utrilla, J., Salhi, E., Von Gunten, U., 2007. Ag-doped carbon
Gaussian 09 Rev. D.01. Gaussian Inc., Wallingford.
aerogels for removing halide ions in water treatment. Water Res. 41 (5), 1031–1037.
Fujiwara, H., 2016. Observation of radioactive iodine (131I, 129I) in cropland soil after the
Sawka, A.M., Thabane, L., Parlea, L., Ibrahim-Zada, I., Tsang, R.W., Brierley, J.D.,
Fukushima nuclear accident. Sci. Total Environ. 566–567, 1432–1439.
Straus, S., Ezzat, S., Goldstein, D.P., 2009. Second primary malignancy risk after
Gao, Y., Guo, Y., Zhang, H., 2016. Iron modified bentonite: enhanced adsorption
radioactive iodine treatment for thyroid cancer: a systematic review and meta-
performance for organic pollutant and its regeneration by heterogeneous visible
analysis. Thyroid 19 (5), 451–457.
light photo-Fenton process at circumneutral pH. J. Hazard Mater. 302, 105–113.
Schlichte, K., Kratzke, T., Kaskel, S., 2004. Improved synthesis, thermal stability and
Grimme, S., Antony, J., Ehrlich, S., Krieg, H., 2010. A consistent and accurate ab initio
catalytic properties of the metal-organic framework compound Cu3(BTC)2.
parametrization of density functional dispersion correction (DFT-D) for the 94
Microporous Mesoporous Mater. 73 (1–2), 81–88.
elements H-Pu. J. Chem. Phys. 132 (15), 154104.
Sellaoui, L., Soetaredjo, F.E., Ismadji, S., Benguerba, Y., Dotto, G.L., Bonilla-
Gryta, M., 2013. The concentration of geothermal brines with iodine content by
Petriciolet, A., Rodrigues, A.E., Lamine, A.B., Erto, A., 2018. Equilibrium study of
membrane distillation. Desalination 325, 16–24.
single and binary adsorption of lead and mercury on bentonite-alginate composite:
Han, S., Um, W., Kim, W.S., 2019. Development of bismuth-functionalized graphene
experiments and application of two theoretical approaches. J. Mol. Liq. 253,
oxide to remove radioactive iodine. Dalton Trans. 48 (2), 478–485.
160–168.
Hansen, V., Yi, P., Hou, X., Aldahan, A., Roos, P., Possnert, G., 2011. Iodide and iodate
Seon, J., Hwang, Y., 2021. Cu/Cu2O-immobilized cellulosic filter for enhanced iodide
(129I and 127I) in surface water of the Baltic Sea, Kattegat and Skagerrak. Sci. Total.
removal from water. J. Hazard Mater. 409, 11.
Environ. 412–413, 296–303.
Shen, N., Yang, Z., Liu, S., Dai, X., Yang, C., Li, J., Zhang, Y., Zhang, M., Zhou, R.,
He, L., Chen, L., Dong, X., Zhang, S., Zhang, M., Dai, X., Liu, X., Lin, P., Li, K., Chen, C.,
Chai, Z., Wang, S., 2020. 99TcO4− removal from legacy defense nuclear waste by an
Pan, T., Ma, F., Chen, J., Yuan, M., Zhang, Y., Chen, L., Zhou, R., Han, Y., Chai, Z.,
alkaline-stable 2D cationic metal organic framework. Nat. Commun. 11, 5571.
Wang, S., 2021. A nitrogen-rich covalent organic framework for simultaneous
Shi, K., Ye, Y., Guo, N., Guo, Z., Wu, W., 2013. Evaluation of Se(IV) removal from
dynamic capture of iodine and methyl iodide. Inside Chem. 7 (3), 699–714.
aqueous solution by GMZ Na-bentonite: batch experiment and modeling studies.
Hoskins, J.S., Karanfil, T., Serkiz, S.M., 2002. Removal and sequestration of iodide using
J. Radioanal. Nucl. Chem. 299 (1), 583–589.
silver-impregnated activated carbon. Environ. Sci. Technol. 36 (4), 784–789.

9
J. Yang et al. Chemosphere 292 (2022) 133401

Sun, L., Li, K., Huang, J., Jiang, Z., Huang, Y., Liu, H., Wei, G., Ge, F., Ye, X., Zhang, Y., modified bentonite: batch, column experiment and mechanism investigation. Chem.
Wu, A., Zhijianwu, 2019. Facile synthesis of tri(octyl-decyl) amine-modified biomass Eng. J. 428 (6), 131333.
carbonaceous aerogel for rapid adsorption and removal of iodine ions. Chem. Eng. Yang, J.Q., Shi, K.L., Sun, X.J., Gao, X.Q., Zhang, P., Niu, Z.W., Wu, W.S., 2020b. An
Res. Des. 144, 228–236. approach for the efficient immobilization of Se-79 using Fe-OOH modified GMZ
Theiss, F.L., Ayoko, G.A., Frost, R.L., 2016. Leaching of iodide (I− ) and iodate (IO3− ) bentonite. Radiochim. Acta 108 (2), 113–126.
anions from synthetic layered double hydroxide materials. J. Colloid Interface Sci. Yazdankish, E., Foroughi, M., Azqhandi, M.H.A., 2020. Capture of I(131) from medical-
478, 311–315. based wastewater using the highly effective and recyclable adsorbent of g-C3N4
Vainikka, P., Hupa, M., 2012. Review on bromine in solid fuels. Part 1: natural assembled with Mg-Co-Al-layered double hydroxide. J. Hazard Mater. 389, 122151.
occurrence. Fuel 95 (5), 1–14. Ye, Z., Chen, L., Liu, C., Ning, S., Wang, X., Wei, Y., 2019. The rapid removal of iodide
Wang, G.-Q., Huang, J.-F., Huang, X.-F., Deng, S.-Q., Zheng, S.-R., Cai, S.-L., Fan, J., from aqueous solutions using a silica-based ion-exchange resin. React. Funct. Polym.
Zhang, W.-G., 2021. A hydrolytically stable cage-based metal–organic framework 135, 52–57.
containing two types of building blocks for the adsorption of iodine and dyes. Inorg. Yu, W., Xu, H., Tan, D., Fang, Y., Roden, E.E., Wan, Q., 2020. Adsorption of iodate on
Chem. Front. 8 (4), 1083–1092. nanosized tubular halloysite. Appl. Clay Sci. 184, 105407.
Wang, G., Qafoku, N.P., Szecsody, J.E., Strickland, C.E., Brown, C.F., Freedman, V.L., Zhang, K., Chen, T., 2018. Dried powder of corn stalk as a potential biosorbent for the
2019. Time-dependent iodate and iodide adsorption to Fe oxides. ACS. Earth. Space. removal of iodate from aqueous solution. J. Environ. Radioact. 190–191, 73–80.
Chem. 3 (11), 2415–2420. Zhang, M., Hou, X., 2021. Rapid analysis of 129I in natural water samples using
Warchoł, J., Misaelides, P., Petrus, R., Zamboulis, D., 2006. Preparation and application accelerator mass spectrometry. At. Spectrosc. 42 (3).
of organo-modified zeolitic material in the removal of chromates and iodides. Zhang, S., Xu, C., Creeley, D., Ho, Y.-F., Li, H.-P., Grandbois, R., Schwehr, K.A.,
J. Hazard Mater. 137 (3), 1410–1416. Kaplan, D.I., Yeager, C.M., Wellman, D., Santschi, P.H., 2013. Iodine-129 and iodine-
Xu, S., Zhang, L., Freeman, S.P.H.T., Hou, X., Watanabe, A., Sanderson, D.C.W., 127 speciation in groundwater at the Hanford site, U.S.: iodate incorporation into
Cresswell, A., Yamaguchi, K., 2016. Iodine isotopes in precipitation: four-year time calcite. Environ. Sci. Technol. 47 (17), 9635–9642.
series variations before and after 2011 Fukushima nuclear accident. J. Environ. Zhang, T., Yue, X., Gao, L., Qiu, F., Xu, J., Rong, J., Pan, J., 2017. Hierarchically porous
Radioact. 155–156, 38–45. bismuth oxide/layered double hydroxide composites: preparation, characterization
Yadollahi, M., Hamadi, H., Nobakht, V., 2020. Capture of iodine in solution and vapor and iodine adsorption. J. Clean. Prod. 144, 220–227.
phases by newly synthesized and characterized encapsulated Cu2O nanoparticles Zhong, X., Liang, W., Wang, H., Xue, C., Hu, B., 2021. Aluminum-based metal-organic
into the TMU-17-NH2 MOF. J. Hazard Mater. 399, 122872. frameworks (CAU-1) highly efficient UO22+ and TcO4− ions immobilization from
Yang, J., Shi, K., Gao, X., Hou, X., Wu, W., Shi, W., 2020a. Hexadecylpyridinium (HDPy) aqueous solution. J. Hazard Mater. 407, 124729.
modified bentonite for efficient and selective removal of 99Tc from wastewater. Zhu, L., Sheng, D., Xu, C., Dai, X., Silver, M.A., Li, J., Li, P., Wang, Y., Wang, Y., Chen, L.,
Chem. Eng. J. 382, 122894. Xiao, C., Chen, J., Zhou, R., Zhang, C., Farha, O.K., Chai, Z., Albrecht-Schmitt, T.E.,
Yang, J., Shi, K., Wu, F., Tong, J., Su, Y., Liu, T., He, J., Mocilac, P., Hou, X., Wu, W., Wang, S., 2017. Identifying the recognition site for selective trapping of 99TcO4- in a
Shi, W., 2022. Technetium-99 decontamination from radioactive wastewater by hydrolytically stable and radiation resistant cationic metal-organic framework.
J. Am. Chem. Soc. 139 (42), 14873–14876.

10

You might also like