Aline
Dressler
,
Antoine
Leydier
* and
Agnès
Grandjean
CEA, DES, ISEC, DMRC, Univ. Montpellier, Marcoule, France. E-mail: antoine.leydier@cea.fr
First published on 24th April 2024
Graphene nanoplatelets (GNPs) were functionalized with an organic ligand to prepare materials for selective extraction of uranium from acidic solution. The effects of di-2-ethylhexylcarbamoylethylbutyl phosphonate (DEHCEBP) ligand concentration on the structure of the final solids and the effect on its extraction capacity were investigated in materials with 0.3, 0.5, 0.8, 1.0 and 1.2 mmol of organic ligand per gram of solid. Raman spectroscopy and X-ray diffraction analysis confirm that impregnation does not modify the carbon network and interlayer distance of the GNPs. As shown with nitrogen adsorption–desorption experiments, the amidophosphonate ligand first fills the micropores then the mesopores of the support while the ligand concentration increases. Acidic uranium solutions with high sulfate content were used to simulate the composition of ore treatment leaching solutions. Increasing the ligand concentration inside the graphene leads to an increase of the equilibration time in batch extraction experiments. This suggests that the DEHCEBP molecules form multiple layers in the materials containing the highest ligand contents. The results also suggest that the ligands located inside the micropores remain inaccessible for extraction. Maximum extraction capacities of the material with 1.2 mmol g−1 to 0.3 mmol g−1 DEHCEBP ranged respectively from 108 mg to 18 mg of uranium per gram of solid. This indicates the high potential of these functionalized graphene nanoplatelets for solid phase uranium extraction.
We have previously shown that impregnation to functionalize silica is an efficient and easy way to obtain a selective sorbent.4–8 Silica was impregnated by bifunctional amidophosphonate ligands developed for the solvent extraction of uranium from acidic solutions (both phosphate9,10 and sulfate11). These studies show a complexation of uranium by phosphonate groups with various complexes depending on the ligand, on the process and on the acidic solution (sulfuric or phosphoric).12 These studies and other results for strontium13 and CO2 adsorption14,15 show that impregnated materials generally have higher adsorption capacities than grafted ones, presumably because of the absence of conformational constraints. More recently,16 we observed in a series of silica-based uranium extractants impregnated with 10–20 wt% amidophosphonate ligand that a higher ligand concentration had a negative impact on the extraction kinetics.
The maximum extraction capacity is closely linked to the surface area of the support. To increase this extraction capacity, without decreasing the efficiency of the process, we need to use a support with a high specific surface allowing high amidophosphonate impregnation rates. Graphene-based materials have gained a lot of attention in this context because they can easily be modified with specific functional groups.17 Along with their chemical stability18 and high specific surface areas (up to 2630 m2 g−1 in theory),19 these characteristics allow the development of hybrid materials with high extraction capacities. However, robust procedures for the large-scale synthesis of single- or few-layer pure graphene are lacking, and the outstanding performances measured in research laboratories have so far not been achieved using mass-produced materials.20,21
Graphene nanoplatelets are a mixture of single-layer, few-layer, and nanostructured graphite22 and are also known as graphite nanoplatelets. Their specific structure and their low costs when produced in large quantities make them an attractive alternative for adsorption applications.23–25 As such, GNPs have found applications in the absorption of organic pollutants,26–29 solar cells,30 and in various polymer nanocomposites with flame retardant,31 shape memory32 and gas barrier33 properties.
The lateral size and thickness of the flakes, and the density of defects and impurities in GNPs depend on the manufacturing technique,34,35 and their selectivity and extraction capacity can be improved by the nature of the functionalization. Surface functionalization by covalent attachment involves rehybridization of one or more of the sp2 carbons in the network into the sp3 configuration, with a simultaneous loss of electronic conjugation.36 Non-covalent functionalization involves three main interactions including hydrophobic interactions, and van der Waals and electrostatic forces.36–38 For post-synthesis functionalization with organic molecules, GNPs do in general have residual oxygen-containing groups but these are not reactive enough for the grafting of organic groups.37 On the other hand, while non-covalent functionalization can be achieved without functional groups, the molecule must be sufficiently attracted to the graphene surface to remain attached.38
Here, graphene-based adsorbents were developed for uranium extraction during the uranium recovery stage from mines. These operations typically involve first ore leaching and then selective extraction of uranium from the leaching solution followed by a concentration step.39–41 Solid-phase extraction (SPE) is an alternative to solvent extraction processes for the selective recovery of low-concentration uranium (up to 1 g L−1) because of the absence of solvent, lower processing times and lower costs.42,43 Solid-phase extraction is also a more compact process, allowing extraction and back-extraction to be performed in separate locations. The uranium industry already uses solid-phase extractants, as for example the resin-in-pulp process, or using organic resin based on ion exchange. However, in the case of high sulfate concentrations and/or in the presence of a high concentration of competing cations in the leaching solution, solid-phase extraction is less efficient and solvent extraction is preferred. The use of a selective ligand linked to a support is the first way to improve the SPE process. In the literature, numerous materials, based on metal oxide particles,44,45 mesoporous silica,46,47 carbon supports,48–51 MOFs,52–54 fibers55,56 or resins41,57–59 were suggested and evaluated for the removal of uranium ions from aqueous effluents by a SPE process.
In the present study, amidophosphonate ligands were impregnated in commercial GNPs. The non-polar sp2 carbon surface of GNPs allows these ligands to be impregnated without prefunctionalization and their high pore volume translates into a high loading capacity. In SPE, however, the accessibility of the ligand molecules is strongly dependent on their arrangement inside the pores of the materials and the mesostructure of the mineral solid support.60 We therefore studied the effects of the ligand concentration on the structure of the functionalized materials and then we assessed the effect of the support on the material extraction properties.
The characteristics of the ligand are presented in Fig. S2 and S3 (mass spectra) and S4–S6 (NMR spectra) (ESI†).
For the TGA experiments, about 20 mg of the functionalized material was placed in a 70 μL alumina pan and heated from 30 to 1000 °C at 5 °C min−1 under a 30 mL min−1 air flow. The ligand concentrations were calculated using eqn (1):
![]() | (1) |
![]() | (2) |
The ligand concentrations obtained from the TGA analysis, using eqn (1), were equal to the amounts of DEHCEBP added during the synthesis of the respective materials.
Organic contents inside the materials calculated from the amount of DEHCEBP added in the impregnation process are in agreement with the results of the gravimetric and thermogravimetric analysis. The fact that all the ligand added in the preparation of the adsorbents is incorporated into the support confirms that the synthesis route is robust.
Attenuated total reflection Fourier transform infrared spectroscopy (ATR-FTIR) data were acquired using a PerkinElmer Spectrum 100 spectrometer from 615 to 4000 cm−1. The scans were quadrupled with a nominal resolution of 4 cm−1 and background correction (atmospheric bands) for each substrate.
Raman spectra were recorded using a Horiba Jobin Yvon device. The data were collected using a 532 nm laser (Olympus MPlan N 100×/0.90) and a 100× objective lens from 500 to 3000 cm−1 with an integration time of 30 s and averaged over 10 scans.
The median particle size by volume, d(50)v, was determined using a laser diffraction particle size analyzer (Mastersizer 3000, Malvern Panalytical) with the refractive index and absorption set to 2.42 and 1, respectively, using water as the dispersant.62 The experiments were carried out without ultrasonication to avoid altering the size of the GNP particles.
The unmodified GNPs and functionalized materials were observed by field emission scanning electron microscopy (FE-SEM, Quanta 200 ESEM FEG, FEI Company) under high vacuum, with a 1.8 kV accelerating voltage. Samples were dispersed in high purity water by shaking without sonication and drop cast on carbon tape without additional preparation.
The carbon interlayer spacing in the materials was determined by powder X-ray diffraction (XRD) with the Bragg Brentano geometry (PANalytical X'Pert PRO MPD; copper anode λKα1 = 1.54056 Å generated at 45 mA and 40 kV, X'celerator detector). The XRD patterns were collected over the 2θ range of 10–70° with 0.017° steps and a measurement time of 0.625 s per step.
Nitrogen adsorption–desorption isotherms were measured at −196 °C using a Micromeritics ASAP 2020 surface area and pore size analyzer. The samples were degassed at 90 °C for 24 h before analysis. The Brunauer–Emmett–Teller (BET) method was used to calculate the specific surface areas. Total pore volumes were determined using the volume of adsorbed gas at P/P0 ≈ 1. Micropore volumes were taken at P/P0 = 0.05 and mesopore volumes were then deduced by subtracting the micropore volume from the total pore volume. This procedure allows the evaluation of the percentage of the volume of each pore type (micro and meso) occupied by the ligand.
The surface density (dsLigand) of the ligand (DEHCEBP nm−2) was calculated from the BET-specific surface area SBET (m2 g−1) of the pristine support and the organic content (τL, mmol g−1, from eqn (1)) of the final materials using eqn (3):
![]() | (3) |
The area occupied by a single DEHCEBP molecule was estimated using the empirical Tanford formula for the length (r, in nm) of a simple hydrocarbon chain of n atoms:
r = 0.154 + 0.1265n | (4) |
After the chosen contact times, and filtration through a 0.22 μm cellulose acetate membrane, the uranium concentration in the liquid phase was measured by inductively coupled plasma atomic emission spectroscopy (ICP-AES; 2% nitric acid; Analytik Jena PlasmaQuant PQ 9000). The mass or amount of uranium extracted per unit mass of solid, that is the extraction capacity at the chosen duration time (QU(t),) was calculated using eqn (5):
![]() | (5) |
The ligand to uranium molar ratio (L/U) in each material under these experimental conditions was determined using the measured ligand concentrations (τL, mmol g−1) inside each solid support, L/U = τL/QU(t).
If all ligand molecules are accessible for extraction, L/U corresponds to the stoichiometric coefficient of the complexes between the ligand and uranium formed inside the pores of the material during extraction.
The phosphorous concentration in the liquid phase after contact with the synthesized materials was also measured by ICP-AES (2% nitric acid; Analytik Jena PlasmaQuant PQ 9000) to determine the amount of ligand leached during the extraction process at different contact times.
![]() | ||
Fig. 2 Fourier-transform infrared spectra of the unmodified GNPs and the impregnated materials with DEHCEBP concentrations of 0.3–1.2 mmol g−1. |
The Raman spectra of the unmodified GNPs and impregnated materials show the three characteristic peaks of graphene (Fig. 3), which provide information on defects (the D band at ∼1340 cm−1), in-plane vibrations of sp2 carbon atoms (the G band at ∼1570 cm−1) and the stacking order (the 2D band at ∼2700 cm−1).63 The position and shape of the three peaks are sensitive to the arrangement of the graphene layers and the level of charge-doping.64,65 In contrast with the results obtained for the GNP-based materials, monolayer graphene has a sharp 2D peak.66 The intensity ratio of the D and G bands (ID/IG) is often used to estimate the defect concentration in carbon materials.63 The fact that the ID/IG ratio does not change significantly after impregnation is a sign of a very poor interaction between ligand and GNP. Indeed, the synthesis route uses direct impregnation, without chemical reaction between the ligand and the GPN surface. Another sign of weak or non-covalent bonding between the ligand and the GPN surface comes from the fact that contacting the functionalized material in an organic solvent enables the recovery of the entire ligand.
![]() | ||
Fig. 3 Raman spectra normalized to the G peak of unmodified GNPs and the impregnated materials with DEHCEBP concentrations of 0.3–1.2 mmol g−1. |
This functionalization does not significantly alter the structural order of the GNPs,67 unlike what has previously been observed for covalent functionalization.68
Material | τ L | %Lb | d(50)vc | S BET | V pores | V micropores | d sLigand |
---|---|---|---|---|---|---|---|
a Concentration (mmol g−1) of DEHCEBP. b Mass percentage of DEHCEBP. c Median particle size by volume (μm). d Specific surface area (m2 g−1). e Total pore volume, measured at P/P0 ≈ 1 (cm3 g−1). f Micropore volume. Measured at P/P0 ≈ 0.05 (cm3 g−1). g Surface density of the ligand (DEHCEBP nm−2) using eqn (3). | |||||||
GNP | 42 (±1) | 675 (±34) | 1.31 (±0.07) | 0.25 | |||
Imp-0.3@GNP | 0.32 (±0.02) | 13.9 (±0.7) | 43 (±3) | 302 (±15) | 0.75 (±0.04) | 0.09 | 0.29 |
Imp-0.5@GNP | 0.52 (±0.03) | 22.5 (±1.1) | 72 (±3) | 153 (±8) | 0.63 (±0.03) | 0.04 | 0.46 |
Imp-0.8@GNP | 0.80 (±0.04) | 34.7 (±1.7) | 99 (±5) | 60 (±3) | 0.35 (±0.02) | 0.01 | 0.71 |
Imp-1.0@GNP | 1.02 (±0.05) | 44.2 (±2.2) | 127 (±3) | 23 (±1) | 0.17 (±0.01) | 0.01 | 0.91 |
Imp-1.2@GNP | 1.15 (±0.05) | 49.8 (±2.5) | 140 (±9) | 6 (±1) | 0.08 (±0.01) | 0 | 1.03 |
Particle sizes were measured by laser diffraction, a technique usually used to determine the mean lateral size of graphene oxide platelets69 and commercial graphene materials62 with high accuracy, but that is known to underestimate the presence of smaller particles.69
Fig. 4 shows SEM images of pristine GNPs and the impregnated materials. Pristine GNPs form agglomerates, as reported previously.70 Increasing the concentration of the organic ligand leads to the formation of a coating layer around the GNPs, which becomes visible on the micrographs at DEHCEBP concentrations above 0.8 mmol g−1 (Fig. 4d–f). The presence of organic molecules on the outer surface of the GNPs suggests that the ligand may act as a bonding agent between initially separated nanoplatelets.
![]() | ||
Fig. 4 Scanning electron micrographs of (a) unmodified GNPs and (b–f) impregnated GNPs with DEHCEBP concentrations of (b) 0.3, (c) 0.5, (d) 0.8, (e) 1.0 and (f) 1.2 mmol g−1. |
This would explain the increase in the mean particle diameter with the DEHCEBP concentration measured by laser diffraction.
Fig. 5 compares the XRD patterns of the materials. The sharp peak at 2θ = 26.36° in the data from the pristine GNPs corresponds to an interlayer distance (d002) of 0.35 nm, close to that of high purity graphene (∼0.34 nm).71 This peak is also present in the data from the pristine samples but decreases in intensity with the organic content from 0.87 to 0.36 of the intensity observed for pristine GNPs between Imp-0.3 and Imp-1.2@GNPs, respectively (Fig. 5). These data show that impregnation of molecule does not affect the interlayer structure of the graphene nanoplatelets. The XRD patterns of the impregnated samples also show a broad peak at 2θ = 20°, indicating the presence of an amorphous phase.
Fig. 6 shows the nitrogen adsorption–desorption isotherms of the GNP support before and after impregnation with different ligand concentrations. The unmodified GNPs and the Imp-0.3, Imp-0.5 and Imp-0.8@GNPs materials have type II isotherms, with an H3 hysteresis loop, typical of particles with a porous network consisting of macropores and non-rigid aggregates of plate-like particles.72 For Imp-1.0 and Imp-1.2@GNPs, the hysteresis loops are closer to type H4, with type I adsorption branches, typical of solids with relatively small external surface areas.72 This can be explained by pore filling due to the high ligand concentration, and also by aggregation of the nanoplatelets as observed previously.
![]() | ||
Fig. 6 Nitrogen adsorption–desorption isotherms of unmodified GNPs and of the impregnated materials with DEHCEBP concentrations of 0.3–1.2 mmol g−1. |
Since unmodified GNPs have a total pore volume of 1.31 cm3 g−1 and the density of DEHCEBP is about 1.0 g cm−3, the maximum ligand concentration in the pore volume of this support is about 57 wt%. However, samples prepared with more than 50 wt% DEHCEBP exhibited a sticky appearance, indicating that a considerable portion of the ligands is probably localized outside the pores and that the maximum reachable loading using this procedure is about 1.2 mmol g−1 or 50 wt%. Some of the pores in the material, presumably the smaller ones, must therefore be inaccessible to the ligand.
The ligand concentrations, pore volumes and specific surface areas of the studied materials are listed in Table 1. The total pore volume and BET specific surface area decrease after functionalization and follow an opposite trend to the ligand concentration. This suggests that at least some of the ligand molecules are located within the pores.
Fig. 7 shows the percentage of filled micropores and mesopores with respect to the unmodified support for all the synthesized materials. The proportion of filled volume is always higher in the micropore volume than in the mesopore volume. This suggests that the organic ligand tends to fill the micropores before the mesopores in the support, as previously observed for silica-based impregnated materials.16 Furthermore, the fact that the pores in samples Imp-1.0 and Imp-1.2@GNPs are almost completely filled by the ligand (Fig. 7) explains why they have a very low specific surface area (Table 1).
![]() | ||
Fig. 7 Percentage of filled micropore and mesopore volume in impregnated graphene nanoplatelets as a function of the concentration of organic ligand (DEHCEBP) in the final materials. |
The total (residual) pore volume of the impregnated materials decreases linearly with the DEHCEBP concentration, indicating that each ligand molecule occupies a similar volume within the pores of the support, regardless of its concentration (Fig. 8).
![]() | ||
Fig. 8 Residual pore volume of the impregnated materials as a function of the concentration of organic ligand (DEHCEBP), R2 = 0.9943. |
Following the same reasoning based on the Tanford formula (eqn (4)) as previously,16 assuming that the movement of the impregnated ligands is not restricted and that a DEHCEBP molecule is about 2 nm long, each molecule occupies a circular area of about 3.3 nm2. The full saturation of the surface of the solid support with one layer of ligand molecules requires theoretically about 0.3 DEHCEBP nm−2. This means that the ligand molecules are disposed in a single layer on the support surface only for DEHCEBP concentrations up to 0.3 mmol g−1 and form multiple layers at higher concentrations.
In summary, these results suggest that during the impregnation process, at low concentrations, the ligand first partially fills the micropores and forms a monolayer in the mesopores. As the ligand concentration is increased, the remaining accessible micropores are filled and the DEHCEBP molecules tend to form multiple layers in the mesopores, the number of layers increasing with the ligand concentration (Fig. 9).
The diffusive model derivate from Weber and Morris (eqn (7))73,74 was used to determine how the diffusion inside the porous structure impacts the adsorption process. This highlights the effect of the amount of the ligand into the diffusive process and in the extraction kinetics.
qt = kidt1/2 + I | (7) |
In this equation, qt is the amount adsorbed at time t, kid is the intra-particle diffusion rate constant (mg g−1 h−1/2) and I (mg g−1) is a constant related to the thickness of the boundary layer. According to this model, if the relationship between qt and t1/2 is linear and passes through the origin it means that the adsorption process is controlled only by intraparticle diffusion in the liquid phase (inside the pores of the samples). On the other hand, a linear relationship that does not pass through the origin indicates that intraparticle diffusion is involved, but is also indicative of some degree of boundary layer control.75 Finally, a piecemeal linear relationship indicates that the adsorption process is governed by two or more mechanisms.76,77
Fits using this model (eqn (7)) of the kinetics data from the five synthesized materials are shown in Fig. 10.
For materials with DEHCEBP concentrations of 0.8 to 1.2 mmol g−1, three distinct linear sections are observed. Assuming well-mixed conditions, such that external mass transport resistance can be neglected,78 the initial linear section (Section I) corresponds to diffusion in the liquid phase (external and in the pore structure); the second section corresponds to intraparticle or diffusion in the boundary layer (here organic layer); and the third section corresponds to the equilibrium state. Section II is the slowest stage in all three cases and is slower in Imp-1.0 and Imp-1.2@GNPs than in Imp-0.8@GNPs. The relationships for the impregnated materials with DEHCEBP concentrations of 0.3 to 0.5 mmol g−1 only consist of two sections, with no section I due to high total pore volume. Decreasing the total pore volume by increasing the amount of ligand in the sample leads to an increase of time to reach the end of section I.
The intercept of the intraparticle diffusion Section (II), which is proportional to the thickness of the boundary layer,77,79 increases with the ligand content of the materials, as does the equilibration time (Table 2). These trends are consistent with the results presented above indicating that the ligand molecules form multiple layers in the pores of the support at concentrations higher than 0.3 mmol g−1. The characteristic rate of adsorption under intraparticle diffusion, kid, decreases from Imp-0.3 to Imp-0.8@GNPs but is similar for Imp-0.8, Imp-1.0 and Imp-1.2@GNPs (Table 2).
Material | Q Umax | t | I | k id |
---|---|---|---|---|
a Maximum uranium extraction capacity (mg g−1). b Estimated equilibration time. c Y-intercept of the intraparticle diffusion section. d Rate of adsorption in the intraparticle diffusion domain. | ||||
Imp-0.3@GNPs | 18 (±1) | 4 hours | 8.9 | 4.8 |
Imp-0.5@GNPs | 31 (±2) | 4 hours | 24.6 | 3.6 |
Imp-0.8@GNPs | 59 (±3) | 2 days | 50.6 | 1.1 |
Imp-1.0@GNPs | 84 (±4) | 7 days | 65.9 | 1.4 |
Imp-1.2@GNPs | 108 (±5) | 7 days | 90.6 | 1.4 |
Fig. 11 shows that while the uranium extraction capacity of the synthesized materials increases linearly with the ligand concentration, from 18 to 108 mg g−1, the ligand to uranium molar ratio follows the opposite trend. In a previous study, we found that silica-based materials impregnated with DEHCEBP had much lower extraction capacities (from 28 to 54 mg g−1). However they had the same equilibrium ligand to uranium molar ratio (L/U ∼ 2) whatever the ligand content was (0.2–0.5 mmol g−1), suggesting minimal hindrance of the silica support in the extractions mechanisms.16 Here, the higher equilibrium L/U ratios (lower uranium extraction efficiency) at lower ligand concentrations (Fig. 11) point toward a possible chemical or physical interaction between the graphene-based support and the DEHCEBP molecules during extraction processes.
![]() | ||
Fig. 11 Equilibrium ligand to uranium molar ratio (red dots) and uranium extraction capacity (blue dots) of the five impregnated materials as a function of the DEHCEBP concentration. |
Assuming that the amidophosphonate molecules form, in equilibrium, the same complex with ligand to uranium molar ratios close to 2, regardless of the support on which they are impregnated, it is possible to suppose that some of the ligands present in the GNP-based materials remain inaccessible during extraction. The proportion of inaccessible ligands (τIL) was estimated by subtracting the ligand concentration present in each material (τL) by the calculated ligand concentration required to reach L/U = 2 (τ(L/U=2)), keeping the experimentally measured QU values fixed.
Eqn (8) and (9) are used to calculate the ligand concentration required to reach L/U = 2 and the concentration of the inaccessible ligand, respectively.
τ(L/U=2) = QU × 2 | (8) |
τIL = τL − τ(L/U=2) | (9) |
The results obtained for each synthesized GNPs-based material are shown in Table 3.
Material | τ LU | Normalized τLUb |
---|---|---|
a Concentration of DEHCEBP molecules (mmol g−1) not involved in U(VI) extraction. b Concentration of DEHCEBP molecules (mmol g−1) not involved in U(VI) extraction normalized to the fraction of filled pore volume. | ||
Imp-0.3@GNPs | 0.17 (±0.03) | 0.27 (±0.04) |
Imp-0.5@GNPs | 0.25 (±0.03) | 0.30 (±0.04) |
Imp-0.8@GNPs | 0.30 (±0.04) | 0.32 (±0.04) |
Imp-1.0@GNPs | 0.30 (±0.05) | 0.31 (±0.05) |
Imp-1.2@GNPs | 0.25 (±0.05) | 0.25 (±0.05) |
The calculation of the concentration of DEHCEBP molecules (mmol g−1) not involved in U(VI) extraction is the lowest for Imp-0.3 and increases with increasing the concentration of ligand apart from Imp1.2.
There are various hypotheses that can explain the presence of an “unused ligand”. The first one would be the inaccessibility of some of ligands due to chemical interactions involving the lone pair of electrons on the phosphorus atom for instance.67 However as previously explained, the material with the lowest ligand content (0.3) contains enough DEHCEBP to cover the entire surface of the GNPs with a monolayer. This material should have the highest concentration of inaccessible DEHCEBP molecules since it has the highest concentration of CNPs. This is not the case however (Table 3) thus discarding our first hypothesis.
The second hypothesis assumes the leaching of some of the ligand during the extraction experiments. However, the phosphorus concentration in solution remained stable after 4 h of extraction, at 0, 0.8, 1.9, 1.8 and 2.4 wt% of DEHCEBP loss for the Imp-0.3–Imp-1.2@GNPs materials, respectively (Fig. S6, ESI†). This suggests that only molecules that interact weakly with the graphene support are lost, invalidating our second hypothesis.
The third hypothesis is that the amidophosphonate molecules located in the smaller pores (micropores) remain trapped and do not therefore contribute to uranium extraction. Normalizing the concentrations of inaccessible ligands to the percentage of filled micropore volume (Fig. 7) yields similar values for all the synthesized materials (Table 3), indicating that the loss of uranium extraction efficiency in these materials could indeed be due to a portion of the ligands being trapped in the micropores of the GNPs. The increase in extraction capacity and extraction efficiency with the ligand concentration can therefore be explained by the increase in the absolute and relative concentrations of accessible ligand molecules in the mesopores of the support.
Uranium extraction tests were performed in high sulfate solution as a simulant of ore leaching conditions. Increasing the amount of DEHCEBP in the sample increases the equilibrium time. Fits of the kinetics data with the Weber and Morris model indicate that the increase of the amidophosphonate content leads to an increase of diffusion inside the porous structure and also an increase of the diffusion in the organic layer. These results are in keeping with the decrease of the total pore volume and the multilayer arrangement of the DEHCEBP molecules in the materials with higher ligand concentrations.
Although results suggest that the ligand molecules located inside the micropores do not contribute to uranium extraction, the amount of accessible ligand in the mesopores increases with the ligand concentration (in absolute and relative terms), and thus the extraction capacity of the materials also increases, reaching 108 mg g−1 at 1.2 mmol g−1 DEHCEBP. These materials exhibit higher extraction capacities than the silica-based materials we previously studied.
Provided they could be obtained in column-suited shapes, they could therefore be used for solid-phase uranium extraction in acidic solutions with high sulfate concentrations. Several shaping methods for these materials, their amenability to recycling, elution and reuse are currently being studied. We strongly believe that the conclusions drawn in this study can be extended to the impregnation of other organic compounds and their various applications in the field of adsorption.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3nj04415g |
This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2024 |