You are on page 1of 10

Separation and Purification Technology 228 (2019) 115709

Contents lists available at ScienceDirect

Separation and Purification Technology


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

Design of an adsorbent-bearing silica Schiff base ligand for the highly T


efficient removal of uranium and thorium in acidic solutions
Laurence Whitty-Léveilléa,b, Cyril Aumaitreb, Jean-François Morinb, Nicolas Reyniera,b,
Dominic Larivièreb,

a
CanmetMINING, Natural Resources Canada, Ottawa, ON, Canada
b
Département de chimie, Université Laval, Québec, QC, Canada

ARTICLE INFO ABSTRACT

Keywords: This report describes the design and development of functionalized silica particles using tethered Schiff base
Actinides ligands for uranium extraction from rare earth element leach liquors. We evaluated different grafting approaches
Selective separation (both one- and two-sided tethering); and found the maximum U(VI) adsorption capacity for the functionalized
Functionalized silica silica was 95 ± 2 mg g−1. The functionalized particles demonstrated a very high degree of affinity for the
Extraction
uranyl and thorium ions with Kd values at pH 6 exceeding 19,000 and 900,000 mL g−1, respectively. This se-
Rare earth elements
lectivity for these ions is better than most metal ions present in leach liquors, and the extraction kinetics are
faster than pure Schiff base crystals. The Schiff-base-functionalized silica particles also exhibited a high degree of
reusability over five cycles of loading-elution in the conditions tested.

1. Introduction (MeOSalophen), for the selective extraction of U in aqueous media [7].


We recently demonstrated that this compound, in its crystalline form,
Population growth while Earth has a limited availability of re- could be used for the selective extraction of U and Th from leach liquors
sources is motivating the identification and use of unconventional re- [8]. While acceptable extraction capacity (63 ± 3 mg g−1) and a high
serves for industrial metals and minerals. This is especially true when degree of selectivity compared to other reported extractants for U and
social, political, or environmental factors could affect supply security, Th over REE were observed, we had issues related to partial solubility of
resulting in severe global economic repercussions [1]. In 2018, the the Schiff base in the mobile phase, long extraction time, and limited
United States Geological Survey (USGS) published a list of 35 minerals, reusability of the crystals. In this study, we calculated the solubility of
among which uranium is listed, as critical resources for energy sover- ca. 130 mg L−1 for MeOSalophen, and found it to be unsuitable for
eignty [2]. While the International Atomic Energy Agency has predicted multiple extraction cycles without the constant addition of the base.
that conventional uranium reserves are sufficient to sustain power Grafting of the ligand on a solid support has been shown as a viable
generation via nuclear fission at the current exploitation rates for the strategy to overcome bleeding of extractant originating from partial
next 80–120 years [3], reliance on unconventional reserves (secondary solubilisation of the ligand in the mobile phase [9]. Among the avail-
resources) and the prospective of thorium-based nuclear fuel cycle as able grafting solid supports, silica gel has exhibited high thermal,
the next generation of nuclear reactors could extend this projected chemical and mechanical stability for metal ion preconcentration
period [4,5]. An example of this type of unconventional reserves is the [10,11]. A number of publications have highlighted that functionali-
rare earth element (REE) concentrates obtained from ore deposits. zation of the silica surface using organic moieties can enhance the ex-
These concentrates can contain significant levels of uranium (up to traction of uranyl ions (UO22+) [12,13,14]. There are many tethering
16%) and thorium (up to 20%) [6]. However, the production of U and approaches to covalently bond silane moieties with hydroxyl groups on
Th from these reserves must be supported by the development of se- the surface of silica, [15,16] so it should be possible to graft the
questration strategies with high degrees of extraction and selectivity for MeOSalophen. Recently, Hu et al. highlighted that the presence of di-
U and Th. clycolamide (DGA) moieties and hydroxyl groups on the surface of
In 2012, Zhang et al. reported the use of a member of the Schiff monolithic silica could enhance the extraction of Th over U with ex-
base family, N,N′-Bis(3-methoxylsalicylidene)-1,2-phenylenediamine traction capacity exceeding 70 mg g−1 [17]. In this publication, a


Corresponding author.
E-mail address: Dominic.larivière@chm.ulaval.ca (D. Larivière).

https://doi.org/10.1016/j.seppur.2019.115709
Received 6 February 2019; Received in revised form 15 May 2019; Accepted 17 June 2019
Available online 21 June 2019
1383-5866/ Crown Copyright © 2019 Published by Elsevier B.V. All rights reserved.
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

tethering strategy to covalently bind both sides of the DGA molecules to CA), whereas anions concentrations were measured using a Dionex ICS-
the silica surface was used in an attempt to enhance selectivity as in- 1600 Ion Chromatography unit (Thermo Fisher, Waltham, MA).
itially demonstrated by Florek et al. for REEs [18].
The objective of this research is to determine if grafting a salophen 2.2. Synthesis of Bis(3-methoxysalicylaldehyde) o-phenylenediimine
Schiff base on a silica support could overcome the limitations men- (MeOSalophen)
tioned above with respect to uranium extraction. First, we compared
the extractive performances of the two types of chemical grafting 2.2.1. Synthesis of compound 1
strategy proposed by Florek et al. on silica support [18]. Then we de- Following a modified procedure from Béreau et al. [21], a suspen-
termined the extraction efficiency of the adsorbent for U(VI) in a solid- sion of 5-methoxy-4-hydroxybenzoic acid (2.70 g, 16.1 mmol) was
liquid system on synthetic solutions and acidic REE leach liquors, and prepared in trifluoroacetic acid (15 mL) under a nitrogen atmosphere
adsorption capacity (Qe) and distribution coefficient (Kd). Finally, we (Scheme 1). A solution of hexamethylenetetramine (4.63 g, 33.0 mmol)
calculated the thermodynamic parameters for the extractive processes, in trifluoroacetic acid (25 mL) was added dropwise to this suspension.
and assessed the reusability of the material. After the addition was completed, the mixture was heated at reflux for
24 h. After cooling to room temperature, a 3 M HCl solution (200 mL)
2. Experimental was added to the clear orange solution, from which a beige solid pre-
cipitated. The solid was filtered, washed with water, and air-dried,
2.1. Preparation of surrogate and samples yielding 1.68 g of compound 1 (54%). 1H NMR (500 MHz, DMSO‑d6) δ
10.99 (m, 1H), 10.29 (s, 1H), 7.87 (d, J = 2.0 Hz, 1H), 7.62 (d,
Nanopure water obtained using a Milli-Q system (Millipore, J = 2.0 Hz, 1H), 3.90 (s, 3H). 13C NMR (126 MHz, DMSO‑d6) δ 191.35,
Bedford, MA) was used to prepare diluted solutions. A uranium stock 166.99, 154.78, 148.79, 122.47, 122.38, 122.03, 116.78, 56.62. HRMS
solution of 1000 mg L−1 was prepared by dissolving the appropriate (ESI–): C9H8O5 (M-H)− 196.03723, found 196.03717.
amount of UO2(NO3)2·6H2O (IBI Labs, Boca Raton, FL) in nanopure
water (18.2 MΩ·cm at 25 °C). High purity (> 99%) Na2SO4, NaCl, or 2.2.2. Synthesis of compound 2
NaNO3, purchased from Sigma-Aldrich (Saint-Louis, MO), were used to A solution of compound 1 (1.0 g, 5.1 mmol) in 12 mL of ethanol was
modify the nature of the aqueous phase in order to achieve a final mixed with a solution of o-phenylenediamine (0.275 g, 2.55 mmol) in
anionic concentration of 4000 mg L−1, mimicking the common con- 12 mL of ethanol. The mixture was stirred, and after about 10 min, a
centrations and media found in the leaching liquors used with REEs. A bright orange product precipitated. The mixture was stirred for another
multi-element standard containing U, Th and REEs at a concentration of 3 h to ensure the completion of the reaction. The solid was filtered,
100 mg L−1 (Inorganic Ventures, Christiansburg, VA) was used to washed with ethanol, and then air-dried in an oven at 40 °C for 12 h,
prepare synthetic solutions. Surrogate solutions were prepared by di- yielding 0.75 g (63%). 1H NMR (500 MHz, DMSO‑d6) δ 13.85 ppm (s,
luting the standard solution with nanopure water. Reagents and sol- 2H), 12.82 ppm (br s, 2H), 9.06 (s, 2H), 7.96 (d, J = 1.9 Hz, 2H), 7.54 –
vents used for the synthesis of the MeOSalophen were obtained from 7.52 (m, 4H), 7.45 – 7.42 (m, 2H), 3.85 (s, 6H). 13C NMR (126 MHz,
Sigma-Aldrich (Millipore Sigma, St. Louis, MO) and used without fur- DMSO‑d6) δ 167.22, 164.68, 156.11, 148.47, 141.44, 128.69, 126.98,
ther purification. An REE leach liquor solution in H2SO4 was provided 121.36, 120.40, 118.72, 115.01, 56.13. HRMS (ESI+): C24H20N2O8 (M
by Search Minerals (Vancouver, BC) and used as received [19]. An +H)+ 464.12271, found 464.12197.
analog leach liquor of a mixture of monazite, bastnasite and xenotime
minerals was also produced in HCl according to a leaching strategy 2.3. Grafting approaches for the modified silica supports
developed recently in our research group on Canadian ores [20]. The
composition of the leach liquors are presented in Table 1. Metals and The synthesis method for the modified silica supports is depicted in
actinides concentrations in H2SO4 and HCl leach liquors were de- Scheme 2. Functionalization of silica beads (SiliaFlash P60, SiliCycle
termined using a 725 ICP-OES (Agilent Technologies, Santa Barbara, Inc, Quebec City, Qc) with the MeOSalophen was performed according
to a procedure developed by Glossop and Lane in dry dimethylforma-
mide (DMF) under reflux and an N2 atmosphere [22]. Surface mod-
Table 1 ification was performed in either one-step, or two-step (step-by-step)
Composition of the major anions and cations present in the leach liquor. sequence. This yielded samples designated as SiO2-MeO-1 and SiO2-
Component H2SO4 leachate solution HCl leachate solution MeO-2, respectively.
(mg L−1) (mg L−1)
2.3.1. One-Step modification procedure (SiO2-MeO-1)
Cl− 28 ± 1 93,600 ± 6500
For the one-step modification of silica, a MeOsalophen derivative
NO3– 5.6 ± 0.3 300 ± 20
SO42− 11,800 ± 600 2200 ± 100
with aminosilane was first prepared (Scheme 3). The synthesis, inspired
PO43− < 0.90 < 0.90 by the work of Florek and coworkers [18], is performed as follows:
U(VI) 2.2 ± 0.1 0.93 ± 0.04
100 mg (0.21 mmol) of compound 2, 0.10 mL (0.44 mmol) of (3-ami-
Th(IV) 13 ± 1 58 ± 2 nopropyl)triethoxysilane (APTS) and 67 mg (0.44 mmol) of hydro-
xybenzotriazole (HOBt) were added to 5 mL of dry dimethylformamide
Si(IV) 220 ± 10 58 ± 2
Ca(II) 160 ± 60 820 ± 20 in an inert atmosphere (N2) at reflux conditions. Then, 84 mg
Mn(II) 140 ± 1 165 ± 6 (0.44 mmol) of N-(3-dimethylaminopropyl)-N′-ethylcarbodiimide hy-
Fe(III) 1020 ± 10 1650 ± 60 drochloride (EDC) was dissolved in 2 mL of dry DMF and added at once
Ni(II) 5.6 ± 0.9 54 ± 2
to the solution of compound 2. N,N-Diisopropylethylamine (i-Pr2NEt)
Cu(II) 3700 ± 230 5.3 ± 0.2
Zn(II) 105 ± 2 30 ± 1
was lastly added to the reaction (0.18 mL). The resulting mixture was
Cd(II) 53 ± 4 0.066 ± 0.004 stirred for 24 h at room temperature in a nitrogen atmosphere. This
La(III) 157 ± 4 80 ± 3 mixture containing the modified silane was then used without further
Eu(III) 1.25 ± 0.03 1.40 ± 0.07 purification.
Dy(III) 17 ± 1 14.2 ± 0.6
For the functionalization of the silica surface, 300 mg of activated
Pb(II) 2.0 ± 0.1 25 ± 1
silica material (treated overnight at 150 °C under vacuum) was dis-
pH = 1.82 ± 0.03. persed in 10 mL of dry DMF, in an inert N2 atmosphere. Then, also in an
inert atmosphere, a catalytic amount of i-Pr2NEt and the above solution

2
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

Scheme 1. Detailed synthesis of compound 2.

of the MeOSalophen modified silane was added at once to the suspen- functionalized silica material by filtration using a 0.45 μm filter (Cole-
sion. The resulting mixture was stirred in reflux conditions for 24 h. Parmer, Vernon Hills, IL). Method blanks were performed by replicating
After cooling to room temperature, the suspended solid product was the experiment with and without unfunctionalized or functionalized
filtered, washed thrice with toluene and water, and then dried at 70 °C silica. The initial and the equilibrium concentration of ions of interest in
overnight in air. the supernatant were determined by ICP-OES.
The adsorption capacity Qe (mg·g−1) was calculated according to
2.3.2. Two-Step (Step-by-Step) functionalization (SiO2 -MeO-2) the following equation:
The surface of silica was modified by following a variation of a
(C0 Ce ) V
protocol published by Florek and co-workers [18]. Briefly, 1.0 g of ac- Qe =
(1)
m
tivated silica material (previously heated-treat overnight at 150 °C
−1
under vacuum) was dispersed in 75 mL of dry toluene in an inert N2 where C0 (mg L ) is the ion concentration in the initial solution, Ce
atmosphere. Then, 2.0 mL (9.0 mmol) of (3-aminopropyl) triethox- (mg L−1) is the equilibrium concentration of ion in the supernatant, V
ysilane (APTS) was added at once to the dispersion. The resulting (L) is the volume of the testing solution, and m is the mass of sorbent
mixture was stirred for 24 h in a reflux apparatus. After cooling to room (g). Intrinsic material characteristics such as the distribution coefficient
temperature, the suspended solid product was filtered, washed twice (Kd) of uranium ions between the solid and aqueous phases, and se-
with toluene and ethanol, and then dried at 70 °C overnight. Unreacted paration factors (SFU/X = KdU/KdX) between the uranium and a number
aminosilane molecules were removed by Soxhlet extraction in di- of elements (X) are also critical when comparing solid phase extraction
chloromethane for 6 h. The resulting material was designated as SiO2- techniques [23,24]. Kd were calculated by the following equation:
APTS and used for the subsequent functionalization step, in which Co Ce V
compound 2 was introduced. For this modification step, 300 mg of the Kd =
Ce m (2)
activated SiO2-APTS material (treated overnight at 70 °C under va-
cuum) was mixed with 100 mg (0.21 mmol) of compound 2 and 32 mg A higher Kd value translates into a higher extraction capacity on the
(0.21 mmol) of HOBt in 15 mL of dry DMF in an inert atmosphere at solid sorbent. Indeed, Fryxell et al. reported that Kd values for a specific
reflux. Then, 40 mg (0.21 mmol) of EDC was dissolved in 2 mL of dry element above 500 are considered acceptable, those above 5000 are
dimethylformamide and added at once to the solution of compound 2. considered very good, and Kd values exceeding 50 000 are considered
N,N-Diisopropylethylamine (i-Pr2NEt) was added as a catalyst outstanding [25].
(0.18 mL). The resulting mixture was left stirring for 24 h at reflux.
After cooling to room temperature, the suspended solid was filtered, 2.5. Reusability study
washed thrice with toluene and water, and then dried at 70 °C over-
night. For this study, SiO2-MeO-2 (0.020 g) was added to a surrogate so-
lution of uranium, thorium and rare earth elements in a sulfate-based
2.4. Adsorption experiments solution (4000 mg L−1) for 6 h at 20 °C. Centrifugation was used in
every step of the extraction scheme to separate the supernatant from
The adsorption of uranium from the dilute aqueous solutions was the functionalized silica. The following extraction scheme was per-
performed using 0.020 g of functionalized silica mixed with 20 mL of an formed: SiO2-MeO-2 was rinsed with Milli-Q H2O (10 mL). Then, the
aqueous solution containing the elements of interest in a conical tube. surrogate sample (20 mL, pH 6) was contacted with the functionalized
The mixture was agitated at 20 °C on an oscillator at 230 S min−1. At silica. Finally, U was eluted from the support with (NH4)2C2O4 (1 M,
the end of the adsorption period, the solution was isolated from the 20 mL) over a contact period of 6 h. The choice of the eluent was

Scheme 2. One-step and two-step modification of the surface of silica to generate the uranium-selective sorbent.

3
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

Scheme 3. Synthesis of modified


MeOSalophen for the one-step grafting.

motivated by the work of Lebed and co-workers, who demonstrated the 190 ppm) of the carboxylic acid group in the organic structure, due to
efficient stripping of U(VI) [26] from a phosphonate-functionalized the strong background noise.
mesoporous silica. The supernatant fractions collected during the The covalent attachments of MeOSalophen units and surface func-
loading and the elution steps were analyzed using an ICP-OES. The tionalization were also investigated by 29Si CP/MAS NMR on both
scheme proposed above was repeated five times. samples. The spectra are shown in Fig. 1B. Tetra-functional silicon
centers were named with the conventional Qn notation where n refers
2.6. Material characterization to the numbers of bridging oxygen atoms surrounding the central si-
licon atom. Q3 band (−103 ppm) can be associated to the Si-OH groups
Thermogravimetric analysis (TGA) was performed to determine the in the surface, while Q4 (−113 ppm) peaks are attributed to the func-
loading of the functionalized Schiff base on the silica material using a tionalized silicon atoms with the triethoxysilane anchoring unit [27].
Netzsch STA 449C thermogravimetric analyzer, with an air flow of With the two-step procedure (SiO2-MeO-2) we mostly observe the Q4
20 mL min−1 and a heating rate of 10 °C min−1. The samples were band and a small shoulder for the Q3 band which confirms the appro-
previously dried at 80 °C under vacuum to minimize water adsorption priate functionalization during the APTS treatment with fewer re-
by the materials. Solid-state magic-angle spinning (MAS) NMR (Bruker maining silanol site on the surface. The one-step procedure (SiO2-MeO-
DRX300 MHz spectrometer) was performed to determine the degree of 1) resulted in a NMR spectra with Q3 and Q4 having a similar intensity
tethering of the ligand onto the silica surface, and the integrity of the compared to the native silica. This observation is in accordance with
structure upon grafting. The 75.4 MHz 13C CP-MAS spectra were re- the synthesis method chosen where the surface is most-likely covered
corded using a 7 mm rotor spinning at 8 kHz. The 29Si CP-MAS NMR by amino-propyl groups for the SiO2-MeO-2 material case whereas for
spectra were measured at 59.60 MHz using 7 mm rotor spinning at the SiO2-MeO-1 material, the surface is only locally functionalized with
5 kHz. The chemical shifts are reported in ppm relative to tetra- the modified salophen, leaving most of the surface intact with silanol
methylsilane for 29Si and adamantane for 13C. Liquid state NMR spectra moieties. The 29Si CP/MAS spectrum also shows information regarding
were recorded using a Varian Inova AS400 spectrometer (Varian, Palo the tethering of the organosiloxane species by exhibiting three char-
Alto, USA) at 400 MHz or an Agilent DD2 (Agilent, Santa Clara, CA) for acteristics peaks T1, T2, T3 (respectively −48, −58 and −68 ppm)
the 500 MHz measures. 1H and 13C NMR chemical shifts are referenced [18]. These peaks correspond to the tri-functionalized organosilane and
to residual protons or carbons in deuterated solvent. Attenuated Total indicate the numbers of ethoxy links between the anchoring unit and
Reflectance (ATR) spectra were recorded using a MB3000 spectrometer the surface. In both samples, a strong T3 peak is observed with a smaller
using a MIRacle single bounce ATR diamond kit. Spectra were obtained contribution of the T2 peak. This confirms the existence of strong
from 128 scans with a 4 cm−1 resolution. X-ray photoelectron spec- covalent bonds with two or three ethoxy bonds between the silica
troscopy (XPS) measurements were obtained on a Kratos AXIS-ULTRA surface and the functional moieties.
spectrometer with a monochromatic Al X-ray source operated at 300 W. IR spectroscopy was also used to validate whether or not the
Survey scans were recorded with a passing energy of 160 eV and in- grafting occurred. The presence of a peak (1654 cm−1) associated with
cremental steps of 1 eV. the amide bond I (C]O stretching) on the modified silica (Fig. 2), and
the presence of a CeH bending peak (2158 cm−1) related to the aro-
3. Results and discussion matic carbon confirmed the successful grafting of the ligand on the
surface [28]. In addition, SiO2-MeO-2 also shows the presence of a
3.1. Surface characterization second amide band (NeH deformation, CeN stretching) at 1537 cm−1,
which is in agreement with the MeOSalophen structures of the grafted
To confirm the covalent attachment of the organic species on the species, as suggested in Scheme 2. Peak identification below 1350 cm−1
silica support, 13C and 29Si solid-state NMR, IR, TGA and XPS analyses was impossible due to the presence of a strong spectral background
were performed. 13C CP/MAS-NMR measurements were performed for originating from the silica support itself.
both tethering strategies, in order to assess the efficiency of the che- Thermogravimetric analysis (TGA) was performed to assess the
mical grafting of salophen units with each method. The spectrums ob- degree of grafting achieved through the proposed procedures of the
tained are presented in Fig. 1A. Regardless of the grafting method used, functionalized silica, which indicates a total weight loss ranging be-
the presence of three peaks (12, 23 and 43 ppm) assigned to the me- tween 9% w/w for SiO2-MeO-1 and 17% w/w for SiO2-MeO-2 (Fig. 3).
thylene carbons in the amino-propyl group of the APTS moiety linked to A broad exothermic effect attributed to the decomposition of the
the silica surface are noticeable [18]. The presence of the MeOSalophen MeOSalophen was observed between 290 and 620 °C for SiO2-MeO-2.
unit was confirmed with the presence of 5 peaks attributed to char- As the slope is not as steep for the materials that were functionalized
acteristic carbon atoms of the Schiff base (56, 113–121, 149 and with the one-step grafting procedure, this implies that SiO2-MeO-1 had
168–169 ppm) according to the 13C spectrum of compound 2 in solu- a lower degree of functionalization.
tion. The band at 56 ppm is characteristic of the CH3 carbon linked to X-ray photoelectron spectroscopy (XPS) analysis, which was per-
the methoxy group in addition to the peak at 149 ppm, which was at- formed on the functionalized silica support, also confirmed the pre-
tributed to the aromatic carbon attached to an oxygen atom. The bands sence of MeOSalophen organic moieties on the surface of the SiO2-MeO.
between 100 ppm and 125 ppm are attributed to the aromatic carbons. Furthermore, the intensities of peaks related to the aromatic carbons
Other imine and amide carbons are located at nearly 168 ppm in good (Figs. S1 and S2) detected in the silica-modified materials are higher
agreement with the literature [18]. In both cases, it is difficult to clearly and better resolved on the sample modified through the two-step pro-
identify the presence or the absence of a characteristic peak (ca. cedure (SiO2-MeO-2), than the one-step procedure (SiO2-MeO-1).

4
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

13 29
Fig. 1. Solid state (A) C CP NMR and (B) Si MSA NMR.

8000
SiO2-MeO-1
7000

6000
Distribution coefficient

5000
Kd (mL/g)

4000
SiO2-MeO-2
3000

2150 2050 1950 1850 1750 1650 1550 1450 1350


2000
Wavenumber (cm-1)
1000
Fig. 2. IR spectra of the different functionalized silica materials.
0
SiO₂-OH
SiO₂-O SiO₂-APTS SiO₂-MeO-1 SiO₂-MeO-2

100 Fig. 4. Impact on the distribution coefficient of the different silica materials.
Initial uranium concentration 100 mg L−1, adsorbent mass 20 mg, pH 6, tem-
perature 20 °C, stirring time 24 h.
95
Mass loss (%)

SiO₂-APTS 3.2. Extraction studies


90
SiO₂-MeO-1
SiO₂-MeO-2 The distribution coefficient (Kd) of U(VI) on the pristine silica sup-
85 port, the APTS-functionalized silica, SiO2-MeO-1, and SiO2-MeO-2 were
evaluated (Fig. 4). The results obtained indicate a Kd ranging from 2300
80 to 6700 mL g−1 for those four materials, with SiO2-MeO-1 and SiO2-
40 140 240 340 440 540 640 MeO-2 providing the highest Kd values (4300 mL g−1 and 6700 mL g−1,
Temperature (°C) respectively). These comparable results between the two types of
grafting are in accordance with the TGA analysis, as they have the very
Fig. 3. Thermogravimetric analysis curves of the MeOsalophen-modified silica
similar MeOSalophen loading. Indeed, even though SiO2-MeO-2 shows
samples.
17.7% w/w mass loss, 9.2% w/w could be attributable to the APTS.
This indicates that 8.5% of MeOSalophen was grafted on SiO2-MeO-2,
Similar results were obtained for the amide bonds. Fig. S2 shows results and 9.4% was grafted on SiO2-MeO-1.
of 10.68% COO bonds, for the two-step procedure while only 3.79% The influence of the residual amine groups of the APTS moieties
COO bonds were observed in the one-step functionalizing method (Fig. present on the SiO2-MeO-2 sample, as shown in Scheme 2, may affect
S1). This confirms the presence of the carboxylic acid group for the two- the extraction of U(VI), as described by others [29,30]. It has been
step procedure. suggested that high uranium physisorption is possible at a near-neutral
pH.

5
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

Scheme 4. Coordination of the uranyl ion by the grafted Schiff base ligand.

Distribution coefficient (Kd) Extraction capacity (Qe) Distribution coefficient (Kd) Extraction capacity (Qe)

100 25000 120 25000

100
20000
95 20000
80
Qe (mg/g)

15000

Qe (mg/g)

Kd (mL/g)
90 15000 60

Kd (mL/g)
10000
40
85 10000
5000
20
80 5000
0 0
1 2 3 4 5 6
75 0 pH
No anions SO₄²¯ Cl¯ NO₃¯
Fig. 6. Effect of pH on sorption of uranium by SiO2-MeO-2. Initial uranium
Fig. 5. Effect of the presence of various anions on uranium sorption on SiO2- concentration in sulfate media 100 mg L−1, adsorbent mass 20 mg, temperature
MeO-2. Initial uranium concentration 100 mg L−1, anion concentration 20 °C, stirring time 24 h.
4,000 mg L−1, adsorbent mass 20 mg, pH 6.0, temperature 20 °C, stirring time
24 h.
3.2.2. Effect of the pH
The pH of the aqueous contacted phase with the adsorbent influ-
Computational modeling performed by Brynda et al. [31] suggested
ences the surface charges on the latter, and on the surface metal binding
that the interactions between the Schiff base ligand and the uranyl ion
sites [33]. To evaluate the role of the pH of the aqueous phase on the
in solution occurred at the coordinating center of the Schiff base [31] as
removal of uranyl ions by SiO2-MeO-2, the uptake of uranium at pH
schematized for our functionalized material (Scheme 4). This com-
values ranging from 1 to 6 was investigated (Fig. 6). This pH range was
plexation geometry was confirmed in one of our previous studies when
chosen to be coherent with the typical acidic conditions used in current
the FTIR spectrums of the MeOSalophen before and after the extraction
hydrometallurgical processes and those reported for the extraction of U
of U were compared [8]. A shift in IR signature of the phenolic groups
by functionalized silica [34]. Li and coworkers [35] also reported that,
(from 1265 cm−1 to 1258 cm−1) confirmed that these groups con-
the formation of soluble carbonated complex at pH ≥ 7, which com-
tributed to the uptake of the uranyl ions with the proposed Schiff base.
bined with the deprotonation of the ligands, could reduce the adsorp-
Following those results, we decided to focus our study on the two-
tion of negatively charged U species [36,37]. Thus, we expect that U
step functionalized silica material, because of its higher U(VI) dis-
(VI) adsorption on the grafted silica beads should be maximized be-
tribution coefficient.
tween pH values of 4 and 6. The results showed that adsorption yield
increased with higher pH values. This pH dependency can be attributed
to the different uranyl species in solution [38,39] and the ionization of
3.2.1. Effect of anion presence
hydroxyl groups on the Schiff base ligands [40]. These results suggest
The impacts of the presence of various counter anions (NO3–, Cl−,
that pH 6 is the most suited for the extraction of U with the proposed
SO42−) on the adsorption of uranium(VI) onto modified silica materials
material.
were evaluated. The medium and the concentrations tested were chosen
as they are representative of leaching strategies commonly used in
hydrometallurgy of REE (Fig. 5). The presence of high concentrations of
nitrate, chloride and sulfate anions had little adverse effect on uranium
sorption. In fact, the extraction of U(VI) ions is slightly enhanced when 120
compared to an extraction performed without any additional anions.
The results obtained show a degree of tolerance of the material to the 100
chemical nature of the counter ions, a welcome feature for industrial
80
applications.
Qe (mg/g)

While the adsorption capacities seem to indicate similar extraction


60
efficiencies in all types of media, the measured Kd values for the
functionalized silica material reveal that U(VI) has a lower retention on 40
the material in presence of chloride ions. This can be explained by the
rigidity of the N2O2 cavity of the grafted MeOSalophen. Rudkevich and 20
coworkers have determined that the UO2-Salophen complex was less
stabilized by anions such as Cl− since the rigid binding pocket is too 0
0 4 8 12 16 20 24
narrow for proper complexation [32]. They highlighted that the Kd
Time (h)
value of chloride-based complexes could be up to 100 times lower than
with larger anions like nitrate, phosphate or sulfate. While the decrease Fig. 7. Adsorption kinetics of U (VI) on SiO2-MeO-2. Initial uranium con-
in measured Kd is not as severe as they reported, the results obtained centration in sulfate media 100 mg L−1, adsorbent mass 20 mg, pH 6.0, tem-
during this investigation tend to support this hypothesis. perature 20 °C.

6
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

Table 2 Ce 1 Ce
= +
Pseudo-second order model regression values for two-step functionalized silica Qe bqmax qmax (4)
support.
where Qe (mg g−1) is the amount of metal ions adsorbed at equilibrium,
Samples Qe exp (mg g−1) Qe cal (mg g−1) r2 k2 (g mg−1 h−1)
qmax (mg g−1) the capacity of adsorbent, and b (L mg−1) is a constant
SiO2-MeO-2 95 97 0.9990 0.015 related to the energy of adsorption. The values of qmax and b can be
obtained by the slope and intercept of the correlation curve (Fig. S5). As
highlighted in Table 3, the Langmuir model predict a theoretical ca-
3.2.3. Kinetic study pacity of uranium sorption of 104 mg g−1, which is close to the Qe
The adsorption kinetics parameters of the functionalized silica ma- obtain experimentally.
terial for the uranyl ion were assessed with a solution containing Modeling of the absorption was also attempted using the Freundlich
100 mg L−1 of U(VI). The values of the adsorbed amount of uranium at equilibrium isotherm. This isotherm describes the multilayer adsorp-
equilibrium (Qe) as a function of the duration of extraction are pre- tion with interactions between adsorbed species. The model assumes
sented in Fig. 7. The adsorption rate is quite rapid: 69 mg of U(VI) is that the energy associated with the adsorption surface is heterogeneous,
adsorbed in the first hour for SiO2-MeO-2. Afterwards, the adsorption and that the adsorption process is proportional to the concentration in
rate decreases at a rate of (0.92 ± 0.05)mg g−1 h−1 until the binding solution [50]. The linear form of this model is:
sites are saturated. These rapid adsorption rates imply that the func-
1
tional groups of MeOSalophen are easily accessible to uranyl ions, and lnQe = lnKF + lnCe
n (5)
that, if capacity is not exceeded, extraction could be performed on a
much shorter time scale than the procedure used during this in- where kF (mg g−1)·(mg L−1)n and n (a dimensionless parameter) are the
vestigation (24 h). constants determining adsorption capacity and adsorption intensity,
To gain better insight into the adsorption process, the adsorption respectively. They can be evaluated from the intercept, and the slope,
data were compared to a pseudo-second-order kinetic model first de- respectively, of the linear plot of ln Qe versus ln Ce (Fig. S6, in the
veloped by Blanchard and coworkers [41]. The linear form of the supporting information) and are reported in Table 3. As detailed by
pseudo-second order equation [42], is mathematically expressed as: Belgacem and coworkers, if 1/n is equal to 1, as observed with this
support, the adsorption is linear. The closer the 1/n value tends towards
t 1 t
= + zero, the more heterogeneous the surface is expected to be [51].
Qt k2 Qe 2 Qe (3)
To assess the goodness of fit of this isotherm, error functions were
where Qe and Qt are, respectively, the adsorbed amount of uranium at calculated in order to facilitate the comparison between the two models
equilibrium and at time t, expressed as mg U g−1, and k2 is the pseudo- tested. The residual root mean square error (RMSE) proposed by
second-order kinetic constant expressed as g·mg−1·h−1. Linear mod- Vijayaraghavan et al. [52], and the Chi-square test established by Ho
eling of the pseudo-second order equation is found in the Supporting et al. were used [53] for that purpose. The RMSE is calculated using the
Information, Fig. S3. following equation:
We also modeled the kinetics of the U(VI) adsorption process using n
1
the pseudo-first-order model [43]. However, the experimental data did RMSE = (Qe, exp Qe, calc )2
n (6)
not fit the model closely as for the pseudo-second-order model (Fig. S4). i=1

Table 2 summarizes the parameters calculated from the regression of t/


The subscripts “exp” and ‘calc” represent the experimental and
Qe; t with a goodness of fit (r2) of 0.999 with similar experimental and
calculated Qe values, while n is the number of times the experimental
calculated equilibrium extraction capacity. This observation indicates
isotherm was replicated. In this study, the value for n is 5. The closer
that the adsorption of U(VI) ions is most likely based on chemisorption
the RMSE value is to zero, the better the fit. The Chi-square test is ex-
processes than physisorption [44].
pressed as:
n
(Qe, exp Qe, calc )2
3.2.4. Adsorption isotherms of uranium 2 =
Qe, exp (7)
The uranium(VI) adsorption isotherms were also measured for the i=1

proposed systems to determine the interaction between the mobile and The model is deemed to be more accurate and valid as the value of
stationary phase at equilibrium. Various adsorption isotherms can be χ2 tends towards zero. Both models tested were subjected to these error
used to model the experimental data [45]. However, the extraction of functions and the results are presented in Table 4. Based on those in-
metallic ions from aqueous solution using various solid supports are dicators, the Langmuir adsorption isotherm demonstrates a better
mostly modeled using the Langmuir isotherm model [46,47]. In this goodness of fit than the Freundlich adsorption isotherm, suggesting that
model, the solid support is assumed to have a limited adsorption ca- the latter model does not adequately describe the adsorption processes
pacity qmax, where each complexation site retains a single metal ion, involved.
and all sites are energetically and sterically independent of the ad- In addition, the goodness of fit of the calculated qmax values with the
sorbed quantity [48]. Thus, the Langmuir isotherm model [49] (Equa- experimental Qe values, as well as the calculated values for the error
tion (4)) was used to model the adsorption process of this support: functions illustrate that the interpretation of the adsorption processes
using the Langmuir adsorption isotherm is appropriate. It can be pos-
Table 3
tulated that the uranyl ions are extracted from the aqueous media to
Parameters determined using Langmuir and Freundlich models.
Langmuir Isotherm Freundlich Isotherm Table 4
Error functions parameters calculated for the Langmuir and Freundlich ad-
Parameter Unit Value Parameter Unit Value sorption isotherm.
R2 — 0.952 R2 — 0.023 Adsorption isotherm RMSE Chi-square
qmax mg g−1 104 KF — 92
b L mg−1 0.791 n — 27 Langmuir 6.24 −0.63
1/n — 0.036 Freundlich 13.7 3.70

7
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

10,000,000
in the conditions tested. However, it is interesting to note that the
SiO₂-OH
1,000,000 presence of hydroxyl groups on the silica surface are providing an ap-
SiO₂-APTS
propriate environment for the complexation of U and Th in the condi-
SiO₂-MeO-2
tions tested as reported in the literature [17,18].
Distribution coefficient

100,000

Based on those data, it is possible to determine a separation factor


10,000
Kd (mL/g)

between U and trivalent REE (SFU/Eu) for the functionalized silica ma-
1,000 terial. The degree of selectivity for U compares favorably to other
functionalized silica supports proposed in the literature for the se-
100
paration of actinides and lanthanides, as summarized in Table 5.
10 Since the proposed functionalized material shows interesting ex-
tractive properties with Th as well (Fig. 8), we evaluated the potential
1 of the adsorbent for selective removal of both Th and U ions. The
La Ce Nd Eu Gd Dy Er Yb Th U
competitive extraction of these elements in the presence of numerous
Fig. 8. Competitive distribution coefficient (Kd) values for the functionalized ions was performed using REE leach liquors produced in sulfuric and
silica bead support for actinides in the presence of REEs. The initial con- hydrochloric media, as presented in Fig. 9. As expected from the
centrations of U and REEs were 1 mg L−1, and the initial concentration of Th complex chemical nature of the leaching solutions, the distribution
was 10 mg L−1, with adsorbent mass 20 mg, pH 6.0, temperature 20 °C, stirring coefficient values for U(VI) decreased for both media (853 mL g−1 and
time 6 h.
362 mL g−1 in sulfuric and hydrochloric media, respectively). This
difference in the Kd values is not unexpected, since we already observed
4000
this behavior in the surrogate solution (Fig. 5). Nevertheless, in the
H₂SO₄ leach liquor
3500 sulfuric leach liquor, the degree of selectivity is preserved for U over
HCl leach liquor
other metals, such as REEs and heavy metals.
3000 Indeed, in a sulfate media, a significant degree of selectivity (70-
Distribution coefficient

2500 fold) for U(VI) over Th is observed. This level of selectivity was not
witnessed during our surrogate investigation. This could be explained
Kd (mL/g)

2000 by the favorable formation of a thorium-neutral complex in a solution


1500 containing a high concentration of sulfate. Below pH 3, the most im-
portant complex of Th(IV) is Th(SO4)2(aq), even in a solution con-
1000 taining various anions such as F−, Cl− and PO43− [57]. As demon-
500
strated by Ang and co-workers, the formation of a bond between Th
(SO4)2 and the ligand is unfavorable because of the neutrality of the
0 complex, whereas the formation of complexes with U(VI) is 5 times
U Th Si Ca Mn Fe Ni Cu Zn Cd La Eu Dy Pb
more favorable [58]. Our experimental results tend to support this
Fig. 9. Competitive adsorption of concurrent ions on Schiff base functionalized observation.
silica in leach liquors. Adsorbent mass 20 mg, pH 1.82, temperature 20 °C, In chloride media, however, Th is more selectively extracted than U,
stirring time 6 h. with an observed Kd value of 3590 mL g−1. This behavior has also been
observed by Panda et al. when they used quadridentate Schiff bases to
form a monolayer that covers the adsorption sites of the functionalized extract Th and U from chloride media [59,60]. They determined that
silica material. the neutral complex formed between UO22+ and Cl− hinders the effi-
cient extraction of U(VI), leading to an incomplete extraction of this
element. Thus, since U(VI) has less affinity for the ligand in the chloride
3.2.5. Selectivity test media, heavy metals can be more easily extracted by the available
The selectivity of the MeOSalophen functionalized adsorbent for U MeOSalophen sites [61]. Nonetheless, those results demonstrate the
(VI) over lanthanides and Th was investigated (Figs. 8 and 9) and the significant selectivity of the salophen-modified silica particle towards
adsorption performance for a specific cation was reported in terms of Kd UO22+ and Th in comparison with other constituents of the leaching
values. In a surrogate solution composed of U, Th, and REEs at pH = 6, solution, even if they are present in much larger concentrations
a high selectivity toward U and Th was observed. This behavior has (Table 1). This experiment also showed that, while lower extraction
already been observed and reported by our group on MeOSalophen capacities were reported in more acidic conditions (Fig. 7), the pro-
crystals [8]. SiO2-MeO-2 shows a Kd value exceeding 900,000 mL g−1 posed functionalized material can effectively extract U(VI) in acidic
and 19,000 mL g−1 for thorium and uranium, respectively, while lower conditions and in the presence of competitive ions.
Kd values were obtained for REEs (< 2000 mL g−1) in the conditions
tested. The silica support also contributes to the extraction with Kd 3.2.6. Reusability test
values ranging from 10 mL g−1 to 90,000 mL g−1. As shown in Fig. 8, It is well-documented that extraction resins where the ligand is only
the addition of APTS groups on the surface of the silica beads does not impregnated on the solid support are generally plagued with issues such
significantly increase the extraction properties compared to bare silica as the lack of reusability because the organic extractant can bleed

Table 5
Comparison of SFU/Eu values for different functionalized silica supports.
Support Functionalizing ligand SFU/Eu Aqueous phase Ref.

KIT-6 N,N,N′,N′-tetra-n-octyldiglycolamide 0.32 n.a. [34]


SiO2 nanoparticles Salicylaldiminepropyltriethoxysilane 1.2 pH = 5.5 [54]
SBA-15 Diethylphosphatoethyltriethoxysilane 4.5 HNO3 1 M [55]
Magnetic mesoporous silica Amidoxime 12 pH = 5 [56]
Silica beads MeOSalophen 14 pH = 6 Present study

n.a. Not Available.

8
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

120 References

100
[1] S. Glöser, L. Tercero Espinoza, C. Gandenberger, M. Faulstich, Raw material criti-
cality in the context of classical risk assessment, Resour. Policy 44 (2015) 35–46.
Uextraction/recovery (%)

80 [2] United States Geological Survey, Interior Releases 2018’s Final List of 35 Minerals
Deemed Critical to U.S. National Security and the Economy. < https://www.usgs.
60 gov/news/interior-releases-2018-s-final-list-35-minerals-deemed-critical-us-
Loading national-security-and > (accessed 9 November 2018).
[3] International Energy Agency, World Energy Outlook 2011, IEA Publications, Paris,
40 Elution ((NH₄)₂C₂O₄, 1M)
2011.
[4] R. Price, J. Blaise, Nuclear fuel resources: enough to last 20, NEA News 34 (2002)
20 10–13.
[5] U.E. Humphrey, M.U. Mayeen, Viability of thorium-based nuclear fuel cycle for the
next generation nuclear reactor: issues and prospects, Renew. Sust. Energ. Rev. 97
0
(2018) 259–275.
1 2 3 4 5
[6] Z. Zhu, Y. Pranolo, C.Y. Cheng, Separation of uranium and thorium from rare earths
Cycle for rare earth production-a review, Miner. Eng. 77 (2015) 185–196.
[7] X. Zhang, C. Jiao, J. Wang, Q. Liu, R. Li, P. Yang, M. Zhang, Removal of uranium
Fig. 10. Reusability tests on SiO2-MeO-2. Loading with initial uranium con- (VI) from aqueous solutions by magnetic Schiff base: kinetic and thermodynamic
centration in sulfate media 100 mg L−1, adsorbent mass 20 mg, pH 6.0, tem- investigation, Chem. Eng. J. 198 (2012) 412–419.
perature 20 °C, stirring time 6 h. [8] L. Whitty-Léveillé, N. Reynier, D. Larivière, Selective removal of uranium from rare
earth leachates via magnetic solid-phase extraction using Schiff base ligands, Ind.
Eng. Chem. Res. 58 (2018) 306–315.
during the extraction and stripping processes [62]. Therefore, grafting [9] F. Ozcan, M. Bayrakcı, Ş. Ertul, Synthesis and preparation of novel magnetite na-
noparticles containing calix [4] arenes with different chelating group towards ur-
the MeOSalophen ligand on a silica support could result in tougher anium anions, J. Macromol. Sci. A 52 (2015) 599–608.
materials. To assess this hypothesis, the reusability of SiO2-MeO-2 was [10] A. Goswami, A.K. Singh, Silica gel functionalized with resacetophenone: synthesis
investigated using the conditions presented in Section 2.5. Based on the of a new chelating matrix and its application as metal ion collector for their flame
atomic absorption spectrometric determination, Anal. Chim. Acta 454 (2002)
results presented in Fig. 10, it is manifest that the functionalized sor- 229–240.
bent can be reused for at least 5 cycles with little degradation of the [11] J.G. Espinola, S.F. Oliveira, W.E. Lemus, A.G. Souza, C. Airoldi, J.C. Moreira,
extraction in the loading and elution phase. These data seem to indicate Chemisorption of CuII and CoII chlorides and β-diketonates on silica gel functio-
nalized with 3-aminopropyltrimethoxysilane, Colloid. Surf. A 166 (2000) 45–50.
a degree of stability for the bond between the MeOSalophen ligand and
[12] K. Venkatesan, V. Sukumaran, M. Antony, P. Vasudeva Rao, Extraction of uranium
the silica surface. The sorbent can be easily reused while maintaining its by amine, amide and benzamide grafted covalently on silica gel, J. Radioanal. Nucl.
extractive properties when eluted with 1 M oxalate solution [26]. These Ch. 260 (2004) 443–450.
data tend to support the reusable and possibly cost-effective character [13] S. Sadeghi, E. Sheikhzadeh, Solid phase extraction using silica gel modified with
murexide for preconcentration of uranium (VI) ions from water samples, J. Hazard.
of the grafted MeOSalophen as solid-phase extractant for the hydro- Mater. 163 (2009) 861–868.
metallurgical operations. [14] Y. Zhao, J. Li, S. Zhang, X. Wang, Amidoxime-functionalized magnetic mesoporous
silica for selective sorption of U (VI), RSC Adv. 4 (2014) 32710–32717.
[15] R. De Palma, S. Peeters, M.J. Van Bael, H. Van den Rul, K. Bonroy, W. Laureyn,
J. Mullens, G. Borghs, G. Maes, Silane ligand exchange to make hydrophobic su-
4. Conclusions perparamagnetic nanoparticles water-dispersible, Chem. Mater. 19 (2007)
1821–1831.
[16] A.K. Tucker-Schwartz, R.A. Farrell, R.L. Garrell, Thiol-ene click reaction as a gen-
Herein, we have presented the design of the first example of Schiff eral route to functional trialkoxysilanes for surface coating applications, J. Am.
base-modified silica particles obtained by controlled ligand grafting, to Chem. Soc. 133 (2011) 11026–11029.
be used for the separation of actinides from lanthanides from leach li- [17] Y. Hu, S. Giret, R. Meinusch, J. Han, F.G. Fontaine, F. Kleitz, D. Larivière, Selective
separation and preconcentration of Th (iv) using organo-functionalized, hier-
quors. One-step and two-step grafting procedures were investigated,
archically porous silica monoliths, J. Mater. Chem. A 7 (2019) 289–302.
and the samples prepared via the two-step method, SiO2-MeO-2, [18] J. Florek, F. Chalifour, F. Bilodeau, D. Larivière, F. Kleitz, Nanostructured hybrid
showed higher extraction performances than its one-step counterpart. It materials for the selective recovery and enrichment of rare earth elements, Adv.
also exhibited acceptable selectivity for Th and U, in the presence of Funct. Mater. 24 (2014) 2668–2676.
[19] S. Alam, H. Kim, N.R. Neelameggham, T. Ouchi, H. Oosterhof, Rare Metal
competitive metals in acidic conditions (pH 1.82). Finally, the effective Technology 2016, John Wiley & Sons, New Jersey, 2016.
reusability of the proposed material was demonstrated through many [20] L. Whitty-Léveillé, N. Reynier, D. Larivière, Rapid and selective leaching of acti-
cycles of adsorption and desorption, suggesting that this material is a nides and rare earth elements from rare earth-bearing minerals and ores,
Hydrometallurgy 177 (2018) 187–196.
promising, sustainable, and green approach for the segregation of Th [21] V. Béreau, H. Bolvin, C. Duhayon, J.P. Sutter, Bi-Compartmental Schiff-Base with
and U from secondary resources. Peripheral Ester Functionalization: Synthesis and Magnetic Behavior of Bimetallic
Zn-Ln Complexes (Ln= Dy, Tb, Gd), Eur. J. Inorg. Chem. 31 (2016) 4988–4995.
[22] P.A. Glossop, C.A.L. Lane, Int. Patent WO2010007561, 2010.
[23] H.V. Lavrov, N.A. Ustynyuk, P.I. Matveev, I.P. Gloriozov, S.S. Zhokhov,
Declaration of Competing Interest M.Y. Alyapyshev, L.I. Tkachenko, I.G. Voronaev, V.A. Babain, S.N. Kalmykov,
Y.A. Ustynyuk, A novel highly selective ligand for separation of actinides and
lanthanides in the nuclear fuel cycle. Experimental verification of the theoretical
The authors declare that they have no competing financial interests.
prediction, Dalton T. 46 (2017) 10926–10934.
[24] J.N. Mathur, M.S. Murali, K.L. Nash, Actinide partitioning-a review. Solvent ex-
traction and ion exchange, Solvent Extr. Ion Exc. 19 (2001) 357–390.
Acknowledgments [25] G.E. Fryxell, Y. Lin, S. Fiskum, J.C. Birnbaum, H. Wu, K. Kemner, S. Kelly, Actinide
sequestration using self-assembled monolayers on mesoporous supports, Environ.
Sci. Techol. 39 (2005) 1324–1331.
The authors would like to thank Yimu Hu for the TGA and surface [26] P.J. Lebed, J.D. Savoie, J. Florek, F. Bilodeau, D. Larivière, F. Kleitz, Large pore
analysis and Pierre Audet for the help with the solid NMR analysis. mesostructured organosilica-phosphonate hybrids as highly efficient and regener-
able sorbents for uranium sequestration, Chem. Mater. 24 (2012) 4166–4176.
Finally, the authors would like to acknowledge the editorial work of [27] M. Colilla, M. Martínez-Carmona, S. Sánchez-Salcedo, M.L. Ruiz-González,
Christa Bedwin on this article. J.M. González-Calbet, M. Vallet-Regí, A novel zwitterionic bioceramic with dual
antibacterial capability, J. Mater. Chem. B 2 (2014) 5639–5651.
[28] J.D. Lunn, D.F. Shantz, Peptide brush-ordered mesoporous silica nanocomposite
materials, Chem. Mater. 21 (2009) 3638–3648.
Appendix A. Supplementary material [29] J.-L. Vivero-Escoto, M. Carboni, C.W. Abney, K.E. deKrafft, W. Lin, Organo-func-
tionalized mesoporous silicas for efficient uranium extraction, Micropor. Mesopor.
Supplementary data to this article can be found online at https:// Mat. 180 (2013) 22–31.
[30] X. Wang, G. Zhu, F. Guo, Removal of uranium (VI) ion from aqueous solution by
doi.org/10.1016/j.seppur.2019.115709.

9
L. Whitty-Léveillé, et al. Separation and Purification Technology 228 (2019) 115709

SBA-15, Ann. Nucl. Energy 56 (2013) 151–157. special focus on the past decade, Chem. Eng. J. 308 (2017) 438–462.
[31] M. Brynda, T.A. Wesolowski, K. Wojciechowski, Theoretical investigation of the [48] G. Limousin, J.P. Gaudet, L. Charlet, S. Szenknect, V. Barthes, M. Krimissa, Sorption
anion binding affinities of the uranyl salophene complexes, J. Phys. Chem. A 108 isotherms: a review on physical bases, modeling and measurement, Appl. Geochem.
(2004) 5091–5099. 22 (2007) 249–275.
[32] D.M. Rudkevich, W. Verboom, Z. Brzozka, M.J. Palys, W.P.R.V. Stauthamer, [49] Z. Aly, V. Luca, Uranium extraction from aqueous solution using dried and pyr-
G.J. van Hummel, S.M. Franken, S. Harkema, J.F.J. Engbersen, D.N. Reinhoudt, olyzed tea and coffee wastes, J. Radioanal. Nucl. Chem. 295 (2013) 889–900.
Functionalized UO2 salenes: neutral receptors for anions, J. Am. Chem. Soc. 116 [50] L. Dolatyari, M.R. Yaftian, S. Rostamnia, Removal of uranium (VI) ions from aqu-
(1994) 4341–4351. eous solutions using Schiff base functionalized SBA–15 mesoporous silica materials,
[33] Y. Liu, Q. Li, X. Cao, Y. Wang, X. Jiang, M. Li, M. Hua, Z. Zhang, Removal of J. Environ. Manage. 169 (2012) 8–17.
uranium (VI) from aqueous solutions by CMK-3 and its polymer composite, Appl. [51] A. Belgacem, R. Rebiai, H. Hadoun, S. Khemaissia, M. Belmedani, The removal of
Surf. Sci. 285 (2013) 258–266. uranium (VI) from aqueous solutions onto activated carbon developed from grinded
[34] J. Florek, S. Giret, E. Juère, D. Lariviere, F. Kleitz, Functionalization of mesoporous used tire, Environ. Sci. Pollut. Res. 21 (2014) 684–694.
materials for lanthanide and actinide extraction, Dalton T. 45 (2016) 14832–14854. [52] K. Vijayaraghavan, T.V.N. Padmesh, K. Palanivelu, M. Velan, Biosorption of Ni(II)
[35] Z. Li, F. Chen, L. Yuan, Y. Liu, Y. Zhao, Z. Chai, W. Shi, Uranium (VI) adsorption on onto Sargassum wightii: application of twoparameter and three-parameter isotherm
graphene oxide nanosheets from aqueous solutions, Chem. Eng. J. 210 (2012) model, J. Hazard. Mater. B 133 (2006) 304–308.
539–546. [53] Y.S. Ho, W.T. Chiu, C.C. Wang, Regression analysis for the adsorption isotherm of
[36] A.K. Crane, M.J. MacLachlan, Portraits of porosity: porous structures based on basic dye on sugarcane dust, Bioresour. Technol. 96 (2005) 1285–1291.
metal Salen complexes, Eur. J. Inorg. Chem. 1 (2012) 17–30. [54] Z. Shiri-Yekta, M.R. Yaftian, A. Nilchi, Silica nanoparticles modified with a Schiff
[37] P. Zhou, B. Gu, Extraction of oxidized and reduced forms of uranium from con- base ligand: an efficient adsorbent for Th (IV), U (VI) and Eu (III) ions, Korean J.
taminated soils: effects of carbonate concentration and pH, Environ. Sci. Technol. Chem. Eng. 30 (2013) 1644–1651.
39 (2005) 4435–4440. [55] S. Yang, J. Qian, L. Kuang, D. Hua, Ion-imprinted mesoporous silica for selective
[38] Q.H. Fan, L.M. Hao, C.L. Wang, Z. Zheng, C.L. Liu, W.S. Wu, The adsorption be- removal of uranium from highly acidic and radioactive effluent, ACS Appl. Mater.
havior of U (VI) on granite, Environ. Sci.-Proc. Imp. 16 (2014) 534–541. Inter. 9 (2017) 29337–29344.
[39] M.J. Kang, B.E. Han, P.S. Hahn, Precipitation and adsorption of uranium (4) under [56] Y. Zhao, J. Li, S. Zhang, X. Wang, Amidoxime-functionalized magnetic mesoporous
various aqueous conditions, Environ. Eng. Res. 7 (2002) 149–157. silica for selective sorption of U(VI), RSC Adv. 4 (2014) 32710–32717.
[40] L. Dolatyari, M.R. Yaftian, S. Rostamnia, Removal of uranium (VI) ions from aqu- [57] P. Di Bernardo, F. Endrizzi, A. Melchior, Z. Zhang, P.L. Zanonato, L. Rao,
eous solutions using Schiff base functionalized SBA-15 mesoporous silica materials, Complexation of Th (IV) with sulfate in aqueous solution at 10–70 °C, J. Chem.
J. Environ. Manage. 169 (2016) 8–17. Thermodyn. 116 (2018) 273–278.
[41] G. Blanchard, M. Maunaye, G. Martin, Removal of heavy metals from waters by [58] K.L. Ang, D. Li, A.N. Nikoloski, The effectiveness of ion exchange resins in separ-
means of natural zeolites, Water Res. 8 (1984) 1501–1507. ating uranium and thorium from rare earth elements in acidic aqueous sulfate
[42] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process media. Part 2. Chelating resins, Miner. Eng. 123 (2018) 8–15.
Biochem. 34 (1999) 451–465. [59] C.R. Panda, V. Chakravortty, K.C. Dash, Solvent extraction of zirconium (iv) mid
[43] J.P. Simonin, On the comparison of pseudo–first order and pseudo–second order uranium (vi) from chloride media with bidentate schiff bases, Solvent Extr. Ion Exc.
rate laws in the modeling of adsorption kinetics, Chem. Eng. J. 300 (2016) 3 (1985) 857–866.
254–263. [60] C. Panda, V. Chakravortty, K.C. Dash, A quadridentate Schiff base as an extractant
[44] J.J. Pignatello, B. Xing, Mechanisms of slow sorption of organic chemicals to nat- for thorium(IV), uranium(VI) and zirconium(IV), J. Radioanal. Nucl. Ch. 108
ural particles, Environ. Sci. Technol. 30 (1995) 1–11. (1987) 65–75.
[45] K.Y. Foo, B.H. Hameed, Insights into the modeling of adsorption isotherm systems, [61] A.E. Gorden, M.A. DeVore, B.A. Maynard, Coordination chemistry with f-element
Chem. Eng. J. 156 (2010) 2–10. complexes for an improved understanding of factors that contribute to extraction
[46] A. Abbas, A.M. Al-Amer, T. Laoui, M.J. Al-Marri, M.S. Nasser, M. Khraisheh, selectivity, Inorg. Chem. 52 (2012) 3445–3458.
M.A. Atieh, Heavy metal removal from aqueous solution by advanced carbon na- [62] N. Kabay, J.L. Cortina, A. Trochimczuk, M. Streat, Solvent-impregnated resins
notubes: critical review of adsorption applications, Sep. Purif. Technol. 157 (2016) (SIRs)–methods of preparation and their applications, React. Funct. Polym. 70
141–161. (2010) 484–496.
[47] M.K. Uddin, A review on the adsorption of heavy metals by clay minerals, with

10

You might also like