Yimu
Hu
abc,
Simon
Giret
abc,
Rafael
Meinusch
d,
Jongho
Han
ef,
Frédéric-Georges
Fontaine
*acg,
Freddy
Kleitz
*h and
Dominic
Larivière
*ac
aDepartment of Chemistry, Université Laval, Québec, G1V 0A6, QC, Canada. E-mail: frederic.fontaine@chm.ulaval.ca; dominic.larriviere@chm.ulaval.ca
bCentre de Recherche sur les Matériaux Avancés (CERMA), Université Laval, Québec, G1V 0A6, QC, Canada
cCentre en Catalyse et Chimie Verte (C3V), Université Laval, Québec, G1V 0A6, QC, Canada
dInstitute of Physical Chemistry, Justus-Liebig-University Giessen, 35392 Giessen, Germany
eDepartment of Chemistry, Korea Advanced Institute of Science and Technology (KAIST), Republic of Korea
fCenter for Nanomaterials and Chemical Reactions, Institute for Basic Science, Daejeon 305-701, Republic of Korea
gCanada Research Chair in Green Catalysis and Metal-Free Processes, Canada
hDepartment of Inorganic Chemistry – Functional Materials, Faculty of Chemistry, University of Vienna, 1090 Vienna, Austria. E-mail: freddy.kleitz@univie.ac.at
First published on 26th November 2018
The potential application of thorium (Th) as nuclear fuel, as well as the environmental and public health concerns associated with it, promotes the development of economic and sustainable materials for the separation and removal of Th(IV) from minerals and environmental samples. In this work, centimeter-size, porous silica monoliths exhibiting hierarchical macroporosity–mesoporosity and a robust silica skeleton were prepared using a sol–gel process combined with post-synthetic hydrothermal treatment in ammonium hydroxide. Upon functionalization with diglycolamide (DGA), the resulting monolithic hybrid material was used as a column-type fixed bed sorbent for continuous flow extraction. An enhanced Th(IV) uptake from aqueous solution was achieved with a high enrichment factor and selectivity in the presence of competitive ions such as rare earth elements (REEs) and uranium (U). Systematic mechanistic studies show that the hierarchical pore system is crucial for enhanced adsorption kinetics and capacity. Two mineral leachates were used to assess the performances of the hybrid material, and despite the complex ion matrix and high ionic composition, the sorbent shows highly efficient recovery of Th(IV). The material was able to undergo 10 extraction–stripping–regeneration cycles, which bodes well for potential industrial applications.
Various methods including extraction, adsorption, ion-exchange, and chemical precipitation have been developed to recover actinides from aqueous solutions, among which the multiple-step liquid–liquid extraction (LLE) is the most widely applied technique on the industrial level, owing to its good selectivity and versatility in regard to solvents and organic ligands.4 However, the large amount of organic solvent consumed and the radioactive waste generated during repetitive extraction steps are the main drawbacks in the process. In comparison, alternative methods based on solid-phase extraction (SPE), which features high enrichment factor, fast kinetics, and significant decrease in solvent consumption, have recently emerged.5–7 A remarkable number of materials and nanomaterials have been developed for actinide and nuclide removal from aqueous solution. Traditional porous materials used for this purpose include clay, bare silica gel, active carbon, and zeolites, but they all suffer from low adsorption capacities and slow kinetics.8–11 Advanced functional porous materials, such as amorphous porous organic polymers (POP),12,13 porous aromatic frameworks (PAF),14–16 ion-exchangers,17,18 metal–organic frameworks (MOF),19,20 and covalent organic frameworks (COF)21,22 have been designed to adapt to various conditions (e.g., pH, ionic strength, and presence of interfering metal ions) for nuclear waste removal. Other promising candidates for stationary phase materials in SPE systems include mesoporous silica and carbon, with a pore size between 2–50 nm. These materials have found wide applications for the capture of important radionuclides (e.g., U and Th) and REEs.6 These materials exhibit a remarkably high specific surface area allowing for high contact efficiency and enhanced mass transfer kinetics between the two hydrophilic phases (sorbent and analyte), as well as easily functionalizable surface properties through post-synthetic grafting23–32 or ion-imprinting technique.33–35
In general, mesoporous materials are suitable as stationary phases in chromatographic applications, as the nano-size pores can significantly enhance the extraction efficiency and reduce the analysis time. However, the use of small size particles as packing materials is often associated with a high backpressure of the column, thus reducing mass transport kinetics and limiting their industrial applicability, especially in high flow-rate chromatography analysis. Under this context, bimodal, hierarchically structured silica monoliths that contain both macropores (pore size > 50 nm) and mesopores are highly desirable. The macroporosity will enable the rapid permeation of the aqueous phase, whereas the mesopores provide high specific surface area and active sites. Ever since the first porous silica monolith prepared by Nakanishi et al. in 1991,36 monolithic silica materials with hierarchical porosity have been synthesized based on the similar sol–gel process using a myriad of templates, including Pluronic-type triblock co-polymers,37 cationic,38 and anionic39 surfactants. Among various commonly known methods relying on the mechanism of sol–gel chemistry (e.g., phase separation, emulsions and foams, and ice templating),40 polymerization induced phase separation is a versatile approach towards hierarchically macroporous–mesoporous networks. During the process, the mesoporosity, the size and shape of the macropores, and the degree of macroscopic phase separation can be easily adjusted by factors such as pH, concentration of the polymer, and gelation time of the sol. Another benefit associated with the use of sol–gel route is that the resulting monoliths can be prepared in a variety of morphologies or shapes (e.g., columns, disks, and capillaries), which is particularly interesting for industrial applications.41,42
Novel extraction chromatographic sorbents would greatly improve the ability to separate and recover actinides, and the overall performance of these sorbents is intimately reliant on functional groups. Nowadays, the most commonly used functional groups are phosphorus-, amine- or sulfoxide-based ligands, inspired by organic extracting agents utilized in LLE.4,6 In particular, derivatives of the diglycolyl amide (DGA) ligands have recently been studied for extraction of REEs;6,32,43 however, studies focusing on their ability for actinide extraction in SPE remain scarce.26 In this work, monolithic silica exhibiting a bimodal, hierarchical macroporous–mesoporous structure was successfully prepared via sol–gel processing. A highly stable silica skeleton was obtained through post-gelation treatment in mildly basic ammonia solution using Ostwald ripening inside the silica backbone. The centimeter-size monolithic body exhibits interconnected and adjustable macropores in the micrometer range in addition to disordered mesopores with pore diameters in the 2–4 nm range. Upon functionalization by DGA, the monolithic sorbent with a high specific surface area demonstrated significant selectivity and extraction capacity toward Th(IV) from aqueous solution in both continuous column extraction and batch conditions. The hierarchical nanostructures considerably accelerate the mass transfer, significantly reduce the backpressure, efficiently promote the full utilization of active sites, and markedly improve the kinetic performance. The applicability of sorbents was demonstrated by the removal of Th(IV) from two mineral leachates: rare earth elements ore (OKA-2) and bauxite residue (red mud). The possibility of regenerating the porous sorbents was demonstrated over up to 10 cycles with no significant loss in Th(IV) extraction capacity, suggesting their potential for industrial applications.
During the synthesis, the size of the mesopores was controlled by the pH (i.e., the concentration of ammonia) and the temperature of the post-synthesis treatment. This behavior arises from the formation of larger nanoparticles inside the skeleton at higher ammonia concentration or higher temperature, which leads to larger pore sizes. After calcination, the mesoporosity features of the monoliths under various solvothermal treatment conditions were thoroughly investigated by N2 physisorption. The silica skeleton of materials without solvothermal treatment (M0) was weakly condensed and shattered during the drying process. Furthermore, M0 exhibits only a small size mesoporosity after drying and calcination (a pore size of 2.6 nm calculated via the NLDFT adsorption branch analysis, Table S1, ESI†), and thus no obvious hysteresis loop associated with capillary condensation in large mesopores was observed on the N2 physisorption isotherm (Fig. S2†). For all of the monoliths treated with ammonia, the size of mesopores increases from 3.7 nm to 4.1 nm at higher ammonia concentration and temperature due to Ostwald ripening process (Fig. S2 and Table S1, ESI†). For M2, for example, a typical type IV N2 adsorption isotherm was obtained with capillary condensation reflecting the structural mesoporosity in the relative pressure range (P/P0) of 0.3–0.4, and the pore size distribution curve showed a sharp peak centered at 4.1 nm with high uniformity (Fig. 2). A slight decrease in specific surface area and pore volume was observed for materials treated with higher temperature or pH, as expected, since harsher solvothermal treatment generates larger pore sizes. The apparent density (i.e., the density of the material excluding the gas-accessible pore volume) was measured from the standard N2 physisorption data as described by Weinberger et al.,45 and was determined to be 1.06 g cm−3. Compared to the values for standard mesoporous silica (e.g., MCM-41, MCM-48, SBA-15, and KIT-6), which vary between 2.37 g cm−3 to 2.61 g cm−3, the significantly lower apparent density of monoliths could be explained by the presence of macropores in the structure (vide infra).
The high-resolution scanning electron microscopy (HRSEM) images highlight the macropores that extend throughout the silica skeleton, and the mesopores that are highly accessible through the macropores (Fig. 1b and c). The transmission electron microscopy (TEM) image of pristine silica monolith M2 shows that the mesopores are disordered and buried in the skeleton (Fig. 1d). The macroporosity parameters of the materials were determined using mercury intrusion porosimetry, and the derived pore size distribution (PSD), pore volume, and apparent and bulk density are compiled in Tables 1 and S1 (ESI†). Mercury porosimetry of the monoliths reveals two intrusion steps: one at lower pressure (<1 MPa) in the domain of micrometer-range macropores and the second one at higher pressure (>30 MPa) in the domain of nanometer-range mesopores (Fig. S4, ESI†). Moreover, for M2, the mercury extrusion branch parallels the intrusion branch, suggesting their good mechanical properties. Well-defined and uniform macropores were obtained through concurrent phase separation and sol–gel transition induced by the polymerization reaction, as confirmed by the very sharp step in the mercury intrusion branch in the micrometer domains and the narrow PSD (Fig. 2c and S4, ESI†). The bimodal porosity of the monoliths was clearly observable from mercury porosimetry analysis; however, the sizes of mesopores obtained from mercury intrusion branch are slightly higher compared to the pore sizes calculated via NLDFT method from N2 physisorption. In general, the pore size derived from mercury porosimetry significantly underestimates the real pore size of narrow mesopores (e.g., for ordered mesoporous silica KIT-6 with 3-dimensional structure).46,47 In this case, although solvothermal treatment significantly enhanced the overall mechanical property and robustness of the silica skeleton, part of the 2-dimensional, MCM-41-type mesopore structure collapsed at high pressure during the mercury intrusion,48 and the pore size is therefore overestimated in the nanometer-range. The apparent density obtained from mercury porosimetry for M2 is 1.32 g cm−3, which is in good accordance with the value measured using N2 physisorption. The significant difference between apparent density and bulk density (0.22 g cm−3) again confirms the highly porous nature of the material.
Samples | N2 physisorption | Hg porosimetry | Bulk density (g cm−3) | |||||
---|---|---|---|---|---|---|---|---|
S BET (m2 g−1) | d meso (nm) | V mesopore (cm3 g−1) | Apparent density (g cm−3) | d macro (μm) | V macropore (cm3 g−1) | Apparent density (g cm−3) | ||
M2 | 621 | 4.1 | 1.36 | 1.06 | 1.65 | 3.87 | 1.32 | 0.22 |
M-DGA | 485 | 3.3 | 0.89 | 1.61 | 1.47 | 2.91 | 1.89 | 0.32 |
The DGA-APTS ligand (see Experimental section) was introduced onto the silica monolith piece using a standard grafting procedure in toluene at 120 °C. Prior to adding the DGA-APTS, the pristine monolith was immersed in dry toluene to ensure the penetration of the ligand into the monolith body and homogeneous grafting inside the pores. After the grafting procedure, the monolithic body became slightly yellow (Fig. 1e). The efficiency of the grafting was evaluated using thermogravimetric analysis–differential thermal analysis (TGA–DTA), which showed a total mass loss of 14.3% (Fig. S5, ESI†). Elemental analysis showed 10.4 wt% of carbon and 1.52 wt% of nitrogen in the functionalized monolith (denoted as M-DGA), corresponding to an apparent surface ligand density of 0.45 nm−2. After anchoring of the organic ligands on the silica surface, the shape of the hysteresis loop was well-maintained compared to the pristine monolithic support, and the loop was shifted to the lower relative pressure range (P/P0) of 0.2–0.3, indicating a decrease in pore size. A surface area of 485 m2 g−1 and a pore volume of 0.89 cm3 g−1 were obtained, and a narrow PSD with a mean value of 3.3 nm could be calculated from the adsorption branch based on the NLDFT model (Fig. 2a and b). The overall hierarchically macroporous–mesoporous structure was also well maintained after functionalization, as shown in the mercury intrusion/extrusion curves (Fig. S4, ESI†) and narrow macropore size distribution of M-DGA (Fig. 2d). Furthermore, the HRSEM images (Fig. 1f and g) and TEM (Fig. 1h) of M-DGA also confirm that the hierarchical structure remains essentially intact and that the mesopores, as well as the functional groups, are easily accessible.
The grafting through covalent bonds is further confirmed by solid state NMR spectroscopy. The peaks corresponding to the DGA ligand can be clearly attributed in the 13C CP/MAS NMR spectrum, as reported previously (Fig. S6, ESI†).43,49 From the 29Si MAS NMR spectra, peaks representing trifunctional (T) silicon between −50 to −65 ppm are observed in addition to the Q3 and Q4 peaks of the pristine silica framework between −90 and −120 ppm (Fig. S7, ESI†). The DGA has been anchored to the silica support mostly through T1((SiO)(OR)2Si–R) and T2((SiO)2(OR)Si–R) species appearing at −52 and −60 ppm, respectively, whereas a portion of the ligands is attached to the monolith surface through T3((SiO)3Si–R) at −65 ppm. In addition, the FTIR spectrum of M-DGA shows new peaks at 1672 and 1550 cm−1, which correspond to amide I (CO stretching) and amide II bands (NH deformation, CN stretching), which are not observed in the pristine monolith (Fig. S8, ESI†).
This extraction pattern is intriguing, since the efficient separation of Th(IV) from U(VI) or REE matrices is a difficult task.50–53 In the solution, the species distribution of actinides largely depends on the pH, and the speciation of Th(IV) and U(VI) as a function of pH has been previously studied in detail. For example, at pH 2–4 and Th(IV) concentration < 5 mmol L−1, Th4+, Th(OH)22+ and Th(OH)3+ are the predominant species in solution,24 whereas similar conditions lead to the formation of UO22+, which is in general less favorable for the sorbents compared to its multinuclear hydroxide complexes species formed at higher pH (e.g., (UO2)3(OH)5+ and (UO2)3(OH)7−).25 The size of UO22+ is significantly larger than that of Th(IV) species present in the medium, which renders the chelation between DGA ligand and the U(VI) species less efficient during the dynamic extraction process. On the other hand, although the DGA-based ligands showed high affinity towards middle-size lanthanides (i.e., Sm(III), Eu(III) and Gd(III)) when grafted on ordered mesoporous silica,43,49,54 strikingly high selectivity was observed here for Th(IV) over REEs (Fig. S9, ESI†), possibly because the spatial arrangement of DGA ligands grafted on the silica surface in M-DGA is optimal for Th(IV) extraction.
The substantial amount of radioactive uranium and thorium associated with rare earth deposits has caused considerable concern in rare earth industry, and the separation of these elements from the matrices is therefore of paramount importance for efficient rare earth production and radioactive nuclides management. A recent study by Hopkins et al. demonstrated high selectivity of DGA covalently bonded to mesoporous silica (KIT-6-N-DGA) for radioactive thorium and protactinium over uranium.26 In the present work, the system based on hierarchically structured silica monoliths functionalized by DGA shows superior uptake capacity and specific selectivity towards Th(IV) under dynamic flow-through conditions over the competitive ions that commonly co-exist with the element, suggesting promising potential for applications in rare earth or nuclear industries.
For comparison, the pristine monolith M2 showed an elution time of 8.90 min for Th(IV), with an enrichment factor of 1820%, which corresponds to 54.4% of the total amount of Th(IV) initially introduced into the column. Furthermore, Sc(III) was also enriched by 660%, while only trace amounts of REEs were retained. Indeed, the high selectivity of M2 toward Th(IV) and Sc(III) can be rationalized by the large surface area of the pristine material (SBET = 621 m2 g−1) as well as the abundant amount of accessible silanol groups. It is well-known that the exposed silanol groups on the mesoporous silica surface serve as O-donors for the complexation with actinides ions.23,24,55 Recently, unmodified mesoporous silica materials (e.g., KIT-6 and SBA-15) were used as an extracting medium in a SPE process for the selective separation and preconcentration of scandium, in which the specific binding sites for Sc(III) was seemingly related to silanol groups on the surface of the material, although the binding mechanism has yet to be understood.56 In this work, the grafting of DGA on monolith not only increases the adsorption capacity but also provides a better selectivity for Th(IV) over Sc(III). This is quite important especially for certain types of real-world samples (such as the bauxite residue in this work), in which large amount of Sc(III) could be easily recovered by pure silica material after Th(IV) is eliminated from the solution. Further, it is suggested that the macropores in the bimodal hierarchically structured silica monoliths facilitate mass transport, and expose more silanol groups to the metal ions, leading to an enhanced and selective sorption of Th(IV) and Sc(III). In the case of M-DGA, after functionalization with the DGA ligand, the oxygen in the carbonyl group readily binds Th(IV), further increasing the uptake capacity of the material.
To investigate the impact of accessible silanol groups on the Th(IV) uptake, a surface passivation using 1,1,3,3-tetramethyldisilazane (TMDS) was carried out for M2 and M-DGA, yielding a surface covered by –SiH(Me)2 groups.57 The resulting monoliths were denoted as M2-TMDS and M-DGA-TMDS, respectively. An additional mass loss of 3% for M2-TMDS and 1.6% for M-DGA-TMDS was observed on TGA after surface passivation (Fig. S5, ESI†), which also results in a slight decrease in the specific surface area and pore width measured by N2 physisorption (Fig. S2 and Table S1, ESI†). After surface passivation, the M2-TMDS became highly hydrophobic, while M-DGA-TMDS was partially hydrophobic (Fig. S10, ESI†). Both M2-TMDS and M-DGA-TMDS were used as stationary phases for automated extraction under the above-mentioned experimental conditions, and the extraction chromatogram of M2 showed significant decrease in Th(IV) retention (Fig. S11, ESI†). The signal for thorium appeared almost at the same time as Al(III), Fe(III), Sc(III) and U(VI), and reached rapidly to >90% after 40 min. Since a very limited number of silanols remain accessible on the surface, the amount of Th(IV) recovered by (NH4)2C2O4 decreased substantially from 1820% (compared to the original concentration) for M2 to only 280% for M2-TMDS. In comparison, the enrichment factor dropped to 950% for M-DGA-TMDS, compared to 3500% for M-DGA. The –SiH(Me)2 groups on the surface of M-DGA-TMDS not only significantly reduce the number of accessible silanol groups, but also enhance the hydrophobicity of the material, thus prohibiting the contact between Th(IV) ions and the DGA ligand. Therefore, the presence of free silanol groups is essential for the efficient extraction of Th(IV), as they provide active binding sites and maintain the hydrophilicity of the material. Indeed, the effect of hydrophilic groups in enhancing the ion uptake of sorbents has been discussed previously. For example, Shi et al. showed that the introduction of hydrophilic amino groups (–NH2) on phosphonate-functionalized mesoporous silica surface enhances the accessibility of Th(IV) ions into the inner cavity of the sorbent, although amino groups do not contribute to the complexation with Th(IV).24
Fig. 4 Effect of the contact time on the Th(IV) sorption with kinetic model fittings of pseudo-first- and pseudo-second-order models on M2 and M-DGA at room temperature. |
To determine the adsorption process and the sorbent uptake capacity for Th(IV), adsorption experiments were performed for M2 and M-DGA with initial thorium concentrations ranging from 15 to 200 mg L−1 at pH 3.0 (Fig. 5). Linear Langmuir and Freundlich adsorption models were used to fit the experimental data of Th(IV) adsorption. The adsorption isotherm was found to fit better with the Langmuir model with a correlation coefficient of 0.99 for both M2 and M-DGA (Fig. S13, ESI†). The maximum sorption capacity (Qm,cal) was calculated to be 54.3 for M2 and 84.5 mg g−1 for M-DGA, which is in good agreement with experimental values (55.1 and 83.6 mg g−1 for M2 and M-DGA, respectively; Table S3, ESI†). The Langmuir isotherm model suggests a mono-layered and uniform sorption, in which Th(IV) ions are adsorbed by the complexation with the hydroxyl/carbonyl groups on the sorbent surface.
Fig. 5 Sorption isotherms of Th(IV) at 298 K for the M2 and M-DGA sorbents, and the corresponding Langmuir and Freundlich models fitting curves. |
As mentioned previously, the high adsorption capacity of M2 and M-DGA could be attributed to the abundant amount of accessible silanol groups and DGA ligands. To identify active sorption sites, FTIR spectra of M2 and M-DGA before and after sorption of Th(IV) were compared, as shown in Fig. 6. For M2, the absorption band at 978 cm−1, which is assigned to free O–H of silanols, shifted to 960 cm−1 when the material is saturated with Th(IV), while the similar blue shift was also observed for M-DGA in the same region (from 976 to 961 cm−1), indicating the interaction between silanol groups and Th(IV). Moreover, in the region around 1672 cm−1, where the carbonyl group in DGA is identified (Fig. 6c), the peak shifted to 1648 cm−1 after adsorption of Th(IV). However, no change was observed for the peak at 1550 cm−1 corresponding to amide II (NH stretching). These observations suggest that both free silanol groups and carbonyl groups contribute to the uptake of Th(IV), whereas amino groups do not seem to complex with Th(IV).
To further understand the nature of the sorption process and investigate the influence of temperature, the extraction of Th(IV) by M2 and M-DGA was carried out at temperatures ranging from room temperature (298 K) to 348 K. As shown in Fig. 7(a), the adsorption capacity of the sorbents was largely affected by temperature in the chosen range, with higher temperature favoring the Th(IV) uptake. Thermodynamic parameters such as the standard free energy change (ΔG°), standard enthalpy change (ΔH°) and standard entropy change (ΔS°) were calculated by following eqn (1) and (2):
ΔG° = ΔH° − TΔS° | (1) |
lnKd = −ΔH°/RT + ΔS°/T | (2) |
Fig. 7 Th(IV) adsorption by M2 and M-DGA at different temperatures (a), and the linear regression of lnKdvs. 1/T (b). |
Materials | ΔH° (kJ mol−1) | ΔS° (J mol−1 K−1) | ΔG° (kJ mol−1) | |||
---|---|---|---|---|---|---|
298 K | 313 K | 333 K | 348 K | |||
M2 | 21.9 | 119.3 | −13.8 | −15.6 | −17.9 | −19.8 |
M-DGA | 16.5 | 109.7 | −16.1 | −17.7 | −19.9 | −21.6 |
The positive value of ΔS° suggests an increase in randomness at the solid/liquid interface during the adsorption process. Indeed, for most of Th(IV) adsorbents depending on the completion of the chemisorption, the adsorption is an endothermic process with an increase of randomness,27,58,61 with only few examples of being exothermic process with negative entropy value.20 Upon complexation with active sites, the Th(IV) species undergo dehydration and release water molecules in the aqueous phase, increasing the overall randomness. In the Th 4f XPS spectra of the Th(IV) loaded adsorbents (denoted as M2_Th and M-DGA_Th, Fig. S14, ESI†), the peaks at the binding energy (BE) of 335.26 eV (M-DGA_Th)/336.20 eV (M2_Th) and 344.46 eV (M-DGA_Th)/345.41 eV (M2_Th) were assigned to Th 4f7/2 and Th 4f5/2, respectively, providing evidence of Th(IV) being successfully adsorbed. The spectra are very similar for both M2_Th and M-DGA_Th, which are very close to those of ThO2 (334.6 eV and 343.9 eV for Th 4f7/2 and Th 4f5/2, respectively).62
The benefit of a bimodal pore structure was further evidenced when compared with standard mesoporous materials. The batch sorption capacity of Th(IV) at the same initial Th(IV) concentration (120 mg L−1, pH = 3.0) by pure MCM-41, KIT-6, and corresponding sorbents functionalized with DGA are shown in Fig. 8. Although MCM-41 has the largest specific surface area (1410 m2 g−1) and provides the highest silanol density on the surface, its 2-D hexagonal structure and relatively small pore size (4.0 nm) are not ideal for the adsorption (Fig. S3, ESI†). In comparison, the interconnected 3-D cubic structure of pure KIT-6 is more favorable, since more silanol groups become accessible to Th(IV). After functionalization with DGA, the MCM-41-DGA and KIT-6-DGA possess a comparable amount of ligands grafted (Fig. S4, ESI†), and the sorbents showed enhanced extraction capacity compared to the pristine materials. For the monolithic materials, the superior sorption capacity can be attributed to their hierarchical porous 3-D structure, which reduces the risk of pore blocking during functionalization, and allows the full use of binding sites to achieve a high ion uptake, meanwhile promoting mass transport to facilitate solution permeation. Furthermore, the macropores in the monolith M-DGA can significantly reduce the back pressure associated with the column compared to mesoporous silica powders such as MCM-41-DGA (Fig. 8b), although the two materials possess similar mesopore size (3.2–3.3 nm). Therefore, the use of silica monoliths exhibiting a bimodal, hierarchical porosity as stationary phase is of particular interest in dynamic separation systems, as it provides fast kinetics and high adsorption capacities, and overcomes the back pressure issue associated with silica particle-based columns.
A crucial aspect for the application of the materials is their reusability under the extraction conditions. Ten cycles of extraction–stripping–regeneration experiments were performed to assess the stability of the functionalized monolith under dynamic conditions. Briefly, in each cycle, 15 mL of a diluted OKA-2 solution was passed through the column during 30 min. The retained elements were then recovered by a solution of (NH4)2C2O4 (0.05 M, 15 mL) at a flow rate of 1 mL min−1, followed by 15 mL diluted HNO3 (pH 3.0) to remove the remaining oxalate salt and regenerate the column. As shown in Fig. 9b, no significant loss in Th(IV) extraction capacity was observed even after 10 extraction–stripping–regeneration cycles. The majority of adsorbed Th(IV) was recovered by the elution of the oxalate salt (>82%). In general, approximately 93% of Th(IV) was extracted from 150 mL diluted OKA-2 solution, in which 86% of Th(IV) was recovered. The unaccounted amount of Th(IV) is likely to be found in the diluted HNO3 used for column regeneration, as a small portion of oxalate was trapped in the dead volume of the material during column extraction. After reusability assessment, the monolith was washed by water and characterized. The results obtained from N2 physisorption analysis showed an increase in both specific surface area (from 485 to 537 m2 g−1) and pore size (from 3.3 to 3.5 nm, Fig. S15, ESI†), which is probably due to loss (leaching) of some organic moieties tethered to the silica surface. The thermogravimetric analysis confirmed this result, as a reduction in the mass loss from 14.3 to 10.1% in the temperature range 100–700 °C (Fig. S16, ESI†). Furthermore, the 13C CP/MAS NMR spectrum showed the decrease in intensity of the peak at 61 ppm and disappearance of the one at 16 ppm, relative to the fresh material, indicating partial hydrolysis of the grafted ligand under the experimental condition, especially the ethoxysilane groups (Fig. S17, ESI†).31 However, the major peaks corresponding to carbonyl group and structural carbon remained well-defined, indicating that the overall structure of the organic ligand is well preserved.
Further work is in progress to prepare carbon-based, hierarchically structured monolithic sorbents with higher acid tolerance for actinide removal at low pH. Furthermore, novel functional groups (extractants) will be designed to further enhance the extraction performance, reusability and radiation stability, and ultimately, the new SPE systems will be assessed for their applicability in separation of actinides from spent fuel and mineral deposits.
The KIT-6 material was synthesized following the method reported by Kleitz et al.64 Pluronic P123 (9.0 g) was dissolved in distilled water (325.0 g), followed by addition of HCl (37%, 17.4 g) under vigorous stirring. After complete dissolution, n-butanol (BuOH, 9.0 g) was added. The reaction mixture was left under stirring at 35 °C for 6 h, after which TEOS (19.4 g) was added at once to the homogeneous solution. The molar composition of the starting mixture was TEOS/P123/HCl/H2O/BuOH = 1.0/0.017/1.83/195/1.31. The mixture was stirred at 35 °C for 24 h, followed by static aging step at 100 °C for 24 h. The resulting solid product was filtered, and dried for 24 h at 100 °C. The template was removed by extraction in ethanol–HCl mixture, followed by calcination in air at 550 °C for 5 h.
Surface passivation of monolithic silica was carried out for M2 and M-DGA. The monoliths were immersed in 20 mL dry hexane, followed by addition of TMDS (2 mL) into the solvent. The mixture was stirred at room temperature overnight. The unreacted TMDS molecules were removed by Soxhlet extraction in CH2Cl2 for 12 h. The passivated materials were denoted as M2-TMDS and M-DGA-TMDS.
Qe = V/m × (C0 − Cf) | (3) |
Thermodynamic data for M2 and M-DGA were obtained by carrying out the sorption of Th(IV) at different temperatures (298, 313, 333, and 353 K). The corresponding thermodynamic parameters were calculated based on the distribution coefficient Kd (L mg−1, eqn (4)):
Kd = V/m × (C0 − Cf)/Cf | (4) |
The digestion of bauxite residue was performed by a DigiPREP block digestion (SCP Science, Montreal). Three samples of bauxite residue (0.1 g each) were mixed with 10 mL of H2SO4 (pH = 1) in 20 mL beakers covered with watch glass. The mixture was heated to 75 °C for 30 min with the temperature ramp of 2 °C min−1. After cooling to room temperature, the mixture was filtered through a 0.45 μm syringe filter. The resulting solutions were combined and diluted to 500 mL and the pH was adjusted by KOH and HNO3. The original composition in digested OKA-2 and bauxite residue solutions was investigated by ICP-MS/MS and ICP-OES (Optima 3000, Perkin-Elmer), and the corresponding concentrations are listed in the Table S4, ESI†.
The digested bauxite residue and OKA-2 solutions were used to evaluate the capture efficiency of M-DGA for Th(IV) under dynamic conditions, as described above. Briefly, M-DGA (0.89 g) was placed inside the column. Then, 15 mL of mineral sample was passed through the column with a nominal flow rate of 0.5 mL min−1 to ensure sufficient contact time between the column and solution (about 30 min for each aliquot). The adsorbed ions were then stripped by loading 15 mL (NH4)2C2O4 (0.05 mol L−1) at a flow rate of 1 mL min−1. After recovery of retained elements, diluted HNO3 (pH 3.0, 15 mL) was used to remove traces of (NH4)2C2O4 and regenerate the column. The total interval for each cycle is 60 min. The same extraction-recovery-regeneration recycle was repeated 10 times using diluted OKA-2 solution to assess the reusability and stability of the material.
Footnote |
† Electronic supplementary information (ESI) available: 1H and 13C NMR spectra, 13C cross polarization (CP) and 29Si magic-angle spinning (MAS) NMR spectra, FT-IR spectra, plots of TGA–DTA, linear regression and the corresponding parameters of the kinetics and adsorption isotherm experiments, N2 sorption isotherms and mercury porosimetry data of the materials, XPS spectra of the monolithic materials saturated with Th(IV), and extraction chromatograms of the functionalized and passivated monolithic materials. See DOI: 10.1039/c8ta07952h |
This journal is © The Royal Society of Chemistry 2019 |