A.
Boda
,
A. K.
Singha Deb
,
Sk. M.
Ali
*,
K. T.
Shenoy
and
S.
Mohan
Chemical Engineering Division, Chemical Engineering Group, Bhabha Atomic Research Centre, Mumbai, India. E-mail: musharaf@barc.gov.in
First published on 18th October 2017
Density functional theoretical modelling was performed to design and screen suitable macrocyclic crown ether functionalized resins for the isotopic enrichment of gadolinium. Theoretical calculations predict the complexation stability order of Gd3+ ion as follows: di-cylohexano-18-crown-6 (DCH18C6) > dibenzo-18-crown-6 (DB18C6) > benzo-15-crown-5 (B15C5), which was experimentally verified. The calculated isotopic separation factor value was shown to be the highest for DB18C6. From the theoretical analysis of both the stability and isotopic separation factor, DB18C6 is predicted to be the most promising candidate for isotopic separation of gadolinium. Hence, DB18C6 was functionalized with chloromethylated polystyrene (CMPS) resin. Subsequently, CMPS-grafted DB18C6 resin was synthesized and characterized. Furthermore, isotopic enrichment of gadolinium was carried out by performing column chromatographic experiments using CMPS-DB18C6 resin. The absorption capacity of the novel CMPS-DB18C6 resin for gadolinium was found to be 1 mg g−1. The separation coefficient, ε × 103, was found to be 6.3, 8.9, 3.4, and 9.7 for Gd-155/158, Gd-156/158, Gd-157/158, and Gd-155/160 isotopic pairs, respectively, and thus hold promise for future isotopic enrichment technology.
Design, System, ApplicationQuantum electronic structure theory based computational methods are becoming increasingly central to understanding the fractionation of stable isotopes in various chemical exchange systems. Development of suitable adsorbents for the separation of isotopes by means of experiment alone is a very difficult task which can be made a little easier by predicting the enrichment factor for the elements of interest employing computational theory. The separation factor α for the isotopic exchange reaction is evaluated by the reduced partition function ratio which needs the harmonic vibrational frequency of the system of interest. Therefore, in the present study a strategy has been conceived for the isotope separation of gadolinium which is of three folds. First, a computational modelling strategy has been envisaged to design a crown ether based functionalized polymeric resin for isotope separation of gadolinium followed by the synthesis of the designed resin and characterization and finally, testing of the synthesized resins for gadolinium isotope separation by performing column chromatography experiments. The work presented in this report holds promise for the future isotopic enrichment technology for gadolinium using column chromatography. The enriched gadolium isotope is of immense useful as the neutron poison in the nuclear reactor. |
Naturally occurring gadolinium consists of 7 stable isotopes: 152Gd, 154Gd, 155Gd, 156Gd, 157Gd, 158Gd, and 160Gd; their abundance ratios are 0.20, 2.18, 14.80, 20.47, 15.65, 24.84, and 21.86 at%, respectively.1 Gadolinium exhibits the highest cross section for the capture of thermal neutrons of any element, which is mainly due to the highest ever cross section of Gd-157 (254000 barn) and Gd-155 (60900 barn). All the remaining isotopes have lower values: Gd-152 (735 barn), Gd-154 (85 barn), Gd-156 (18 barn), Gd-158 (2.2 barn), and Gd-160 (1.4 barn). At present, natural gadolinium is used as a burnable poison in nuclear fuel, but the use of Gd-155/157 would create even more effective burnable poison2 due to their high neutron cross section and will thus reduce the core inventory. Apart from the use of Gd-155 and Gd-157 as a neutron poison, Gd-152, which is a precursor for the production of radioactive Gd-153, is widely used for osteoporosis research and bone density measurements.
There are numerous methods for the preparation and production of isotopes of interest at desired isotopic concentrations. Among various available methods, three methods being commonly practiced for the production of stable isotopes are: distillation,3,4 centrifuge-based enrichment,5,6 and electromagnetic enrichment7–11 (using a calutron). Distillation-based enrichment methods only work effectively when there is a large relative mass difference between the different isotopes of an element. Therefore, this method is used for the isotopes of lighter elements, such as He, Li, B, and C, which are typically separated through distillation or diffusion enrichment techniques. The isotopes of elements that are too heavy to be separated through distillation are enriched using centrifuge-based methods. Furthermore, an appropriate gaseous compound of the element is essential but it may not always be available, and thus, there are limitations to this method. Although electromagnetic-based enrichment will sufficiently separate the isotopes of nearly all elements, it is very expensive for large quantity production, as it has an extremely low throughput, but it can allow very high purities to be achieved. This method is often used for the processing of small amounts of pure isotopes for research or specific use, but is impractical for industrial use. The electromagnetic-based calutron is used for the production of isotopes of Pd, Sr, Ca, and the lanthanides. Other methods such as laser enrichment, photochemical enrichment, and plasma separation are also being studied and tested for feasibility. Although significant separation has been observed using the laser isotope method, it is not widely used because enrichment by laser would be highly expensive for commercial production of isotopes.
Another important isotope separation process is the chemical exchange method. Because chemical exchange isotope separation methods typically display small separation coefficients, a very long time is required to produce isotopes of a desired concentration. Nevertheless, separation of isotopes by ion exchange chromatography is considered to be one of the most effective chemical exchange methods, which is based on the chemical equilibrium between isotopic species distributed between the stationary resin phase and the mobile solution phase. Chemical exchange techniques have been practiced quite successfully for the separation of isotopes of various elements using metal ion ligand exchange systems, namely: Li,12 Ce,13 Zn,14,15 Eu,16,17 Cu,18,19 Gd,20–22 V,23 and Nd.24,25 Ion exchange is considered to be one of the processes based on chemical exchange in the equilibrium state and a highly promising technique for the separation of various isotopes.
Macrocyclic polyethers (crown ethers) and aminopolyethers (cryptands) have found widespread application in different fields of separation science and technology since their introduction through the work of Pederson26 and Lehn.27 One of their most remarkable properties is the capability to complex metal cations based on their cavity size. During complexation, the vibrational energy levels of the cation are affected. In the resultant cage, the cation is fixed but able to vibrate. The nature of these bonds and their combined effect in a macrocyclic ether produce a large number of vibrational degrees of freedom that should therefore be suitable for the detection of isotopic effects. This cavity-fitting selectivity can also be exploited for isotope separation by means of chemical reactions with cyclic polyether and bi-cyclic amino polyethers.26–29 In order to accomplish an efficient isotopic fractionation for the enrichment of isotopes, different isotopic distribution must take place between two phases, which then can be easily separated. Most of the macrocyclic crown ethers are soluble in various solvents.29 These compounds can also be functionalized as anchor groups on organic or inorganic matrices.30 Therefore, a heterogeneous exchange reaction for metal ions in liquid–liquid as well as liquid–solid molecular systems can be performed for a required level of enrichment for the isotopes of interest.
The separation of metal ion isotopes using macrocyclic crown ethers was first demonstrated by Jepson and Dewitt in 1976 using dibenzo-18-crown-6 and dicyclohexyl-18-crown-6 compounds.31 After a gap of ten years since the original Pederson discovery of the unusual binding ability of the crown ethers towards metal cations,26 it was shown that crown ethers were also quite effective for the separation of calcium isotopes. An imperative conclusion was the comprehension that all the crown ether and cryptand ligands demonstrate some degree of isotopic recognition, albeit small in most cases. Later, Heumann32 comprehensively reviewed most of the literature on the isotopic separation of various metal ions using crown ethers and cryptand compounds.
Since very high isotopic separation was realized in ion exchange equilibrium with crown ethers and cryptands, interest has recently been shown for these isotopic enrichment systems. Hence, crown ethers are being tested as a highly promising isotopic enrichment agent from a mixture of different isotopes having a very small difference in size.
The application of computational chemistry-based molecular engineering is rapidly growing with the continuing development of computer power, new and robust algorithms, and the availability of fast software. Today, molecular modelling can sometimes provide useful determination of the properties and behaviour of materials even before they have been synthesized, as well as useful estimations of the parameters and behaviour needed for traditional chemical engineering process development and design.33 From an experimental point of view, choosing a suitable solvent/extractant from a myriad of solvents and extractants is a very time-consuming and tedious affair. It will be of tremendous help if the screening of the experiment is done beforehand by means of other easy and less time-consuming techniques for the prediction of this parameter. Development of suitable adsorbents for the separation of isotopes by means of experiment alone is a very difficult task that can be made easier by employing computational theory to predict the enrichment factor for the elements of interest. In order to achieve that goal, molecular modelling studies have previously been performed for solvent extraction and isotope separations.34–38 Therefore, in the present study, a strategy has been conceived for the isotopic separation of gadolinium consisting of three parts. First, a computational modelling strategy has been envisaged to design a crown ether based on functionalized polymeric resin for the isotopic separation of gadolinium followed by the synthesis of the designed resin and characterization, and finally, testing of the synthesized resins for gadolinium isotope separation by performing column chromatography experiments.
The complexation energies are calculated for the complexation of gadolinium metal ions (Gd3+) with ligands (L) as given below:
Gd3+ − (H2O)9 + L → Gd3+ − L + 9H2O | (1) |
The complexation energy, ΔE, is defined by the following general relation:
ΔE = (EGd3+−L + 9EH2O) − (EGd3+−(H2O)9 + EL) | (2) |
All the Hessian calculations were done at a temperature of 298.15 K and pressure of 1 atm. The value of R is taken as 8.314 J mol−1 K−1. The free energy of complexation was then evaluated using the earlier standard reported methodology.44 The solvent phase effect in the energy was accounted by employing a conductor-like screening model (COSMO).45 The dielectric constants of 80 and 34.81 were used for water and nitrobenzene, respectively. The default COSMO radii were used for all the elements.
The isotopic exchange reaction for gadolinium metal ions (Gd3+) between aqueous and complex phase is expressed as:
zGd3+aq. + z+ΔmGd3+compl ↔ z+ΔmGd3+aq. + zGd3+compl | (3) |
Here, z = atomic mass, solv denotes solvation by solvent, and compl denotes complexation by the organic ligand. The separation factor α for the isotopic exchange reaction (eqn (3)) can be written in terms of the reduced partition function ratio (fr) as47
α = frGd3+aq./frGd3+compl | (4) |
fr = Zvibz+ΔmGd3+Πi(hνiz+ΔmGd3+/kBT)/ZvibzGd3+Πi(hνizGd3+/kBT) | (5) |
Here, Z = vibrational partition function, ν = vibrational frequency, and h = Planck constant. kB is the Boltzmann constant, and T is the temperature. The main input required for the evaluation of the reduced partition function ratio and hence for the estimation of the separation factor is the harmonic vibrational frequency of the hydrated Gd3+ ion and Gd3+–crown ether complex.
COSMOtherm48 software was used for the calculation of solubility and partition coefficients of various neutral ligands in the water–organic bi-phasic system.
The grafting of the resin was monitored by Fourier transform infra-red spectroscopy (FTIR, ABB-MB-3000, spectral range between 500 cm−1 and 4000 cm−1 with 4 cm−1 resolution), nuclear magnetic resonance (13C-NMR, Bruker Avance 500, frequency 500 MHz, chemical shifts (δC) are quoted in parts per million (ppm) downfield from trimethylsilane (TMS), and coupling constants (J) are quoted in Hertz (Hz)), spectroscopy, and thermogravimetric analysis (TGA, Netzsch-STA 449-Jupiter F3). The structure and morphology of the grafted resin were analysed by scanning electron microscopy (SEM, Camscan MV2300CT/100) and X-ray diffraction (XRD, Philips 1729).
First, a nitro derivative of DB18C6 was prepared by refluxing DB18C6 using nitric acid/acetic acid in chloroform medium followed by amination using Pd/C (10%) in hydrazine hydrate, following a reported procedure.49 The aminated crown ether was functionalized in CMPS resin using a condensation reaction of methyl chloride resin and amine (crown).50
The distribution coefficient (Kd) of metal ion, between organic and aqueous solution, is determined using the following equation:
(6) |
Fig. 2 (a) Optimized geometry of nona-hydrated Gd3+(H2O)9 and (b) the calculated IR spectra of Gd3+(H2O)9. |
The calculated Gd–O distance (2.49 Å) is well matched with the EXAFS data (2.39 Å).51 The corresponding IR frequency with atomic mass of Gd of 157 amu is also presented in Fig. 1. Five crown ethers, namely B15C5, DB15C5, B18C6, DB18C6, and DCH18C6, were used in the complexation studies of the Gd3+ ion. In the case of B15C5, DB18C6, and DCH18C6, we have taken the most stable conformers from the results of our earlier study, where the conformational search was performed.36,52 In the case of DB15C5, we have added one more phenyl group to the most stable B15C5, and hence, no conformational search was done. For B18C6, we have also taken the geometry from our earlier study and reoptimized it.38 The optimized structures of B15C5 and DB15C5 and their complexes with Gd3+ ion are given in Fig. 3.
Fig. 3 Optimized structures of (a) B15C5, (b) DB15C5, and their complexes with Gd3+ ion optimized at the BP/SVP level of theory. |
As discussed earlier, the vibrational IR frequency is the central quantity for the determination of the separation factor, and hence, the IR frequency of the B15C5 and DB15C5 Gd3+ ion complexes is plotted in Fig. 4.
Next, the optimized structures of B18C6, DB18C6, and DCH18C6 and their complexes with Gd3+ ion are displayed in Fig. 5.
Fig. 5 Optimized structures of (a) B18C6, (b) DB18C6, (c) DCH18C6, and their complexes with Gd3+ ion at the BP/SVP level of theory. |
The reported ionic radii of Gd3+ ion is 2.16 Å. From the structural parameters, it was found that the metal ion oxygen bond distance is smaller in the case of B15C5 (2.368 Å) as compared to B18C6 (2.465 Å), DB15C5 (2.387 Å), DCH18C6 (2.402 Å), and DB18C6 (2.625 Å). Furthermore, the position of the Gd3+ ion with respect to the mean plane is also displayed in Fig. 6.
The calculated IR frequency of Gd3+ ion complexes of B18C6, DB18C6, and DCH18C6 is plotted in Fig. 7.
Description | Charge | s | p | d | f |
---|---|---|---|---|---|
Gd3+–(H2O)9 | 1.889 | 4.38 | 11.99 | 10.71 | 7.016 |
B15C5 | 2.305 | 4.33 | 11.98 | 10.35 | 7.018 |
DB15C5 | 2.314 | 4.33 | 11.98 | 10.34 | 7.018 |
B18C6 | 2.204 | 4.36 | 11.99 | 10.41 | 7.018 |
DB18C6 | 2.227 | 4.35 | 11.99 | 10.40 | 7.019 |
DCH18C6 | 2.121 | 4.35 | 11.99 | 10.51 | 7.022 |
Complex of Gd3+ | E HUMO | E LUMO | E LUMO–HOMO | χ | η | ΔN |
---|---|---|---|---|---|---|
(H2O)9 | −9.80 | −7.90 | 1.89 | 8.85 | 0.94 | |
B15C5 | −5.80 | −0.22 | 5.58 | 3.01 | 2.79 | 0.78 |
DB15C5 | −5.86 | −0.40 | 5.45 | 3.13 | 2.72 | 0.89 |
B18C6 | −5.76 | −0.18 | 5.57 | 2.97 | 2.78 | 0.78 |
DB18C6 | −5.79 | −0.20 | 5.58 | 3.00 | 2.79 | 0.78 |
DCH18C6 | −6.38 | 0.88 | 7.26 | 2.75 | 3.63 | 0.88 |
Fig. 9 Calculated LUMO–HOMO gap of (a) B15C5, (b) B18C6, (c) DB15C5, (d) DB18C6, and (e) DCH18C6 in nitrobenzene. |
The amount of charge transfer is expressed in the following equation: , where χ and η denote absolute electronegativity and hardness, respectively; M and L represent the expression of metal ion and ligand, respectively. The charge transfer was determined for the metal ion–ligand acceptor donor interaction, and the calculated values are given in Table 3. It is worth noting that the calculated large ion-ligand free energy (ΔG) can be qualitatively correlated with the higher amount of charge transfer, ΔN, of the crown ethers as follows: DCH18CH > B18C6 > DB18C6 > B15C5 except DB15C5. DB15C5 displays the highest ΔN but the lowest free energy (ΔG).
To gain further insights into the nature of bonding, the distribution of HOMO and LUMO orbitals in complexes of Gd with different crown ethers in nitrobenzene is displayed in Fig. 10. From the figure, it is seen that the distributions of HOMO and LUMO orbitals of Gd complexes in nitrobenzene are dissimilar, and hence, their LUMO–HOMO gaps are different. The participation of Gd f orbitals is manifested in the distribution of LUMO and HOMO orbitals. For all cases, the same iso-surface value of 0.011 (e bohr−3) has been used.
Fig. 10 The calculated LUMO–HOMO gap of complexes of Gd3+ with (a) B15C5, (b) B18C6, (c) DB15C5, (d) DB18C6, and (e) DCH18C6 in nitrobenzene. |
Complex | RPFR(f) | α 155/160 = faq/fcomp | α 157/160 = faq/fcomp | |
---|---|---|---|---|
Gd155/160 | Gd157/160 | |||
Gd–9w | 1.0024 | 1.0014 | ||
EDTA–Gd3+ | 1.00151 | 1.000914 | 1.00089 (1.000064) | 1.000485 (1.000040) |
DCH18C6–Gd3+ | 1.001890 | 1.001083 | 1.000509 | 1.000317 |
B18C6–Gd3+ | 1.00179 | 1.00104 | 1.000608 | 1.000359 |
DB18C6–Gd3+ | 1.000647 | 1.000391 | 1.001752 | 1.001009 |
B15C5–Gd3+ | 1.002016 | 1.001213 | 1.000383 | 1.000187 |
DB15C5–Gd3+ | 1.00176 | 1.001051 | 1.000638 | 1.000348 |
CMPS–DB18C6 | 1.001191 | 1.000712 | 1.0012 | 1.00068 |
The results show that the separation factor for DB18C6 is higher when compared to B15C5, DB15C5, B18C6, and DCH18C6. With its comparatively higher separation factor, DB18C6 is considered to be the most promising ligand for isotopic enrichment of gadolinium. However, in order to obtain a considerable amount of enriched isotopes, the crown ether ligand should be grafted on a solid matrix, and then the isotopic separation can be carried out in column chromatography mode. The predictability of the computational protocol was validated using the reported results for commercially available DOWEX resin with EDTA as the complexing/enrichment agent. Furthermore, the calculated separation factor for the 152/160 pair for Gd3+–DCH18C6 (α = 1.0012) was found to be in good agreement with the reported experimental results (α = 1.002)22 and thus confirms the acceptability of the present methods for computation of the RPFR and separation factor.
System | Solubility (gm cc−1) | logP |
---|---|---|
B15C5 | 5.8 × 10−5 | 2.27 |
DB15C5 | 4.2 × 10−7 | 4.87 |
B18C6 | 4.0 × 10−5 | 1.87 |
DB18C6 | 1.4 × 10−3 | 0.21 |
DCH18C6 | 4.2 × 10−7 | 2.60 |
Fig. 11 Optimized structures of DB18C6 grafted to CMPS with Gd3+ ion optimized at the BP/SVP level of theory and position of the ion with respect to the mean plane. |
In order to study the effect of CMPS resin functionalization on the structure of DB18C6, the cavity size of DB18C6 was evaluated. The cavity of the crown ether has been considerably reduced (2.421 Å) from the free DB18C6 (2.820 Å) after functionalization with CMPS resin, which perhaps resulted in a higher Gibbs free energy of complexation. Furthermore, from the figure, it is seen that the Gd3+ ion is located at the centre of the cavity of the CMPS-DB18C6 moiety in a highly symmetric fashion. The average Gd–O bond distance in the case of CMPS-DB18C6 was found to be 2.521 Å, which is shorter than that observed with DB18C6 (2.625 Å), and thus, it experiences a larger interaction. Another interesting point to be noted is that in the case of complex of Gd3+ with DB18C6, the two benzene rings were highly non-planar, whereas with functionalized CMPS-DB18C6, the non-planarity of benzene rings has been drastically reduced.
The effect of functionalization on energy was examined by determining the adsorption energy (ΔE) and free energy (ΔG) of Gd3+ ion with CMPS-DB18C6 resin in the aqueous phase. The calculated values of zero point energy-corrected adsorption energy (ΔE) and free energy (ΔG) of Gd3+ ion with CMPS-DB18C6 resin in the aqueous phase are presented in Table 6. The free energy of adsorption with functionalized CMPS-DB18C6 resin was found to be higher than that of DB18C6. This might be due to the reduced LUMO–HOMO energy gap (4.74 eV) for the CMPS-DB18C6 resin compared to DB18C6 (5.58 eV). The distribution of LUMO–HOMO orbitals for the CMPS-DB18C6 resin and its complex with Gd3+ is displayed in Fig. 12. Furthermore, the smaller residual charge on the Gd atom in the complex of CMPS-DB18C6 compared to that of DB18C6 also supports the higher free energy. The higher amount of charge transfer, ΔN, also confirms the higher Gibbs adsorption free energy for the CMPS-DB18C6 resin over DB18C6.
Description | ΔE | ΔG | ΔS* | Charge | ΔN |
---|---|---|---|---|---|
CMPS-DB18C6 | 25.52 | −43.59 | 0.247 | 2.191 | 0.896 |
Fig. 12 Calculated LUMO–HOMO gap of (a) CMPS-DB18C6 and (b) its complex of Gd3+ ion in an aqueous environment. |
Next, the reduced partition function ratio (RPFR) and isotope separation factor (α) for hydrated Gd3+ ion and complexes of Gd3+ with CMPS-DB18C6 resin were determined using the vibrational IR spectrum. The calculated vibrational IR spectrum for the Gd3+–CMPS-DB18C6 resin is plotted in Fig. 13, whereas the calculated value of the RPFR and α are presented in Table 4. The value of the RPFR increased compared to that of the Gd3+–DB18C6 complex, which resulted in a slight decrease in the separation factor compared to DB18C6. However, the calculated separation factor for the CMPS-DB18C6 resin was higher than that of the other crown ethers.
Group | IR peaks (cm−1) |
---|---|
–CH2Cl | Absent (758) |
–NH | 3078, 3093 |
–CH2– | 2861, 1473 |
C–O–C | 1134 |
The 13C NMR spectrum of CMPS and CMPS-DB18C6 resin are shown in Fig. S2 (ESI†), and the interpretation of the individual chemical shift values are depicted in Table 8. The grafting of the DB18C6 on the CMPS resin can be seen from the C–N bond, C(aromatic)–O, and C(aliphatic)–O chemical shift values in the CMPS-DB18C6 resin.
Compound | Chemical shift (ppm) | Assignments |
---|---|---|
CMPS | 48 | CH2Cl |
128 | Protonated aromatic carbon | |
149 | Quarternary aromatic carbon | |
DB18C6 functionalized CMPS | 32 | CH2 |
38 | CH | |
42 | C–NH | |
48 | CH2Cl | |
66 | C(aliphatic)–O | |
128 | Protonated aromatic carbon | |
149 | Quarternary aromatic carbon | |
165 | C(aromatic)–O |
The TGA curves of CMPS and CMPS-DB18C6 resin are shown in Fig. 14. The lower decomposition temperature of CMPS-DB18C6 resin compared to CMPS resin indicates the surface anchoring of DB18C8.
The SEM images of CMPS and CMPS-DB18C6 are shown in Fig. 15. The morphological change in the surface of the resin beads is clear from the images. The broad peak centered at the 2θ value of approximately 20°, observed in the XRD pattern of CMPS-DB18C6 resin (Fig. 15), is characteristic of the polystyrene–divinylbenzene cross-linked resin.56
The IR and NMR spectra are presented in the ESI.†
The distribution coefficient (Kd) of metal ion, between the resin and solution phase, was determined using the following equation:
(7) |
The Kd values are plotted in Fig. 17, which indicates that there is an increase in the Kd value with increasing concentration of metal ion in the solution.
Furthermore, an isotope separation study was conducted by connecting two columns that were 6 mm in diameter and 1 m in length with CMPS-grafted DB18C6 resin for Gd3+ ion using 0.005 M Gd(NO3)3 solution. The pH of the solution was adjusted with NH4OH to 6.50. Excess Gd3+ was continuously passed to test the feasibility of the isotope exchange. Samples were collected after each 20.0 ml using an automatic fraction collector (Buchi, Switzerland). The concentration of the samples was analyzed using ICP-MS, and isotopic analysis of last few samples was carried out using TIMS. The loaded Gd was further eluted using 0.5 M HCl. The breakthrough curve is displayed in Fig. 19. The estimated values of isotope composition are presented in Tables 9 and 10.
Fig. 19 Breakthrough curve using CMPS-grafted DB18C6 resin in two 1 meter columns connected in a series for Gd3+ ion using 0.005 M Gd(NO3)3. |
Sample | Gd-155/158 | Gd-156/158 | Gd-157/158 | Gd-158/160 |
---|---|---|---|---|
Natural | 0.59742 ± 0.04% | 0.82546 ± 0.04% | 0.63212 ± 0.05% | 1.13564 ± 0.05% |
B1 | 0.59940 ± 0.03% | 0.82740 ± 0.02% | 0.63410 ± 0.02% | 1.13090 ± 0.01% |
B2 | 0.59670 ± 0.01% | 0.82480 ± 0.01% | 0.63220 ± 0.02% | 1.13030 ± 0.01% |
B3 | 0.59831 ± 0.03% | 0.82585 ± 0.01% | 0.63236 ± 0.01% | 1.13143 ± 0.01% |
B4 | 0.59766 ± 0.03% | 0.82594 ± 0.02% | 0.61782 ± 0.02% | 1.12946 ± 0.03% |
Sample | Gd-155/158 | Gd-156/158 | Gd-157/158 | Gd-158/160 |
---|---|---|---|---|
E1 | 0.59721 ± 0.09% | 0.82499 ± 0.06% | 0.63169 ± 0.1% | 1.131028 ± 0.05% |
E2 | 0.60181 ± 0.002% | 0.82941 ± 0.2% | 0.61518 ± 0.1% | 1.13949 ± 0.5% |
E3 | 0.59510 ± 0.015% | 0.82283 ± 0.025% | 0.63114 ± 0.026% | 1.12962 ± 0.023% |
E4 | 0.60102 ± 0.02% | 0.82710 ± 0.1% | 0.63639 ± 0.1% | 1.12939 ± 0.04% |
The single stage separation factor value, α = 1 + ε, for each Gd isotope can be estimated by the following expression:
(8) |
The separation coefficient, ε, was calculated using the isotopic enrichment curves of the front and rear boundaries according to the equation developed by Spedding et al.57 and Kakihana and Kanzaki:58
(9) |
The measured Gd isotopic ratios of samples collected in this breakthrough operation are presented in Table 9. The symbols B1, B2, B3, and B4 represent the different volume fraction during the breakthrough operation. From Table 9, it is observed that the heavier isotope, i.e., Gd-160, is enriched in volume fractions, which indicates that the frontal zone is enriched with heavier isotope compared to natural gadolinium. From the isotopic values in Table 9, the isotope separation coefficient was calculated. The separation coefficient, ε × 103, was found to be 6.3, 8.9, 3.4, and 9.7 for Gd-155/158, Gd-156/158, Gd-157/158, and Gd-155/160, respectively.
Next, the measured Gd isotopic ratios of samples collected by eluting with HCl are presented in Table 10, where E1, E2, E3, and E4 represent the selected eluted volumes. From Table 10, it is observed that lighter isotope, i.e., Gd-155, is enriched in the collected fractions, which indicates that the rare zone is enriched with lighter isotope. It should be noted that from E1 to E4, the concentration of Gd-155 increased compared to natural gadolinium.
From the isotopic ratio measurements, it was observed that slight isotopic separation may be taking place in the bed. Because the absorption capacity is low, more material is needed for the achievement of higher separation.
Footnote |
† Electronic supplementary information (ESI) available: Table S1: calculated values of energies, enthalpy, entropy, and Gibbs free energy of the chemical species; Table S2: optimized coordinates of the complexes; Fig. S1. IR spectrum of (a) DADB18C6, (b) CMPS resin, and (c) DB18C6 functionalized CMPS resin; Fig. S2. 13C NMR of (a) CMPS and (b) DB18C6 modified CMPS resin. See DOI: 10.1039/c7me00076f |
This journal is © The Royal Society of Chemistry 2017 |