Tatiana G. Levitskaia*a,
Sayandev Chatterjeea,
Bruce W. Areya,
Emily L. Campbella,
Yongchun Hongb,
Libor Kovarikb,
James M. Petersona,
Natasha K. Pencea,
Jesus Romeroa,
Vaithialingam Shutthanandanb,
Birgit Schwenzerc and
Tamas Vargab
aEnergy and Environment Directorate, Pacific Northwest National Laboratory, Richland, WA 99354, USA. E-mail: Tatiana.Levitskaia@pnnl.gov
bEnvironmental and Molecular Sciences Laboratory, Pacific Northwest National Laboratory, Richland, WA 99354, USA
cPhysical and Computational Sciences Directorate, Pacific Northwest National Laboratory, Richland, WA 99354, USA
First published on 18th July 2016
The radioactive contaminant iodine-129 (I-129) is one of the top risk drivers at radiological waste disposal and contaminated groundwater sites where nuclear material fabrication or reprocessing has occurred. However, currently there are very few options available to treat I-129 in the groundwater, partially related to the complex biogeochemical behavior of iodine in the subsurface and occurrence of I-129 in the multiple chemical forms. We hypothesize that layered hydrotalcite materials containing redox active transition metal ions offer a potential solution, benefiting from the simultaneous adsorption of iodate (IO3−), and iodide (I−) anions, which exhibit different electronic and structural properties and therefore may require dissimilar hosts. To test this hypothesis, Cr3+-based materials were selected based on the rationale that Cr3+ readily reduces IO3− in solution. It was combined with either redox-active Co2+ or redox-inactive Ni2+ so that two model materials were prepared by hydrothermal synthesis including Co2+–Cr3+ or Ni2+–Cr3+ (M–Cr). Obtained M–Cr materials composed of Co2+–Cr3+ or Ni2+–Cr3+ layered hydrotalcite and small fractions of Co3O4 spinel or Ni(OH)2 theophrastite phases were structurally characterized before and after uptake of periodate (IO4−), IO3−, and I− anions. It was found that the IO3− uptake is driven by its chemical reduction to I2 and I−. Interestingly, in the Co2+–Cr3+ hydrotalcite, Co2+ and not Cr3+ serves as a reductant while in the Ni2+–Cr3+ hydrotalcite Cr3+ is responsible for the reduction of IO3−. A different uptake mechanism was identified for the IO4− anion. The Co2+–Cr3+ hydrotalcite phase efficiently uptakes IO4− by a diffusion-limited ion exchange mechanism and is not accompanied by the redox process, while Cr3+ in the Ni2+–Cr3+ hydrotalcite reduces IO4− to IO3−, I2 and I−. Iodide exhibited high affinity only to the Co–Cr material. The Co–Cr material performed remarkably well for the removal of IO3−, I− and total iodine from the groundwater collected from the US DOE Hanford site, WA, USA, outperforming non-redox active hydrotalcites (e.g., Mg2+–Al3+) reported previously. This work demonstrates that redox-controlled sorption can be a highly effective method for the treatment of anions based on elements with mobile oxidation states. Further, multiple anions of interest could be simultaneously removed through a combination of approaches.
For several decades considerable research effort has been directed toward identifying synthetic materials for removing or attenuating transport of iodine in the groundwater.8–10 Clay-like hydrotalcite materials commonly known as layered double hydroxides (LDHs) are attractive candidates as they have been extensively studied for the removal of harmful anions from contaminated water.11 The LDH materials can be represented by the general formulae [M1−x2+Mx3+(OH)2]x+(An−)x/nyH2O, where M2+ is a divalent and M3+ is a trivalent cation; the value of x is equal to the molar ratio of M3+/(M2+ + M3+), whereas An− is the interlayer anion. These materials usually uptake anions through ion exchange. Variation of the identities of M2+, Mm+, and An− give rise to a large class of isostructural materials with physicochemical properties that can be tailored to achieve high selectivity and capacity for the uptake of an anion of interest12 and also be regenerated for reuse.13 Application of hydrotalcites for the selective sorption of various anions of iodine has been recently reviewed.8,9 Among tested hydrotalcites, Mg2+–Al3+ materials showed promise for iodine removal and have been preferentially investigated. For instance, iodide has been immobilized using Mg2+–Al3+ LDH by exchanging the parent Cl− anion with I−.14 In another study, calcined Mg2+–Al3+–CO32− material was used to remove iodide from contaminated ground water.15 Hydrotalcite-type compounds consisting of brucite-like positively charged layers with the similar Mg2+–Al3+–CO32−–NO3− chemical composition showed promise for IO3− uptake.16,17 The main drawback of the Mg2+–Al3+ LDH materials is insufficient selectivity toward iodide and iodate anions. Several studies demonstrate that iodine removal efficiency of uncalcined Mg2+–Al3+ materials is limited by their high affinity toward carbonate12,13 present in groundwater, raising concerns regarding their applicability for environmental separations.8,9 Presence of other anions also affects iodine removal. Kang et al. reported that Mg2+–Al3+ hydrotalcite exhibits high efficiency of iodide uptake in competing anion free solutions.18 However, when this and other hydrotalcites were tested in dilute sodium salt (nitrate, nitrite, aluminate, hydroxide) solutions the iodide uptake was significantly reduced.19,20 A similar effect was observed for iodide adsorption using Zn-bearing hydrotalcites in sodium chloride solutions.21
One potential approach to increase LDH selectivity toward iodine adsorption is to introduce soft polarizable metal ions into the LDH structure which have tendency to strongly bind soft bases such as iodine anions.22 Indeed, Hg and Ag-bearing sulfide minerals demonstrate high affinity for iodine anions as reported in the review by Mattigold et al. (and references therein),10 however toxicity limits their utilization for groundwater treatment. We hypothesize that the presence of the borderline soft redox-active metal ions in the LDH structure will have a weaker but similar effect and may improve their selectivity toward adsorption of the iodine anions. Another objective of this work is to examine if the application of LDH material containing redox-active transition metals is beneficial for selective uptake of the iodine anions due to their potential conversion to the oxidation state and chemical form most complimentary with the electronic and structural environment of the LDH material. Introduction of a secondary redox uptake mechanism in addition to anion exchange can be particularly useful in designing hydrotalcite materials for the simultaneous adsorption of iodate and iodide anions, which exhibit different electronic and structural properties and therefore may require dissimilar hosts. In fact, this can pave the way for a multi-modal approach to removal of multiple anions, where some of the anions can be removed by a simple ion-exchange mechanism, while the others can be removed through more complicated chemical or redox pathways. As proof-of-concept, we are examining borderline soft, redox-active transition metals such as Co2+ and Cr3+ for the selective uptake of anions of iodine, and their subsequent redox change. Indeed, our previous exploratory work has indicated that among the large array of LDH compounds synthesized and tested for their affinity toward iodate and iodide, Cr3+-containing hydrotalcite materials exhibited effective IO3− uptake from the Hanford groundwater at high sorption capacity.23 These scoping results prompted us to elucidate and compare uptake mechanisms for iodine anions using two materials based on Co2+–Cr3+ and Ni2+–Cr3+ (referred in the text as Co–Cr and Ni–Cr or jointly as M–Cr) as model redox-active compounds containing a combination of the borderline soft Co2+ and Ni2+ divalent transition metal ions. In addition, Co2+ can be easily oxidized to Co3+ while Ni2+ is expected to retain a +2 oxidation state; this difference allows elucidation of the role of the model redox mobile Cr3+ in the uptake of the iodine anions. It is understood that Cr-based materials are unlikely candidates for the practical applications due to the toxic properties of Cr(VI). Rather it serves as a convenient redox-active model allowing benchmarking of redox-driven ion exchange mechanisms and design of environmentally friendly advanced hydrotalcite materials to manage challenging redox mobile subsurface contaminants that include not only iodine, but As, Hg, Tc, and others as well. In this work, the mechanism and sorption behavior of M–Cr toward three different anionic species of iodine including IO4−, IO3− and I− was explored. Even though periodate is anticipated to be the only minor iodine species in the subsurface, its investigation is essential to understand redox behaviour of iodine upon incorporation in the hydrotalcite structure.
XRD patterns of the samples were recorded on a Philips X'pert Multi-Purpose Diffractometer (MPD) (PANAlytical, Almelo, The Netherlands) equipped with a fixed Cu anode operating at 45 kV and 40 mA. XRD patterns were collected in the 5–100° 2θ-range with 0.04° steps at a rate of 5 s per step. Phase identification was performed using JADE 9.5.1 from Materials Data Inc., and the 2012 PDF4+ database from ICSD. The lattice parameters and volume-averaged crystallite sizes were calculated from whole-pattern fitting TOPAS v5 (Bruker AXS GmbH, Germany).
XPS data were recorded on a Phi5000 Versa Probe system equipped with a monochromatic Al Kα X-ray source (1486.7 eV) and a hemispherical analyzer. Powder samples were mounted using double-sided carbon conductive tape attached to a stainless steel sample holder. The instrument was calibrated for Cu (2p3/2) at 932.6 (±0.1 eV) with the FWHM (full with at half maximum) of 0.98 eV. The surface charge was eliminated by charge neutralizer and correction of data referencing the 284.5 eV C 1s peak. The percentages of individual elements detected were determined from the relative composition analysis of the peak areas of the bands on the basis of the relative peak areas and their corresponding sensitivity factors to provide relative compositions. XPS peak fitting was done using the software XPSPEAK41 with Shirley type background and 20% GL (Gaussian–Lorentzian ratio). The satellite peaks were fitted with parameters given in literature.27,28
SEM analysis was performed using the FEI Quanta 3DFEG Dual Beam microscope operated at 10–20 kV. The samples were prepared by two independent methods; in the first method, the sample particles were dispersed onto carbon tape and coated with ∼5 nm of carbon to minimize charge effects. In the second method, samples were mounted on the tape and polished using typical metallographic techniques to avoid colloidal silica for polishing. Compositional analysis was performed with Oxford 80 mm2 SDD electron dispersive spectrometry (EDS) detector. For quantitative EDS analysis, calculated K factors provided by INCA software were used. No correction for absorption within the specimen was performed. Both secondary electron images (SE) and backscatter electron images (BSE) were recorded.
TEM analysis was performed using the FEI Titan 80-300 operated at 300 kV. The information limit of the microscope was below 0.1 nm. Imaging of the focused ion beam (FIB)-prepared samples was done under low electron dose conditions to minimize beam-induced damage. All images were digitally recorded using a CCD camera and analyzed using Gatan digital micrograph. Compositional analysis was performed with EDAX Si(Li) EDS detector.
![]() | (1) |
Iodine analysis was done using ICP-MS. The instrument was calibrated using high-purity National Institute of Standards and Technology (NIST)-traceable calibration standards to generate calibration curves and verify continuing calibration during the analytical run. All standard preparations and sample dilutions were done using a 0.5% Spectrasol solution CFA-C to prevent build-up/memory of iodine in the ICP-MS introduction system.
IR spectra of the obtained solid materials exhibited a moderately strong band at ∼1350 cm−1 as well as bands in the 820–850 cm−1 region which are assigned to the asymmetric stretching and the out-of-plane deformation of free interlamellar carbonate anions with D3h symmetry in the composites.31 Additional bands observed at 1480–1490, 1388 (shoulder), and 1067 cm−1 are also attributed to intercalated carbonate ions, but in the lower C2v symmetry,32,33 suggesting binding of these carbonates by the hydrotalcites. A broad band centered at ca. 3440 cm−1 is due to the νO–H stretching vibration of hydrogen-bonded geminal OH groups and water molecules, and that at 1636 cm−1 corresponds to the bending of water.34 Other IR bands appeared at the 500–700 cm−1 spectral region, and were assigned to the vibrations of the metal–oxygen bonds.35 Interestingly no presence of nitrate anion in the prepared materials was observed by IR spectroscopy suggesting preferential incorporation of the carbonate anion available due to the aerated conditions of the synthesis.
XRD studies of the materials were conducted to identify the structure of the composites. The XRD patterns of the Co–Cr and Ni–Cr materials (Fig. 1A and B) are characterized by the presence of two distinct phases, one phase being significantly more crystalline than the other, so that M–Cr have composite structures. The diffractograms of the highly crystalline phases in the Co–Cr and Ni–Cr materials are consistent with the respective structures of the cubic spinel Co3O4 (JCPDS no. 09-0418)36 and theophrastite, the anhydrous form of Ni(OH)2 (JCPDS no. 59-0463).36 Based on the relative intensities, it should be noted that while a significant fraction of the cubic spinel Co3O4 is present in the Co–Cr composite, the Ni(OH)2 phase constitutes only a tiny portion of the Ni–Cr composite. The second phase in both composite materials exhibits a nearly identical broad and highly polycrystalline diffraction pattern that is strikingly similar to the previously reported hydrotalcite composite of the formula Mg6Al2CO3(OH)16·4H2O (JCPDS no. 14-0191)36 characterized by a layered structure with intercalated CO32− anions. The diffraction pattern of this polycrystalline phase also matches that of previously observed hydrotalcite-like Co–Cr and Ni–Cr carbonates obtained by the coprecipitation method.32,33,37 The XRD profiles of the polycrystalline phases obtained in this work are sharper and more intense than ones obtained by the simple co-precipitation method as illustrated in above reports by del Arco et al. and Kooli et al.,32,33,37 which is attributed to the improved crystallinity of the hydrotalcite products resulting from the prolonged hydrothermal syntheses. In unison with literature, the XRD results indicate that the hydrotalcite phases of the prepared M–Cr composites possess highly layered structural framework.
![]() | ||
Fig. 1 XRD patterns of the M–Cr materials before and after exposure to the aqueous solution of 10−1 M of I−, IO3−, or IO4−. (A) Co–Cr, (B) Ni–Cr. |
XPS measurements were conducted to identify the oxidation states associated with the key components of the prepared composites. Due to the high redox mobility of Co and Cr transition metals (at pH = 14, standard electrode potentials of the Co3+/Co2+ and CrO42−/Cr(OH)3 couples are 0.17 V and −0.13 V, respectively38), oxidation of the starting Co2+ and Cr3+ ions can be expected during the hydrothermal synthesis. The obtained XPS spectra for the M–Cr composites are shown in Fig. 2 (black traces), observed atomic compositions and the binding energy assignments are given in Tables 1 and S2,† respectively. For the Co–Cr composite, the Co 2p3/2 spectrum shows a broad peak, which envelops the 780.9 and 779.8 eV regions corresponding to Co2+ (87%) and Co3+ (13%), respectively (Fig. 2 and Table 2).39,40 Likewise, Cr 2p1/2/2p3/2 spectral region shows bands with the maxima at 586.8 and 577.2 eV suggesting presence of Cr3+ (77%)41 as well as shoulders at 588.7 and 579.2 eV suggesting presence of Cr6+ (23%)42 (Fig. 2 and Table 2). The O 1s spectrum (Fig. S1A,† black trace; Table S2†) shows a broad peak enveloping bands at 531.4 eV assigned to a mixed +2/+3 cobalt oxide43 and 530.7/530.6 eV assigned to Cr2O3 (ref. 44) and CrO3.42 The low intensity Cl 2p3/2 spectrum at 198.7 eV (ref. 45) is assigned to the presence of chloride in the hydrotalcite phase. For the Ni–Cr material, the Ni 2p3/2 spectrum is attributed to Ni(OH)2 (Fig. 2 and Table 2).28 The Cr spectral region is nearly identical to that of observed for the Co–Cr material and indicates that Cr3+ and Cr6+ contribute respective 78 and 22% to the overall intensity (Fig. 2 and Table 2).41,42 The O 1s spectrum (Fig. S1B,† black trace; Table S2†) shows a broad peak as well, centered around 530.7/530.6 eV assigned to Cr2O3 (ref. 44) and CrO3 (ref. 42) with a shoulder at 532.0 eV corresponding to Ni(OH)2.46
![]() | ||
Fig. 2 XPS spectral overlays for the M–Cr materials before (black traces) and after 24 hour exposure to 10−1 M I− (violet traces), IO3− (red traces) or IO4− (orange traces). (A) Co 2p3/2 region for Co–Cr, (B) Cr 2p1/2 and 2p3/2 regions for Co–Cr, (C) I 3d5/2 region for Co–Cr, (D) Co 2p3/2 region for Ni–Cr, (E) Cr 2p1/2 and 2p3/2 regions for Ni–Cr, (F) I 3d5/2 region for Ni–Cr. The respective composition percentages obtained from deconvoluting the spectra are presented in Table 2. |
Composite | C 1s | O 1s | Cl 2p | Co 2p | Cr 2p | I 3d |
---|---|---|---|---|---|---|
Co–Cr | 24.3 | 50.7 | 3.2 | 17.2 | 4.5 | NA |
Co–Cr/I− | 23.4 | 52.4 | 1.4 | 17.4 | 4.8 | 0.6 |
Co–Cr/IO3− | 27.0 | 50.6 | 1.5 | 15.5 | 4.2 | 1.3 |
Co–Cr/IO4− | 26.0 | 51.4 | 1.7 | 14.8 | 3.1 | 3.1 |
Composite | C 1s | O 1s | Ni 2p | Cr 2p | I 3d |
---|---|---|---|---|---|
Ni–Cr | 27.1 | 52.8 | 16.6 | 3.4 | NA |
Ni–Cr/I− | 21.7 | 55.6 | 18.9 | 3.2 | 0.0 |
Ni–Cr/IO3− | 23.0 | 54.3 | 17.0 | 3.9 | 1.7 |
Ni–Cr/IO4− | 23.5 | 53.9 | 16.8 | 3.6 | 2.3 |
Sample | Co% | Cr% | I% | |||||
---|---|---|---|---|---|---|---|---|
Co2+ | Co3+ | Cr3+ | Cr6+ | IO4− | IO3− | I2 | I− | |
Co–Cr | 87 | 13 | 77 | 23 | NA | |||
Co–Cr–I− | 91 | 9 | 76 | 24 | — | — | — | 100 |
Co–Cr–IO3− | 78 | 22 | 75 | 25 | — | 46 | 26 | 28 |
Co–Cr–IO4− | 87 | 13 | 80 | 20 | 100 | — | — | — |
Sample | Ni% | Cr% | I% | ||||
---|---|---|---|---|---|---|---|
Ni2+ | Cr3+ | Cr6+ | IO4− | IO3− | I2 | I− | |
Ni–Cr | 100 | 78 | 22 | NA | |||
Ni–Cr–I− | 100 | 70 | 30 | — | — | — | — |
Ni–Cr–IO3− | 100 | 69 | 31 | — | 44 | 24 | 32 |
Ni–Cr–IO4− | 100 | 60 | 40 | 48 | 27 | 13 | 12 |
SEM images of the Co–Cr composite (Fig. 3A) shows powdery particles that aggregate together to form a cake-like heterogeneous matrix. The powdery morphology of the composite indicates a large surface area facilitating the uptake of anions. Upon further magnification, microcrystalline granules embedded onto the heterogeneous matrix are observed. EDS measurements indicate that the composition of the hydrotalcite-like heterogeneous matrix is distinctly different from the granules largely comprised of the Co3O4 spinels (Table 3). The granules are enriched in Co so that Co:
Cr ratio is about 6.4
:
1 compared to the heterogeneous phase with Co
:
Cr ratio of about 2.4
:
1. The hydrotalcite phase showed the presence of Cl, suggesting intercalated chloride anion (Table S3†). These results are further supported by the TEM structural data (Fig. 4), which reveal a two-phase system comprising of particles with cuboidal microcrystalline structure embedded in the heterogeneous matrix. The heterogeneous matrix with significant presence of both Cr and Co, shows a highly diffused diffraction (Fig. 4F) further supporting the polycrystalline structure, characterized by small (>5 nm), highly distorted polycrystalline sheets observed by HAADF imaging (Fig. 4E). The HAADF and diffraction patterns of the Co-enriched cuboids confirm a highly ordered Co3O4 spinel structure with Co2+ and Co3+ being located at the tetrahedral and octahedral sites respectively (Fig. 4C and D).47 Small amounts of Cr and Cl associated with the Co3O4 spinel phase is presumably due to their bleeding from the adjacent heterogeneous matrix. The cuboidal particles are observed to agglomerate and form porous regions (Fig. 4B).
M–Cr | Analyte | Dominant phase | M![]() ![]() ![]() ![]() |
Details of elemental mapping |
---|---|---|---|---|
a The region measured was cracked on the surface. | ||||
Co–Cr | — | Hydrotalcite | 2.4![]() ![]() ![]() ![]() |
Regions 1 and 2 in Fig. S2, Table S3 |
Spinel | 6.4![]() ![]() ![]() ![]() |
Regions 3, 4 and 5 in Fig. S2, Table S3 | ||
Ni–Cr | — | Hydrotalcite | 2.6![]() ![]() ![]() ![]() |
Regions 6 and 7 in Fig. S2, Table S3 |
Theophrasite | 6.3![]() ![]() ![]() ![]() |
Region 8 in Fig. S2, Table S3 | ||
Co–Cr | IO4− (24 hours) | Hydrotalcite | 2.5![]() ![]() ![]() ![]() |
Region 1 in Fig. S4, Table S4 |
Spinel | 6.1![]() ![]() ![]() ![]() |
Region 2 in Fig. S4, Table S4 | ||
Not identifieda | 1.9![]() ![]() ![]() ![]() |
Region 3 in Fig. S4, Table S4 | ||
IO4− (7 days) | Hydrotalcite | 1.7![]() ![]() ![]() ![]() |
Regions 5 and 6 in Fig. S4, Table S4 | |
Spinel | 6.9![]() ![]() ![]() ![]() |
Region 4 in Fig. S4, Table S4 | ||
IO3− (24 hours) | Hydrotalcite | 2.7![]() ![]() ![]() ![]() |
Regions 1 and 2 in Fig. S6, Table S5 | |
Spinel | 6.6![]() ![]() ![]() ![]() |
Region 3 in Fig. S6, Table S5 | ||
IO3− (7 days) | Hydrotalcite | 1.7![]() ![]() ![]() ![]() |
Regions 5 and 6 in Fig. S6, Table S5 | |
Spinel | 6.7![]() ![]() ![]() ![]() |
Region 4 in Fig. S6, Table S5 | ||
I− (24 hours) | Hydrotalcite | 2.6![]() ![]() ![]() ![]() |
Regions 1 and 2 in Fig. S8, Table S6 | |
Spinel | 6.4![]() ![]() ![]() ![]() |
Regions 3 and 4 in Fig. S8, Table S6 | ||
I− (7 days) | Hydrotalcite | 2.5![]() ![]() ![]() ![]() |
Regions 5 and 6 in Fig. S8, Table S6 | |
Ni–Cr | IO4− (24 hours) | Hydrotalcite | 2.6![]() ![]() ![]() ![]() |
Regions 1 and 3 in Fig. S10, Table S7 |
Theophrasite | 6.7![]() ![]() ![]() ![]() |
Region 2 in Fig. S10, Table S7 | ||
IO4− (7 days) | Hydrotalcite | 5.5![]() ![]() ![]() ![]() |
Regions 4 and 5 in Fig. S10, Table S7 | |
IO3− (24 hours) | Hydrotalcite | 2.6![]() ![]() ![]() ![]() |
Regions 1 and 2 in Fig. S11, Table S8 | |
Theophrasite | 7.0![]() ![]() ![]() ![]() |
Region 3 in Fig. S11, Table S8 | ||
IO3− (7 days) | Hydrotalcite | 5.1![]() ![]() ![]() ![]() |
Regions 4–6 in Fig. S11, Table S8 | |
I− (24 hours) | Hydrotalcite | 2.8![]() ![]() ![]() ![]() |
Region 2 in Fig. S12, Table S9 | |
Theophrasite | 6.4![]() ![]() ![]() ![]() |
Regions 1 and 3 in Fig. S12, Table S9 | ||
I− (7 days) | Hydrotalcite | 4.4![]() ![]() ![]() ![]() |
Regions 4 and 5 in Fig. S12, Table S9 |
SEM image of the Ni–Cr composite (Fig. 3E) shows a cake-like morphology similar to the Co–Cr composite, displaying microcrystalline granular particles being embedded onto a hydrotalcite like heterogeneous matrix. The Ni:
Cr ratio in the heterogeneous matrix and in the granules is about 2.6
:
1 and 6.3
:
1, respectively (Table 3). The microscopy results are consistent with the X-ray diffractograms for both M–Cr composites and demonstrate the presence of two distinct phases in each material, the highly crystalline phase being consistent with a cubic spinel Co3O4 for Co–Cr material or with Ni(OH)2 theophrastite phase for Ni–Cr, while the less crystalline heterogeneous phase is consistent with a hydrotalcite like structure.
Overall, based on the elemental analysis XPS, and microscopy results, as well as structural similarity of the heterogeneous M2+–Cr3+ phases of the materials obtained in this work with the corresponding literature hydrotalcite carbonates,30 it is concluded that hydrotalcite phases mainly have general composition of [M1−xCrx(OH)2]Ax/2·nH2O, where A is intercalated anion; predominantly carbonate for Ni2+–Cr3+ and carbonate/chloride mixture for Co2+–Cr3+. These polycrystalline phases incorporate nearly entire stoichiometric amount of Cr used in the synthesis of M–Cr materials, with a negligible amounts getting incorporated in the second crystalline phases (as evidenced by the XPS and TEM results above).
Composite | Crystallite sizes, nm. Standard deviation is given in the parenthesis | |||
---|---|---|---|---|
Untreated | IO4− treated | IO3− treated | I− treated | |
Co–Cr | 3.6(1) | 1.3(2) | 6(1) | 3.4(2) |
Ni–Cr | 2.9(1) | 1.7(2) | 2.8(1) | 2.6(1) |
Exposure of Ni–Cr composite to IO4−, IO3− and I− for 24 hours results in lowering of intensity of the XRD pattern attributed to Ni(OH)2, along with broadening of the hydrotalcite phase pattern (Fig. 1B) indicating decrease in crystallinity of the entire composite material upon exposure to these anions. Upon prolonged exposure to the iodine anions for 7 days, the hydrotalcite phase is observed to further reduce crystallinity, while the Ni(OH)2 phase is completely absent suggesting that both phases are involved in the iodine uptake. As in the case of Co–Cr, periodate incorporation reduces the planar distance within the layers of the Ni–Cr hydrotalcite phase indicating the strong interaction between this anion and Ni–Cr cationic layers (Table 4). On the other hand, no significant changes are observed in the crystallite sizes of this phase upon IO3− or I− exposure.
The Ni–Cr composite exhibited high affinity for IO4− and IO3− anions but uptake of I− was negligible (Fig. 2D–F and Table 1). Sorption of IO4− and IO3− anions were accompanied by their partial reduction to IO3−/I2/I− and I2/I−, respectively, and oxidation of Cr3+ to Cr6+, whose content increased by about 10–20% (Table 2). As expected, there was no change in the oxidation state of Ni2+ upon iodine exposure.
SEM analysis on Co–Cr demonstrates that exposure of the composite to IO4−, IO3− or I− for 24 hours results in noticeable changes in the morphology of the heterogeneous phase (Fig. 3B–D). Exposure to IO4− results in pancake like morphology where the individual powdery units are stuck together. Exposure to IO3− makes the matrix more compact while giving it a coarser landscape. Exposure to I− generates a crust-like morphology. Such significant morphological changes support chemisorption of iodine by the heterogeneous matrix.
Elemental mapping of 24 hour IO4− exposed Co–Cr composites shows no presence of iodine in the bulk material (Table 3). However, elemental analysis of the cracks on the surface shows a significant concentration of iodine. The TEM analysis (Fig. S5†) is consistent with the SEM results and demonstrates that neither Co3O4 nor Co–Cr hydrotalcite phases incorporate IO4− during first 24 hours of exposure. This, in unison with the fact that XPS of IO4−-exposed Co–Cr composite exhibits a significant fraction of iodine suggests that initially sorption of IO4− is surface-limited.
Elemental analysis of 24 hour IO3−-exposed Co–Cr composite demonstrates a high concentration of iodine in the regions dominated by the hydrotalcite matrix, while only small iodine amounts are observed to be associated with spinels (Table 3). This is further illustrated by elemental mapping, suggesting that IO3− has high affinity to the polycrystalline matrix, with the Co3O4 cuboids showing insignificant, presumably surface, uptake. The TEM analysis of IO3−-exposed Co–Cr composite (Fig. S7†) confirms the SEM/EDS results (Fig. S6†) and reveals that the elemental composition and morphology of Co3O4 spinels remain largely intact upon IO3− exposure.
Elemental analysis of 24 hour I−-exposed Co–Cr composites reveals that both the polycrystalline hydrotalcite phase and the spinels demonstrate affinity to iodide, with the affinity of the hydrotalcite phase being slightly higher (Table 3). TEM analysis of the polycrystalline layer (Fig. S9†) shows small (∼1 nm) particles dispersed within the matrix. High diffraction contrast suggests that these are amorphous iodide-containing particles. Elemental mapping also indicates a significant absorption of iodine around the spinel nano-particles (Table 3). As discussed above, this high affinity of the Co3O4 phase towards I− is also reflected in its complete disappearance from the X-ray diffractogram of Co–Cr composite after 7 day exposure to I− due to its presumable reaction with I− (Fig. 1A).
The microscopy of the Co–Cr composite after exposure to IO4− or IO3− for 7 days displays very little visible change either in heterogeneous matrix or the cuboidal granules (Fig. S3†). Elemental mapping results reveal that the composition of the cuboidal spinel phase remains almost invariant upon prolonged exposure to these anions, with slight incorporation of iodine presumably due to surface sorption. On the other hand, significant changes are observed in the elemental composition of the hydrotalcite phase (Table 3). This result is consistent with X-ray diffractograms (Fig. 1A) demonstrating that Co3O4 phase is not affected by prolonged exposure to these anions, while crystallinity of the heterogeneous hydrotalcite phase is greatly reduced. It is noteworthy that while there was no incorporation of IO4− in the Co–Cr hydrotalcite phase during first 24 hour of exposure, significant IO4− incorporation occurs in 7 days. This is indicative of a diffusion-limited ion exchange mechanism. In case of exposure to IO3−, fast incorporation of this anion into the hydrotalcite phase occurs so that significant iodine loading is achieved within first 24 hours and only small increase of iodine loading is measured during the following days. These points out a different sorption mechanism involving chemical or redox interaction between IO3− and hydrotalcite phase. This redox mechanism is supported by the XPS results. The elemental mapping of the Co–Cr composite exposed to I− for 7 days shows the disappearance of Co3O4 phase. For example, region 5 in the microscopic profile that resemble granular particles (Fig. S8†), shows near identical elemental composition as that of the hydrotalcite matrix (region 6 in Fig. S8† and Table 3), thereby supporting the chemical reaction of Co3O4 phase with I−. This, in unison with the XRD results on the Co–Cr composite after 7 day I− exposure suggests that I− uptake by the composite is associated with a chemical reaction of I− with both the hydrotalcite as well as the Co3O4 phases of the composite.
SEM analysis performed on the Ni–Cr composite exposed to 10−1 M IO4−, IO3− or I− for 24 hours shows very little visible changes in the morphologies either in the polycrystalline matrix or on the embedded granules (Fig. 3F–H). However, the elemental mapping clearly indicates prominent changes in the elemental composition upon exposure to IO4− and IO3− resulting in iodine incorporation in the composite.
Elemental mapping of the 24 hour IO4−-exposed Ni–Cr composite shows almost equal distribution of iodine between the bulk hydrotalcite phase and the granules enriched with theophrasite phase (Table 3). On the other hand, elemental mapping of the 24 hour IO3− exposed composite shows a greater tendency of the absorbed iodine to be distributed in the bulk hydrotalcite matrix as compared to the granules matrix. Elemental analysis of 24 hour I−-exposed Ni–Cr show no incorporation of iodine either on the granules or within the heterogeneous matrix (Table 3) confirming weak affinity of this composite toward uptake of the I− ions observed by XPS.
The elemental mapping of the Ni–Cr composite after exposure to IO4− or IO3− for 7 days shows complete disappearance of the Ni(OH)2 rich granular phases upon prolonged iodine exposure. While the microscopic profiles shown in Fig. S10† for IO4− and Fig. S11† for IO3− depict profiles similar to the starting materials with granular particles embedded in the hydrotalcite, elemental mapping demonstrates a nearly identical elemental composition of the granules and the hydrotalcite phases (Table S8† for IO4− and Fig. S11† for IO3−), indicating that the granules are merely morphological variations of the matrix, and not a different phase or a different chemical form. Prolonged exposure of the Ni–Cr composite to I− solution did not facilitate sorption of this anion (Fig. S12 and Table S9†).
Iodine concentration in groundwater, μg L−1 | Kd(Iodine) mL g−1 | Percent iodine sorbed, % | ||||
---|---|---|---|---|---|---|
Initial | Post contact | Co–Cr | Ni–Cr | Co–Cr | Ni–Cr | |
Co–Cr | Ni–Cr | |||||
Iodate (IO3−) | ||||||
130![]() |
48![]() |
37![]() |
363 | 528 | 63 | 71 |
96![]() |
32![]() |
24![]() |
424 | 615 | 67 | 74 |
62![]() |
18![]() |
13![]() |
507 | 740 | 70 | 78 |
32![]() |
7340 | 6090 | 740 | 933 | 77 | 81 |
12![]() |
2420 | 1600 | 885 | 1450 | 81 | 86 |
10![]() |
1970 | 1130 | 888 | 1680 | 80 | 87 |
6400 | 1190 | 676 | 937 | 1810 | 81 | 89 |
2330 | 375 | 314 | 1120 | 1890 | 83 | 90 |
1240 | 212 | 133 | 1020 | 1770 | 86 | 89 |
1010 | 137 | 88 | 1350 | 2240 | 84 | 91 |
613 | 99 | 58 | 1090 | 2050 | 84 | 91 |
130 | 18.2 | 13 | 1320 | 1950 | 86 | 90 |
74 | 8.9 | 7.4 | 1540 | 1940 | 88 | 90 |
![]() |
||||||
Iodide (I−) | ||||||
120![]() |
83![]() |
104![]() |
94 | 33 | 31 | 13 |
79![]() |
58![]() |
78![]() |
77 | 2 | 26 | 1 |
54![]() |
28![]() |
46![]() |
199 | 37 | 48 | 15 |
27![]() |
2780 | 24![]() |
1900 | 27 | 90 | 11 |
13![]() |
17.6 | Not measured | 156![]() |
— | ∼100 | — |
10![]() |
5.7 | 9890 | 401![]() |
20 | ∼100 | 8 |
8710 | 12.5 | 7550 | 147![]() |
33 | ∼100 | 13 |
5480 | 5.5 | 5130 | 212![]() |
15 | ∼100 | 6 |
2910 | 5.4 | 2390 | 113![]() |
46 | ∼100 | 18 |
1050 | 4.3 | 948 | 51![]() |
23 | ∼100 | 10 |
837 | 2.3 | 743 | 79![]() |
27 | ∼100 | 11 |
584 | 2.1 | Not measured | 57![]() |
— | ∼100 | — |
292 | 2.2 | 240 | 28![]() |
46 | 99 | 18 |
116 | 1.8 | 98 | 13![]() |
41 | 98 | 16 |
59 | 1.6 | 51 | 7800 | 32 | 97 | 13 |
![]() |
||||||
Total iodine in unmodified groundwater | ||||||
8.6 ± 0.9 | 1.5 | 3.1 | 1010 | 390 | 82 | 64 |
Both M–Cr composite materials exhibited efficient sorption of the IO3− anion. The Ni–Cr composite was found to be more effective sorbent as evident from the Kd(IO3−) values across entire IO3− concentration range. The maximum Kd(IO3−) values observed for the Co–Cr and Ni–Cr composites are about 1500 and 2200 mL g−1, respectively. As the IO3− concentration in the groundwater decreases, Kd(IO3−) values first increase due to the reduced loading of the sorbents and then gradually plateau. Overall it is observed that Co–Cr and Ni–Cr composites remove respective 80–88 and 85–90% IO3− from the groundwater initially containing 74–12500 μg L−1 IO3−. This confirms high selectivity of the M–Cr materials for IO3− anion as the groundwater contains large access of other anions including nitrate (317
000 μg L−1), chloride (181
000 μg L−1), carbonate (116
000 μg L−1), and sulfate (50
000 μg L−1) (Table S1†).
The Co–Cr composite demonstrated very high affinity and selectivity for I−, so that its nearly quantitative removal from the groundwater was observed for the concentration below 13000 mL g−1 (Table 5) The Ni–Cr exhibited poor sorption of iodide as reflected by low Kd(I−) values. This result is consistent with the XPS and SEM/EDS measurements demonstrating absence of iodine in the I− exposed Ni–Cr.
The performance of the inorganic composite materials tested for the uptake of iodine from the unmodified Hanford groundwater is of particular interest and can help to predict the removal of I-129 isotope. The Kd(iodine) values for the total iodine obtained using the unmodified groundwater generally follow the combination of the IO3− and I− sorption profiles. To this end, the Co–Cr composite exhibited a Kd(Iodine) value of ∼1000 mL g−1 and removed 82% of total iodine. Ni–Cr composite exhibited modest performance with Kd(Iodine) value of ∼400 mL g−1 and removal of 64% of total iodine. These results can be attributed to the iodine speciation in the Hanford groundwater. Previous literature indicates that the Hanford groundwater contains predominantly IO3−, accounting for up to 70% of total iodine.4 The minor iodine species include iodide and organoiodine. It is presumed that Co–Cr composite, which has high affinity for both IO3− and I− anions, effectively uptakes them from groundwater accounting for 82% removal of total iodine. The iodine not removed from groundwater most likely consists of organoiodine species. This is consistent with the Ni–Cr performance, which exhibits high affinity for IO3− but not I− anion, and removes only 64% of total iodine from the groundwater.
These results also demonstrate high selectivity of the M–Cr composites towards the anions of iodine (total iodine concentration in groundwater is 8.6 μg L−1, Table S1†) even in presence of the bulk concentrations of coexisting anions including Cl− (181000 μg L−1), NO3− (317
000 μg L−1), CO32− (116
000 μg L−1) and SO42− (50
000 μg L−1) that are present in Hanford ground water as shown in Table S1.† This demonstrates the high selectivity of the composites and emphasizes the importance of the redox mechanism in the uptake of IO3.
Overall sorption efficiency of IO4−, IO3− and I− anions by Co–Cr and Ni–Cr composites is summarized in Table 6. Structural characterization study of the composite materials before and after loading with iodine revealed distinct uptake mechanism for each anion/composite system as evident from the mutually consistent XRD, XPS, and microscopy analyses.
Composite | Anion | Sorption | Variation of Cr oxidation state | Variation of Co or Ni oxidation state | Iodine speciation in the composites |
---|---|---|---|---|---|
Co–Cr | IO4− | Moderate | No change | No change | IO4− |
IO3− | High | No change | Decrease in Co2+ increase in Co3+ | IO3−, I2, I− | |
I− | High | No change | No change | I− | |
Ni–Cr | IO4− | High | Decrease in Cr3+ increase in Cr6+ | No change | IO4−, IO3−, I2, I− |
IO3− | High | Decrease in Cr3+ increase in Cr6+ | No change | IO3−, I2, I− | |
I− | None | No change | No change | — |
The Co–Cr and Ni–Cr hydrotalcite phases showed effective uptake of the IO3− anion, the most abundant chemical species of iodine in the contaminated Hanford groundwater. It is significant to note that IO3− uptake by both Co–Cr and Ni–Cr composites was accompanied by its partial reduction to elemental iodine and iodide. In the Co–Cr hydrotalcite, Co2+ served as a reductant oxidizing to Co3+. Based on the relative reduction potentials for Co(OH)3/Co(OH)2 (0.42 V vs. NHE) and IO3−/I2 (1.2 V vs. NHE) couples, this process is not entirely surprising. However, it is surprising that there is no accompanying oxidation of Cr3+ in this process in spite of IO3− being a strong enough oxidant to oxidize Cr3+ to Cr6+ [E0′(CrO42−/Cr3+) = 0.11 V vs. NHE].38 In fact, when an aqueous solution of Cr(NO3)3 was allowed to react with an aqueous stock solution of IO3− as an experimental control, rapid oxidation of Cr3+ to Cr6+ was observed (data not shown). The fact that a similar oxidation process is not observed in the Co–Cr composite implies that Cr3+ incorporation in composite raises its oxidation potential by such a significant amount that IO3− is no longer able to oxidize it. The uptake of the IO3− anion generates large distortion of the hydrotalcite phase so that its crystallite size is significantly increased (Table 4) presumably due to oxidation of Co2+. The IO3− uptake mechanism by the Ni–Cr hydrotalcite is different, as for this phase the reduction of IO3− was accompanied by oxidation of Cr3+ to Cr6+. Interestingly, the hydrotalcite crystallite size is preserved. Overall these results suggest that chemical reduction is an important driving force in the uptake of IO3− by the redox mobile hydrotalcite phases resulting in highly efficient and selective removal of this anion from such a complex solution matrix as contaminated groundwater. Under similar conditions, redox inactive Ni2+–Al3+ and Mg2+–Al3+ composites demonstrates low affinity for IO3−.51 The specificity of the IO3− uptake by the composites in presence of bulk concentrations of coexisting redox inactive anions NO3−, CO32− and SO42− is illustrated from the uptake data from Hanford groundwater, and further emphasizes the importance of redox affinity for this uptake. The mechanism of the iodate uptake can be characterized as redox-controlled ion exchange.
The M–Cr composites exhibit a different uptake mechanism for IO4− anion. For the Co–Cr composite, SEM/EDS analysis suggests that the IO4− is initially adsorbed on the surface only and slowly penetrates into the core of the material with time. This process is not associated with reduction of IO4− or oxidation of Co2+/Cr3+ indicating a diffusion-limited ion exchange mechanism. On the other hand, uptake of IO4− by the Ni–Cr composite is accompanied by the reduction of IO4− to IO3−, I2 and I− so that the significant fraction of Cr3+ is oxidized to Cr6+.
Iodide exhibited affinity to both hydrotalcite and spinel phases of the Co–Cr composite. On the other hand, Ni–Cr composite failed to incorporate I−. It is an interesting observation that while the hydrotalcite phase of Co–Cr can uptake IO4−, IO3− and I−, the Ni–Cr hydrotalcite phase can uptake only IO4− and IO3−. It is likely that the geometries of IO4− (distorted tetrahedral), IO3− (trigonal pyramidal) and I− (spherical) influence the complimentarily of the uptake.
Both Co–Cr and Ni–Cr composites performed remarkably well for the removal of IO3− from the Hanford groundwater demonstrating exceptional selectivity over other anions common in the subsurface. Only Co–Cr demonstrated effective uptake of I−. Overall the Co–Cr composite removed major fraction of the total iodine from the unmodified Hanford groundwater performing significantly better than previously reported LDH materials in the complex solutions.8–10
This work presents a new approach of multi-anion removal where the sorbent can remove multiple anions of interest through different mechanisms; where some can be removed through a simple ion exchange mechanism while others can be removed through a redox based approach. This work also concludes that Co-based hydrotalcite/spinel composites are highly promising candidates for the selective removal of the dominant inorganic forms of iodine from groundwater, as borderline-soft Co ion present in the crystalline different phases is responsible for the uptake of both IO3− and I− species. This work also emphasizes the importance of designing redox active composite materials preferentially containing dissimilar structures to ensure uptake of various iodine species and maximize removal of total iodine from the groundwater. Layered hydrotalcites containing redox active components activate two uptake mechanisms driven by ion exchange and the redox reactions, greatly enhancing affinity and selectivity toward redox mobile anions. It appeared that for Co–Cr material, Cr does not interact with incorporated iodine maintaining a +3 oxidation state and most likely plays only indirect role in iodine uptake serving as a building block of hydrotalcite phase. Future investigation of Co-based composites containing environmentally friendly trivalent transition metal ions, e.g. Fe3+ and Al3+ is warranted.
Footnote |
† Electronic supplementary information (ESI) available: Composition of 299-W19-36 Well of Hanford, WA; assignment of XPS data as well as additional XPS spectra; microscopy and elemental mapping of the M–Cr composites after exposure to anions of iodine. See DOI: 10.1039/c6ra13092e |
This journal is © The Royal Society of Chemistry 2016 |