Debajit
Sarma
a,
Christos D.
Malliakas
a,
K. S.
Subrahmanyam
a,
Saiful M.
Islam
a and
Mercouri G.
Kanatzidis
*ab
aDepartment of Chemistry, Northwestern University, 2145 Sheridan Road, Evanston, IL 60208, USA. E-mail: m-kanatzidis@northwestern.edu
bMaterials Science Division, Argonne National Laboratory, Argonne, IL 60439, USA
First published on 27th October 2015
The fission of uranium produces radionuclides, 137Cs and 90Sr, which are major constituents of spent nuclear fuel. The half-life of 137Cs and 90Sr is nearly 30 years and thus that makes them harmful to human life and the environment. The selective removal of these radionuclides in the presence of high salt concentrations from industrial nuclear waste is necessary for safe storage. Here we report the synthesis and crystal structure of K2xSn4−xS8−x (x = 0.65–1, KTS-3) a material which exhibits excellent Cs+, Sr2+ and UO22+ ion exchange properties in varying conditions. The compound adopts a layered structure which consists of exchangeable potassium ions sandwiched between infinite layers of octahedral and tetrahedral tin centers. K2xSn4−xS8−x (x = 0.65–1, KTS-3) crystallizes in the monoclinic space group P21/c with cell parameters a = 13.092(3) Å, b = 16.882(2) Å, c = 7.375(1) Å and β = 98.10(1)°. Refinement of the single crystal diffraction data revealed the presence of Sn vacancies in the tetrahedra that are long range ordered. The interlayer potassium ions of KTS-3 can be exchanged for Cs+, Sr2+ and UO22+. KTS-3 exhibits rapid and efficient ion exchange behavior in a broad pH range. The distribution coefficients (Kd) for KTS-3 are high for Cs+ (5.5 × 104), Sr2+ (3.9 × 105) and UO22+ (2.7 × 104) at neutral pH (7.4, 6.9, 5.7 ppm Cs+, Sr2+ and UO22+, respectively; V/m ∼ 1000 mL g−1). KTS-3 exhibits impressive Cs+, Sr2+ and UO22+ ion exchange properties in high salt concentration and over a broad pH range, which coupled with the low cost, environmentally friendly nature and facile synthesis underscores its potential in treating nuclear waste.
The most commonly used technique for the separation of radioactive elements from industrially produced nuclear waste is solvent extraction using liquid phase organic compounds.8–10 The use of ion exchange media is another alternative for the removal of radionuclides from the nuclear waste,11–22 however, they are relatively less explored due to certain drawbacks: the organic ion exchange materials are efficient but costly, whereas the inorganic ion exchange materials are cheaper but they are less efficient because of low selectivity for the ions of interest. So, there is a growing need to develop efficient inorganic ion exchange materials for radioactive species.
Over the past decade or so metal sulfides have emerged as a selective class of ion exchangers for capturing soft metal ions such as Hg, Cd, Ag etc.20,23,24 Chalcogenide open-framework compounds, such as K6Sn[Zn4Sn4S17]17 and (NH4)4In12Se20 (ref. 19) present unique advantages over their oxide analogues. The layered thiostannates are particularly interesting because they exhibit open accessible structures where ion-exchange chemistry can occur readily.25–30 In previous work, we proposed that layered metal sulfides K2xMxSn3−xS6 (M = Mn, KMS-1; M = Mg, KMS-2) can be used for facile ion exchange of Sr2+, Cs+ and UO22+.18,19,31,32 A variety of synthetic parameters were explored in the search for new compounds based on a tin sulfide layer structure to modulate the ion exchange properties. The advantage of the chalcogenide materials stems from the fact that they are based on softer chalcogen ligands (in the Lewis base sense) which can induce high selectivity for heavy metal ions, Cs+, Sr2+, UO22+ against co-present hard ions such as Na+, Al3+ and Ca2+.17,19,31–33
Herein, we report a new ternary layered compound, K2xSn4−xS8−x (x = 0.65–1, KTS-3) and its promising selectivity for removing Cs+, Sr2+ and UO22+ species via ion exchange processes. Specifically, we find that KTS-3 exhibits high distribution coefficients (Kd) for the capture of Cs+ (5.5 × 104), Sr2+ (3.9 × 105) and UO22+ (2.7 × 104) over a broad pH range (V/m ∼ 1000 mL g−1). We find that KTS-3 remains highly effective for these ions even in presence of a large amount of Na+ ions.
The distribution coefficient Kd, used for the determination of the affinity and selectivity of KTS-3 for Cs+, Sr2+, or UO22+ is given by the equation: Kd = (V/m)[(C0 − Cf)/Cf] where, V is the volume (mL) of the testing solution, m is the amount of the ion exchanger (g), C0 and Cf are the initial and equilibrium concentration of a given ion An+ (ppm).
The individual Cs+, Sr2+, or UO22+ uptake from solutions of various concentrations were studied with V/m ≈ 1000 mL g−1, at room temperature and 15 h contact time. The data obtained were used for the determination of the sorption isotherms. All ion exchange experiments reported in this work were performed by the batch method in 20 mL scintillation vials.
The individual ion exchange experiments for the Cs+, Sr2+, or UO22+ ions at different pH and salt concentration were also carried out. The required pH values (2, 4, 6, 8, 10 and 12) were achieved by diluting the commercial standards (1000 ppm) with HCl or NaOH solution to ∼6 ppm. The ion exchange experiment at different Na+ concentration was done by dissolving the required amount of NaCl in 10 mL solution of An+ ion (∼6 ppm). The exchange experiments were performed with V/m ≈ 1000 mL g−1, at room temperature and 15 h contact.
Competitive ion exchange (Cs+ and Sr2+) experiments of KTS-3 were also carried out with a V/m ratio of 1000 mL g−1, at room temperature with 15 h of contact time. The initial concentration was approximately ∼6 ppm for both the ions. The competitive ion exchange experiments were similar to those of the individual ion exchange experiments except they contained both Cs+ and Sr2+ ions in solution.
The kinetic studies of the adsorption of ions by KTS-3 were carried out as follows: ion-exchange experiments of various reaction times (5, 15, 30, 60, 120, 300 and 1200 min) were performed. For each experiment, 10 mg of KTS-3 was weighed into a 20 mL vial. A 10 mL sample of water solution containing ∼1 ppm of Cs+/Sr2+/UO22+ was added to each vial, and the mixtures were kept under magnetic stirring (pH ∼ 7). The suspensions were filtered after the designated reaction time and the filtrates were analyzed by inductively coupled plasma-mass spectroscopy (ICP-MS).
K2Sn3S7 | Subcell | Supercell |
---|---|---|
a R = ∑‖Fo| − |Fc‖/∑|Fo|, wR = {∑[w(|Fo|2 − |Fc|2)2]/∑[w(|Fo|4)]}1/2 and w = 1/[σ2(Fo2) + (aP)2 + bP] where P = (Fo2 + 2Fc2)/3. The twin law [1 0 0.5 0 −1 0 0 0 −1] was used with a refined fraction of 45.0(3)%. | ||
Formula weight | 658.69 | 661.31 |
Wavelength | 0.71073 Å | |
Crystal system | Orthorhombic | Monoclinic |
Space group | Cmcm | P21/c |
Unit cell dimensions | a = 3.6831(2) Å, α = 90° | a = 13.092(3) Å, α = 90° |
b = 25.8877(19) Å, β = 90° | b = 16.882(2) Å, β = 98.100(15)° | |
c = 16.8155(11) Å, γ = 90° | c = 7.3748(13) Å, γ = 90° | |
Volume | 1603.31(18) Å3 | 1613.7(5) Å3 |
Z | 4 | |
Density (calculated) | 2.729 g cm−3 | 2.722 g cm−3 |
Absorption coefficient | 6.026 mm−1 | 6.029 mm−1 |
F(000) | 1200 | 1204 |
Color | Yellow | |
Crystal size | 0.320 × 0.160 × 0.050 mm3 | |
Index ranges | −4 ≤ h ≤ 4, −34 ≤ k ≤ 34, −22 ≤ l ≤ 22 | −19 ≤ h ≤ 19, −25 ≤ k ≤ 25, −10 ≤ l ≤ 11 |
Reflections collected | 13142 | 37136 |
Independent reflections | 1145 [Rint = 0.0590] | 5540 [Rint = 0.094] |
Completeness to θ = 25.242° | 99.7% | 100.0% |
Refinement method | Full-matrix least-squares on F2 | |
Data/restraints/parameters | 1145/0/45 | 5540/1/121 |
Goodness-of-fit | 1.140 | 1.135 |
Final R indices [I > 2σ(I)] | R obs = 0.0749, wRobs = 0.2183 | R obs = 0.1063, wRobs = 0.2807 |
R indices [all data] | R all = 0.0796, wRall = 0.2237 | R all = 0.1410, wRall = 0.3100 |
Extinction coefficient | 0.0019(5) | |
Largest diff. peak and hole | 3.868 and −1.905 e Å−3 | 4.342 and −2.783 e Å−3 |
Weighting scheme | a = 0.1110, b = 74.4974 | a = 0.141, b = 21.5465 |
Thermogravimetric (TG) analysis of the KTS-3 compound was carried out in flowing nitrogen gas (flow rate = 20 mL min−1) in the temperature range 20–600 °C (heating rate of 10 °C min−1). The TG studies indicate that KTS-3 exhibits a single-step weight loss of ∼10% up to 235 °C which corresponds to the loss of adsorbed water molecules. The compound remains stable up to 525 °C, after which it starts to decompose (Fig. S1b†) into K2Sn2S5 and SnS2 as determined by powder XRD. Differential thermal analysis (DTA) of the samples shows no sign of melting up to 600 °C (Fig. S2†).
X-ray photoelectron spectroscopy performed on KTS-3 (Fig. 2) shows peaks at 292.5 and 295.6 eV which are characteristic for 2p3/2 and 2p1/2 of K+ cations.39 The peaks at 486.0 and 494.5 eV are consistent with the 3d5/2 and 3d3/2 levels observed for Sn4+ cations.39 The sulfur 2p orbital excitations appear as a broad peak in the range 158–165 eV. The deconvolution of the broad band gives two bands centered at 161.5 and 162.7 eV which are characteristic of 2p3/2 and 2p1/2 sulfide anions, respectively.39,40
Fig. 2 X-ray photoelectron spectra of (a) potassium, (b) tin, (c) sulfur, and (d) survey spectrum for KTS-3. Dotted and solid lines represent experimental and deconvoluted spectra, respectively. |
The orthorhombic cell can be refined in Cmcm with a stoichiometry of ‘K2Sn4S8’ but this is problematic because this composition does not charge balance assuming K+, Sn4+, and S2− ions. Furthermore, the agreement factor for the ‘K2Sn4S8’ refinement was very high at ∼14.5% with large negative residual electron density around the Sn(2) and S(4) sites. Upon refinement of the occupancy of Sn(2) and S(4) (50% disorderly occupied) but omitting the supercell reflections, the agreement factor improved significantly (7.5%, see Table 1) and the refined composition becomes K1.92Sn3.04S7.04 which is charge balanced. By subsequently introducing the intensity of the supercell reflections into the refinement, an additional long range ordering of the vacancies in the Sn(2) and S(4) sites was found. The supercell of KTS-3 was solved using the monoclinic spacegroup P21/c and twining was required for a successful refinement. A refined twin fraction of 45.0(3)% was determined using a twin law of 180 degrees rotation along the c-axis, Table 1. The final agreement factor is satisfactory given the very broad and diffuse nature of the supercell reflections, Fig. 3b.
The asymmetric unit of the KTS-3 supercell has 15 atoms. Four crystallographically independent Sn4+ atoms (two sites are partially occupied), eight sulfide atoms (two sites are partially occupied) and three K+ ions (two sites are partially disordered). The Sn(1) and Sn(2) ions are octahedrally coordinated by six sulfur atoms, and Sn(3) and Sn(4) atoms are tetrahedrally coordinated by four sulfur atoms. The [SnS6] and [SnS4] units are shared through S(5)/S(6) edges, the [SnS6] units are edge-shared through S(1)–S(2), and the [SnS4] units are edge-shared through S(7)/S(7) and S(8)/S(8) edges. The Sn(1), Sn(2) distorted octahedra have Sn–S distances in the range of 2.504(2)–2.621(2) Å and the Sn(3) distorted tetrahedral have Sn–S distances in the range of 2.288(2)–2.486(2) Å. Because of the partial ordering of vacancies, Sn(3) and Sn(4) are disordered with a refined fractional occupancy of 73.0(4) and 31.1(5)%, respectively. The same occupancy values were used for the S atoms that edge-share the [SnS4] tetrahedra, i.e., the occupancy factor of S(7) and S(8) was constrained at 73.0(4) and 31.1(5)%, respectively. All K atoms have relatively large thermal factors which is characteristic for loosely bound intercalated atoms found in ion-exchanged materials.17,18,32 K(1A) and K(1B) are delocalized with an average disordered distance of 2.32(1) Å and fractional occupancy of 60.7(5) and 39.3(5)%, respectively where K(3) fully occupies its own site.
The basic difference between the structure of KTS-3 and that of so-called KMS structures which are also layered (K2xMxSn3−xS6; M = Mn, KMS-1; M = Mg, KMS-2)18,32 is in the structure of the layers themselves. The layers of KMS-1 and KMS-2 are essentially derived from the SnS2 structure by replacing randomly some of the octahedral Sn4+ ions by either Mn2+ (KMS-1) or Mg2+ (KMS-2) ions, where all the Sn/M (M = Mn or Mg) ions occupy octahedral sites and the sulfur ions are three coordinated.18,32 However, in case of KTS-3 there are both octahedral and tetrahedral centers that are connected by three and two coordinated sulfur atoms to form the layer structure. Disordered potassium ions are located between the SnS2 or Sn3S7 layers, Fig. 3c.
K1.92Sn3.04S7.04 + 1.92CsCl → Cs1.92Sn3.04S7.04 + 1.92KCl | (1) |
K1.92Sn3.04S7.04 + 0.96SrCl2·6H2O → [Sr(H2O)y]0.96Sn3.04S7.04 + 1.92KCl | (2) |
K1.92Sn3.04S7.04 + 0.96UO2(NO3)2·6H2O → [UO2(H2O)y]0.96Sn3.04S7.04 + 1.92KNO3 | (3) |
Fig. 4 Powder X-ray diffraction patterns of pristine KTS-3 and the exchanged materials. The (020) and (040) reflection peaks for the exchanged materials shift towards lower 2θ (higher d spacing). |
For the Cs+ and Sr2+ exchanged samples the PXRD analysis showed a shift of the (020) and (040) basal Bragg peaks to lower 2θ values (higher d-spacing). The interlayer spacing of the material increases from 8.441 Å to 8.632 Å (Sr2+) and 8.813 Å (Cs+). The PXRD analysis of the UO22+ exchanged sample shows the presence of a mixture of layered phases, which are mainly due to the different degrees of hydration of the UO22+ ions. The interlayer spacing of the UO22+ exchanged material increases from 8.441 Å to 9.966 Å and 10.250 Å. The change in the interlayer spacing follows the order UO22+ > Cs+ > Sr2+ > K+, which is consistent with the ionic size of the ions. The TG analysis (Fig. S4†) showed that the degree of hydration for the exchanged materials follows the order Sr2+ > UO22+ > Cs+ > K+.
The band gap of the pristine KTS-3 material is 2.38 eV and the yellow color of the material changes marginally upon exchange with Cs+ and Sr2+ ions. The exchanged materials show a small increase in absorption and the measured band gaps were 2.54 eV (Cs+) and 2.56 eV (Sr2+). With UO22+ exchange, the yellow color slowly changes to a darker orange color and the band gaps red shifted to 2.30 eV and 2.40 eV (Fig. 5a). This can be attributed to partial dehydration of the UO22+ ions and the presence of U⋯S interactions. The presence of two band gaps for the UO22+ exchanged material was attributed to the differently hydrated UO22+ ions.
The infra-red spectrum of the uranyl exchanged KTS-3 material shows a strong peak at ∼910 cm−1, which is not found in pristine KTS-3 (Fig. 5b). This peak at ∼910 cm−1 is assigned to the antisymmetric vibration of [OUO]2+ group and is significantly red shifted compared to the peaks found for aqueous [OUO]2+ ions (∼963 cm−1).41
The XPS spectra of the Cs+ exchanged samples show the characteristic 3d5/2 and 3d3/2 for Cs+ at 724.7 and 738.7 eV (Fig. 6a).39 The Sr2+ exchanged samples show peaks at 133.9 and 135.7 eV characteristic for 3d5/2 and 3d3/2 of Sr2+ cations (Fig. 6b).39 The UO22+ exchanged samples show two peaks at 379.6 and 390.6 eV characteristic for 3f7/2 and 3f5/2 of U6+ centers (Fig. 6c).31,39 All exchanged samples showed the characteristic peaks for tin and sulfur ions as observed for the pristine compound. The peaks for the potassium 2p3/2 and 2p1/2 could not be found in the exchanged samples (Fig. 6d), which confirms their complete exchange from the KTS-3 compound.
Cs+ ion exchange | Sr2+ ion exchange | UO22+ ion exchange | ||||
---|---|---|---|---|---|---|
Langmuir | Langmuir–Freundlich | Langmuir | Langmuir–Freundlich | Langmuir | Langmuir–Freundlich | |
q e (mg g−1) | 280(11) | 304(22) | 102 (5) | 113(14) | 287(15) | 358(61) |
b (L mg−1) | 0.09(2) | 0.07(2) | 0.20(8) | 0.19(15) | 0.23(6) | 0.09(6) |
n | — | 1.37(23) | — | 1.81(54) | — | 1.52(24) |
R 2 | 0.965 | 0.972 | 0.923 | 0.928 | 0.952 | 0.960 |
Langmuir isotherm
(4) |
Langmuir–Freundlich isotherm
(5) |
The Langmuir isotherm describes adsorption on a homogenous surface and the maximum adsorption corresponds to a saturated monolayer. This model is based on the assumptions that (a) the adsorption sites are equivalent and each site can only accommodate one molecule, (b) the energy of adsorption is constant and independent of surface coverage, and (c) there is no transmigration of adsorbate from one site to another.42–44 The Langmuir–Freundlich isotherm is an extension of the Langmuir model, which reduces to Freundlich isotherms at low surface coverage and to Langmuir isotherms at high surface coverage.42
The equilibrium data for Cs+ ion exchange (Fig. 7a) could be fitted with both Langmuir, and Langmuir–Freundlich isotherm models with a good agreement (R2 ≥ 0.97). The value of the Langmuir–Freundlich constant n = 1.37(23) was found to be closer to 1 which suggests that the adsorption behavior of Cs+ ion exchange follows the Langmuir adsorption model. The agreement of the Langmuir adsorption isotherm with the Cs+ ion exchange can be rationalized by taking into consideration the structural features of KTS-3. The [Sn3S7]2− layers of KTS-3 are separated by layers of disordered potassium ions, so the exchangeable Cs+ ions form a layer between the [Sn3S7]2− layers that corresponds to the monolayer of Langmuir isotherms. The adsorption sites for the exchangeable ions are fixed (S2− ions) and chemically equivalent. Moreover, once the ions are exchanged it is not possible to migrate to other sites. The equilibrium data for Sr2+ (Fig. 7b) and UO22+ (Fig. 7c) were also fitted with Langmuir (R2 = 0.92 and 0.95, for Sr2+, UO22+ respectively) and Langmuir–Freundlich adsorption (R2 = 0.92 and 0.96, for Sr2+, UO22+ respectively) isotherms in good agreement. The value of Langmuir–Freundlich constant [n = 1.81(54) for Sr2+ and 1.52(24) for UO22+] shows that it deviates from the Langmuir isotherm model (n = 1). The behavior of Cs+, Sr2+ and UO22+ vis-a-vis their isotherms can be rationalized by the fact that the number of ions exchanged in the case of Cs+ is twice that of bivalent Sr2+ and UO22+ and hence it has higher surface coverage and tends to follow better the Langmuir model. The ion exchange of bivalent metal ions often follow the Langmuir–Freundlich model rather than the Langmuir model.45
The maximum ion exchange capacities, qm were found to be 280(11) mg g−1 (2.10 mmol g−1) for Cs+, 102(5) mg g−1 (1.16 mmol g−1) for Sr2+ and 287(15) mg g−1 (1.20 mmol g−1) for UO22+ from the Langmuir isotherm model. The theoretical capacities for K2xSn4−xS8−x (x = 0.96) considering all the K+ ion are exchanged are 2.90 mmol g−1 (385 mg g−1) for Cs+ and 1.45 mmol g−1 for Sr2+ (127 mg g−1), UO22+ (347 mg g−1). The observed Cs+ exchange is about 72%, Sr2+ exchange ∼ 80% and UO22+ exchange ∼ 83% of the theoretical capacity. All K+ ions are exchanged after the reaction and the observed exchange capacity is due to the fact that the polycrystalline sample (K2xSn4−xS8−x, KTS-3) has a range of x values from 0.65–1. The observed ion exchange capacity of KTS-3 compares well with well-known Cs+ and Sr2+ sorbents (e.g., zeolites, sodium silicotitanates and zirconium titanium silicates; 1.86–4.1 mmol g−1 for Cs+ and 1.0–2.0 mmol g−1 of Sr2+).46–49
The Langmuir constants b (L mg−1) for the Cs+, Sr2+ and UO22+ were found to be 0.09(2), 0.20(8) and 0.23(6) L mg−1, respectively. The value of b is an indicator for the affinity towards a particular ion. Higher b values for Sr2+ and UO22+ ions indicate that KTS-3 has larger affinity towards them compared to Cs+. The affinity of a sorbent towards a particular ion can also be expressed in terms of distribution coefficient (Kd),
(6) |
The Kd values were found to be 5.5 × 104 mL g−1 for Cs+, 3.9 × 105 mL g−1 for Sr2+ and 2.7 × 104 mL g−1 for UO22+ (∼6–8 ppm, V/m = 1000 mL g−1 and pH = 7). Kd values in the 104 or 105 ranges are considered to be very good for ion exchange processes.12,13,50–52
The stability of the KTS-3 phase over a range of pH values (2–12) was tested and was found to be impressive. The compound remains crystalline (3 ≤ pH ≤ 11) and retains the layered structure for days when suspended in solution. Even in highly acidic (pH ≤ 2) or basic conditions (pH ≥ 12) it remains stable for hours; a small decomposition of the compound can be seen if kept for more than 24 h (Fig. S5†).
Fig. 8 represents the variation of Kd values for individual and competitive Cs+ and Sr2+ ion exchange with pH. KTS-3 shows excellent Cs+ ion exchange capacity over a pH range of 2–12. It absorbs over 97% of the ions from pH 4 to 10 and it absorbs around 53% of the ions even in a highly acidic environment (pH 2). The KCsd values were found to be ∼3.4 × 104 to 5.5 × 104 in the pH range of 4–10. However, there is slight decrease in the KCsd values at pH = 2 (1.1 × 103) and it falls sharply at pH = 12 (253) (7.4 ppm, V/m = 1000). The decrease in KCsd values may be due to partial decomposition of KTS-3 in these regions of pH. The presence of Sr2+ in solution does not affect the ion exchange of Cs+, as the KCsd values found are comparable with those of the individual values (9.8 × 102 to 6.7 × 104). The small increase in KCsd in the presence of Sr2+ can be attributed to the overall increase in ionic charge of the solution.
KTS-3 exhibits remarkable Sr2+ capture capacity with more than 98% of the ions absorbed between pH 4 to 10. It decreases slightly at pH = 2 (81%) and 12 (88%) but is still much higher than Cs+. The KSrd values for the Sr2+ ion exchange over the pH range 2–12 were found to be 4.2 × 103 to 3.9 × 105 mL g−1 (6.9 ppm, V/m = 1000) (Fig. 8). The presence of Cs+ does not induce an appreciable change as the KSrd value remains almost same 4.5 × 103 to 3.9 × 105 mL g−1.
The Kd value for Cs+ in the presence of a huge excess of Na+ ions decreases slightly from 5.5 × 104 mL g−1 at 0 M concentration to 4.4 × 103 mL g−1 at 0.1 M concentration (Fig. 9a). Further increase in the Na+ concentration reduced the Kd values and even at a Na+ concentration of 1 M, KTS-3 exhibited a reasonable Kd value of 644 mL g−1. The Kd values in the presence of both Sr2+ and Na+ ions vary from 5.5 × 104 mL g−1 at 0 M Na+ to 501 mL g−1 at 1 M Na+ concentration. The Kd value for Sr2+ in both individual and competitive (Cs+) ion-exchange reactions drops sharply in the presence of Na+. Namely, it decreases from 3.9 × 105 mL g−1 at 0 M to 201 at 1 M Na+ concentration (Fig. 9a).
The kinetics of Cs+ adsorption for low concentration (∼1.2 ppm) solutions showed that 94% of the ions were absorbed within 5 min. The competitive Cs+ adsorption (in the presence of Sr2+) showed ∼90% adsorption within 5 min, which remains unchanged thereafter (Fig. 9b). The Sr2+ adsorption 92% (individual) and 92% (competitive in presence of Cs+) occurred within 5 min and with a longer time it increases to 97% (individual and competitive). The small decrease in ion exchange with time for Cs and increase for Sr2+ are due to the dynamic ion exchange process between K+ and Cs+/Sr2+ and the higher affinity of KTS-3 towards Sr2+. Upon contact with KTS-3, the Cs+ replaces the K+ ions immediately and only a small amount of K+ ions gets reabsorbed in the interlayer spaces to release some of initially absorbed Cs+ ions back to solution. However, in the case of Sr2+ the higher affinity of KTS-3 shuts down this dynamic ion exchange.
The ion exchange capacity of K2xSn4−xS8−x (x = 0.65–1, KTS-3) is excellent and compares well with K2xMxSn3−xS6 (M = Mn, KMS-1; M = Mg, KMS-2). The Cs and UO22+ ion exchange capacity of KTS-3 (qm = 226 mg g−1 for Cs+ and 382 mg g−1 for UO22+) is comparable with KMS-1 (qm = 280 mg g−1 for Cs+ and 287 mg g−1 for UO22+) and the Cs+ ion exchange capacity is much lower than KMS-2 (qm = 532 mg g−1 for Cs+). However, KTS-3 (qm = 102 mg g−1 for Sr2+) outperforms both KMS-1 (qm = 77 mg g−1 for Sr2+) and KMS-2 (qm = 87 mg g−1 for Sr2+) in terms of Sr2+ ion exchange capacity. Moreover, the relative ease and inexpensive synthesis of K2xSn4−xS8−x make it a promising material for future studies.
Our work shows that the metal chalcogenide family can provide promising ion exchange materials for the selective removal of radionuclide from nuclear waste. Further work is to assess the utility of KTS-3 in remediation applications of nuclear wastes is justified.
Footnote |
† Electronic supplementary information (ESI) available: Raman spectra, thermogravimetric analysis, scanning electron microgram, X-ray crystallographic file (CIF) containing crystallographic refinement details, atomic coordinates with equivalent isotropic displacement parameters, anisotropic displacement parameters, and selected bond distances for KTS-3. See DOI: 10.1039/c5sc03040d |
This journal is © The Royal Society of Chemistry 2016 |