Layered A2Sn3S7·1.25H2O (A = Organic Cation) as Efficient Ion

Mar 6, 2017 - PDF. ja7b00565_si_001.pdf (838.61 kB). Citing Articles; Related Content. Citation data is made available by participants in Crossref's C...
3 downloads 14 Views 2MB Size
Subscriber access provided by University of Newcastle, Australia

Communication

The layered (A)2Sn3S7·(A=organic cation) as Efficient Ion-Exchanger for Rare Earth Element Recovery XingHui Qi, Ke-Zhao Du, Mei-Ling Feng, Yu-Jie Gao, Xiao-Ying Huang, and Mercouri G. Kanatzidis J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.7b00565 • Publication Date (Web): 06 Mar 2017 Downloaded from http://pubs.acs.org on March 6, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of the American Chemical Society is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

The layered (A)2Sn3S7·1.25H2O (A = organic cation) as Efficient IonExchanger for Rare Earth Element Recovery Xing-Hui Qi,†, Ke-Zhao Du,†, Mei-Ling Feng,*,†,‡ Yu-Jie Gao,† Xiao-Ying Huang,† and Mercouri G. Kanatzidis*,‡ §

§

†State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou, Fujian 350002, P. R. China ‡Department of Chemistry, Northwestern University, 2145 Sheridan Road, Evanston, Illinois 60208, United States Supporting Information Placeholder ABSTRACT: Exploring new ion-exchangers for the recovery of rare earth elements (REE) and recycling is worthwhile for the high-tech industry and an eco-friendly sustainable economy. The efficient enrichment of low concentration REE from complex aqueous solutions containing large excess of competitive ions is challenging. Here we present a chalcogenide example as a superior REE ion-exchanger efficiently removing them from very complex aqueous solutions, (Me2NH2)1.33(Me3NH)0.67 Sn3S7·1.25H2O (FJSM-SnS). The material exhibits fast and efficient ion exchange behavior with short equilibrium times ( 99%) at low concentrations. Moreover, after ion-exchange the REE in corresponding exchanged products could be easily recovered and the material regenerated by elution. FJSM-SnS has superior capacity and faster absorption kinetics than other states of the artificial REE sorbents such as Al2O3/EG, clay minerals, zeolite and activated carbon.

Rare earth elements (REE) have wide high-tech applications in permanent magnets, luminescent materials, catalysts 1-3 and other fields. Because no decisive substitutes of REE have been found, if REE shortages emerge they will hinder 4 the development of industry fields involved. On the other hand, mining, extraction and roasting stages for REE production involve high energy consumption and have detrimental 5 environmental impact such as acidification and eco-toxicity. 6 Discarding REE-containing wastes is wasteful. REE recovery could help recycling the valuable resources relieving their high environmental burden and easing a potential supply 7, 8 crisis. There is a long standing and continuous research effort to develop methods for highly efficient REE recovery, 9 10-12 13 including precipitation, adsorption, solvent extraction, 14 15 liquid membrane, ionic liquid extraction and biosorp-

16

tion. Most of them can only be used to recover high concentration REE and involve painstaking procedures, high 17, 18 cost and energy consumption. However, low REE concentrations (one to hundreds of ppm) and large excess of non3+ 3+ lanthanide competitive ions such as Al and Fe in rare earth tailings or REE containing waste are more common and 5, 17, make the REE adsorption and separation a desirable task. 19 Thus, efficient adsorbents must have rapid kinetics, high adsorption capacity, wide pH operation range and high selectivity in REE from low concentration solutions. 20 21 Traditional adsorbents including zeolites, active carbon, 22 10-12 share the problems of low capacity, clay and metal oxides poor selectivity, narrow pH range and interference from humic substances. To overcome the limitations of traditional adsorbents, several novel adsorbents have been developed for REE recovery from solution, such as functionalizing gra23 24 18 phene oxide (GO), resins, mesoporous silicates and oxy17 hydroxides. Nevertheless, the efficient enrichment of low concentration REE from the solution containing large excess of competitive ions is still challenging. There is still enormous space for seeking better adsorbents that overcome the existing shortcomings including the slow kinetics aggrega3+ 3+ 17, 23, 24 tion and non-selectivity against Al and Fe . To date, the crystalline metal sulfides have been investigated as excellent ion-exchangers for radioactive elements, 25-29 heavy metals, and actinides removal. However, as far as we know that the application of crystalline sulfides as ionexchangers for the sorption of REE is unexplored. The metal sulfide ion-exchangers possess a rich diversity in composition and structure as well as broad structural flexibility observed 25, 28, 30 upon cation-exchange. Therefore, the crystalline chalcogenides, as a class of materials, deserve investigation as ion-exchangers for REE recovery. In this work, we report the efficient REE recovery performance of a layered thiostannate, (Me2NH2)1.33(Me3NH)0.67 Sn3S7·1.25H2O (FJSM-SnS), which is the first chalcogenide example as a superior REE ion-exchanger from very complex aqueous solutions. Previously, we have shown this material + 2+ to be a very efficient ion-exchanger capturing Cs , Sr and 1

ACS Paragon Plus Environment

Journal of the American Chemical Society 2+

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

31,

UO2 ions which are relevant to nuclear waste remediation. Herein, we carried out comprehensive ion-exchange experiments to determine the influence of operational conditions including contact time, initial ion concentration, competitive ions, pH and bed volumes on the REE recovery. Our results indicate that FJSM-SnS exhibits rapid kinetics, high capacity and good selectivity as a REE ion-exchanger and discrimina3+ 3+ + tion against Al , Fe and Na ions. In addition, the high recovery rate (>99%) in ion-exchange column experiments and the efficient regeneration by elution with KCl solution indicate that FJSM-SnS is an excellent REE trapping material. Yellow hexagonal crystals of FJSM-SnS (Figure S1) were 31 synthesized on large scale as we previously reported. Its 2nstructure is based on a two-dimensional (2D) [Sn3S7]n anionic layer constructed with SnS5 polyhedra in which large windows formed by twenty-four-membered Sn12S12 rings were + + observed (Figure 1a). The [Me3NH] , [Me2NH2] cations and 31 lattice H2O molecules reside at the interlayer space. The 2ninterlayer space between the [Sn3S7]n layers can accommo+ 33 + 34 + 35 date many cations, including Cs , [Me4N] , [Me3N] , + [DABCOH] (DABCOH = protonated 1,836 37 diazabicyclooctane), TBA (TBA = tert-butylamine), + + + 38 [Et4N] ([Et4N] = tetraethylamine) and [NH4] . The structural feature of FJSM-SnS affords the prerequisite for REE enrichment (Figure 1b). 32

Page 2 of 5

which is attributed to the strong coulombic affinity for high 3+ valence state Ln , high species diffusion in the microporous framework, tunable interlayer distance and intralayer win25, 42 dow. For the ion-exchange capacity experiments, the Langmuir 43 isotherm equation was adopted to fit the experimental data by non-linear regression, equation (2).

q = qm

bCe 1 + bCe

(2)

Where qm represents the maximum exchange capacity of FJSM-SnS, and b (L/mg) is the Langmuir constant related to the energy of the adsorption; q (mg/g) is the exchange capacity at the equilibrium concentration Ce (ppm). 2 3+ The fitting result of R = 0.969, 0.981 and 0.976 for Tb , 3+ 3+ Eu and Nd suggests that the Langmuir model is appropriate to depict the REE ion-exchange equilibrium data for FJSM-SnS (Figure 2a). And the corresponding parameters obtained from the fit are shown in Table S3. The maximum 3+ 3+ 3+ sorption capacities of Eu , Tb and Nd were 139.82 ± 3.42 mg/g, 147.05 ± 4.53 mg/g and 126.70 ± 3.90 mg/g, respectively, which were close to their theoretical ion-exchange capacity (around 150 mg/g). It is worth noting that the capac12 ities are much better than those of Al2O3/EG, clay miner22 20 21 als, zeolite and activated carbon with the capacities of 3+ Eu ranging from 3.28 to 46.5 mg/g. Detailed comparisons on REE adsorption capacity with different adsorbents are listed in Table S4.

2n-

Figure 1. (a) A 2D [Sn3S7]n anionic layer with its large windows of 24-membered Sn12S12 rings viewed parallel to the ab plane, SnS5 polyhedra are shaded in purple; (b) schematic of the REE insertion process by FJSM-SnS through exchange + + of its interlayer [Me2NH2] and [Me3NH] cations. The ion-exchange kinetics and capacity (q) (equation S1) were studied as key parameters to assess the potential of the sulfide material for application of large scale REE entrap39 ment. The kinetics experiments showed that during the initial 5 min, the concentrations dropped drastically from 3+ 3+ 5300 ppb to 23 ppb for Eu , 5730 ppb to 12 ppb for Tb , respectively (Figure S3a). The fast kinetics indicate the ionexchange reaction is driven by chemical sorption rather than 40 physical sorption according to equation (1). It is interesting that both kinetic data can be fitted by pseudo second-order kinetic model (Figure S3b, c). This further confirms that the ion-exchange process involves a chemical adsorption. 3+

+ 0.67Eu + (Me2NH2)1.33(Me3NH)0.67Sn3S7·1.25H2O + + 8.75H2O  (Eu)0.67Sn3S7·10H2O + 1.33Me2NH2 + 0.67Me3NH (1) Comparisons of the kinetics of lanthanide adsorption with 41 other sorbents, such as Fe3O4@HA MNPs, attapulgite-iron 40 12 oxide composites, Al2O3/EG hybrid adsorbent and D113-III 24 resin were listed in Table S2. Clearly, FJSM-SnS has a faster adsorption ability than the materials mentioned above,

Figure 2. (a) The capacity (mg ions removed/g of ionexchanger) plotted against the ion concentration at equilibrium; (b) the Kd of ions in competitive ion-exchange 3+ experiments (Competitive experiments containing Eu (6.01 3+ 3+ + ppm), Al (118.3 ppm), Fe (53.17 ppm) and Na (330.37 ppm) in the solution were carried out); (c) removal rates as a func3+ tion of the Ln ion by FJSM-SnS and of bed volumes in the ion-exchange chromatographic column experiments. The distribution coefficient (Kd) (equation S2) as a barometer for selectivity was also tested. Competitive experiments in 3+ 3+ 3+ the solution containing Eu (6.01 ppm), Al (118.3 ppm), Fe 2

ACS Paragon Plus Environment

Page 3 of 5

Journal of the American Chemical Society +

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(53.17 ppm) and Na (330.37 ppm) were carried out (Figure 2b). The pH of these solutions was ~5.0. In the coexistence with large excess of competitive trivalent metal cations, the 3+ 3 Kd of Eu could reach 5.4 × 10 mL/g which was 43.60 times + 2 3+ 3 of Na (1.2 × 10 mL/g), 4.00 times of Al (1.3 × 10 mL/g), 3+ 2 12.48 times of Fe (4.3 × 10 mL/g). Since the radius of triva3+ 3+ lent Ln ion (r = 0.09 ~ 0.115 nm) is nearly twice that of Al 3+ 44 (r = 0.05 nm) and Fe (r = 0.06 nm), it has lower charge-to3+ 3+ 3+ radius ratio than Al and Fe . Thus Ln is softer Lewis acid3+ 3+ ic ion than Al and Fe . Thus, the good selectivity of FJSM3+ SnS for Eu may originate from the soft Lewis basic character of the chalcogenide framework with a larger affinity for 3+ 3+ 3+ + Ln than for the smaller Al , Fe and Na . The pH resistance is another important evaluative parameter for a good ion-exchanger. As shown in Figure S4, the Kd rapidly increased with the increase of pH for all REE ions tested until pH~8.5. In the solutions with pH~1.9, the Kd 3 3+ maximum was observed with 1.1 × 10 mL/g for Nd due to + the high density H as competitive ions. When the pH increased to 2.8, the Kd maximum value was observed at 3.9 × 3 3+ + 10 mL/g for Nd owing to the decrease of H density. The Kd maximum increased at higher pH values to ~5.9 (Eu: 1.2 × 5 6 10 mL/g) and ~8.5 (Yb: 6.5 × 10 mL/g), respectively. In 4 general, a material with a Kd >10 mL/g is regarded as an excellent adsorbent. When pH reached to 11, precipitates of REE oxyhydroxides formed. Compared to other REE sorbents 3 18 including KIT-6-N-DGA-1 (Kd maximum ≈ 6.8 × 10 mL/g) 5 45 and Ac-Phos SAMMS (Kd maximum ≈ 1.82 × 10 mL/g), the FJSM-SnS has notably higher affinity to REE. Because ion-exchange columns are commonly used in practical wastewater treatment, we fabricated such columns and used for all REE elements except Pm (Figure 2c). It is noted that some nano materials, such as dispersed GO nanosheets, may not be suitable for column use because of stacking and congesting tendency creating high back pres23 sure during wastewater flow. The FJSM-SnS bulk polycrystalline material showed notable advantages when applied in a column. The recovery rates (R) (equation S3) are extremely high around 99.99% in low concentrations (0.59 ~ 1.14 ppm) during processing of 6024 bed volumes of solution (bed volume = 0.498 mL). The high recovery rates underscore the high potential of FJSM-SnS for application in the recovery of REE from large volumes of aqueous solutions. We chose the Eu-loaded sample (FJSM-SnS-Eu) as an example for investigating the mechanism of trapping REE ions. The FJSM-SnS-Eu was obtained by soaking in high concen3+ tration Eu solution. Scanning electron microscope (SEM) and X-ray photoelectron spectroscopy (XPS) were employed to examine the morphology and Eu species of FJSM-SnS-Eu. The SEM image indicated that the sheet-like crystal shape after ion-exchange was retained (Figure 3a). The corresponding element distribution maps of Sn, S and Eu were recorded using SEM (Figures 3b-3d). The green color area in Figure 3d shows the homogeneous distribution of Eu in the FJSM-SnSEu crystal. It is important to note that the REE in the corresponding exchanged products can be eluted by treatment with concentrated KCl solution and the FJSM-SnS material regenerated. 3+ For example, the Eu in Eu-exchanged FJSM-SnS (FJSM-SnSEu) can be recovered by soaking in 2 mol/L KCl solution. The energy dispersive spectroscopy (EDS) results obtained after

3+

soaking showed that there was no Eu remaining in the eluted solid products (FJSM-SnS-K) (Figure S5). SEM imaging of the solid FJSM-SnS-K indicated that the crystal maintains the sheet-like morphology (Figure 3e). The red color area in Figure 3h shows the homogeneous distribution of K in the FJSM-SnS-K crystal. The XPS data on the samples show characteristic peaks for Eu 3d (Figure 4a). The 3d5/2 peak is located at 1137.68 eV with a satellite at 1127.9 eV. The binding energy value is slightly 3+ 10 3+ higher than for Eu /TiO2 (1135.1 eV) and Eu /hydrous alu11 mina (1135.3 eV). Transmission electron microscopy (TEM) results also confirmed that FJSM-SnS-Eu maintained its crystallinity. The selected area electron diffraction pattern from a sample of FJSM-SnS-Eu exhibited strong diffraction rings (Figure S6a) consistent with the crystal structure and highresolution TEM images showed a 0.64 nm interlayer space (Figure S6b).

Figure 3. SEM images of FJSM-SnS-Eu (a) and FJSM-SnS-K (e). (b) to (d) and (f) to (h) show the corresponding EDS element distribution maps of Sn, S and Eu/K for FJSM-SnSEu and FJSM-SnS-K, respectively. FJSM-SnS-Eu samples came from the ion-exchanged experiment (V : m = 6000 3+ mL/g, contact time 12h, at room temperature, initial Eu concentration was 1935 ppm).

Figure 4. (a) The XPS spectrum of FJSM-SnS-Eu. The samples came from the ion-exchanged experiments (V : m = -1 o 3+ 90.25 mL g , contact time 24 h, at 65 C and initial Eu concentration of 21025 ppm). (b) Solid-state optical absorption 3+ 3+ spectra of the pristine FJSM-SnS, Eu and Tb -exchanged o products (V : m = 1000 mL/g, contact time 20 min, at 65 C, 3+ 3+ initial Eu and Tb concentrations of 1000 ppm). The fitting of the Langmuir equilibrium isotherm equation 3+ + (2) gives b (Eu , 3.92 L/mg) > b (Cs , 0.0038 L/mg) indicates 31 a significantly larger absorption energy for REE. In addition, the optical absorption edges of the Tb and Eu-exchanged products are red-shifted at 2.77 eV and 2.52 eV, respectively, compared to the 2.92 eV of the pristine compound. The re3+ 3+ sults are consistent with the deeper colors of Eu and Tb exchanged crystals (Figure 4b). The red shift of the absorption edge shows that there is a binding interaction between 3+ the Ln ion and S atoms of the framework. Similar red-shifts have been reported in other chalcogenide ion-exchange ma29, 30, 32 terials involving soft metals such as Hg, Pb, Ag etc We have demonstrated that the 2D crystalline thiostannate FJSM-SnS is an efficient ion-exchanger for REE recovery. 3

ACS Paragon Plus Environment

Journal of the American Chemical Society

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Importantly, the material shows rapid kinetics, high capacity, wide pH resistance and facile, cheap and efficient elution. It 3+ 3+ + also shows high selectivity against Al , Fe and Na and high recovery rate (> 99%) in low concentration solutions. The high capacity and excellent selectivity for REE originate partly from the strong affinity of its soft Lewis basic sulfide 3+ framework to high valence state of Ln , coupled with the flexible nature of the microporous framework and its tunable interlayer distance. An additional strong driving force for the selectivity is electrostatic coulombic attraction associated with the high charges cations with the anionic framework. Our study shows that chalcogenides can be highly effective in the field of REE recovery and recycling from complex aque3+ ous solutions. The high affinity for Ln , the low cost and environmentally friendly character of FJSM-SnS are highly attractive features. This work distinguishes the crystalline microporous sulfide ion-exchanger materials as new types of effective media for REE recovery and recycling.

Physical measurements, ion-exchange experiments details, EDS, TEM, PXRD, TGA and tables for comparisons on kinet3+ ics and capacities of Ln adsorbents. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author *[email protected] *[email protected]

Author Contributions §Xing-Hui Qi and Ke-Zhao Du contributed equally.

Notes The authors declare no competing financial interests.

ACKNOWLEDGMENT Financial support from the National Science Foundations of China (NSFC; Nos. 21521061 and 21373223), 973 programs (Nos. 2014CB845603 and 2012CB821702), Chunmiao project of Haixi institute of CAS (Chinese Academy of Sciences) (CMZX-2014-001) and National Science Foundation of United States (Grant DMR-1410169) are gratefully acknowledged.

REFERENCES

5. 6. 7. 8.

11. 12.

13. 14. 15. 16. 17. 18. 19.

21.

Supporting Information

4.

10.

20.

ASSOCIATED CONTENT

1. 2. 3.

9.

Leventis, N.; Gao, X. R., J. Am. Chem. Soc. 2002, 124, 1079-1088. Binnemans, K., Chem. Rev. 2009, 109, 4283-4374. Zhang, N.; Salzinger, S.; Deubel, F.; Jordan, R.; Rieger, B., J. Am. Chem. Soc. 2012, 134, 7333-7336. Alonso, E.; Sherman, A. M.; Wallington, T. J.; Everson, M. P.; Field, F. R.; Roth, R.; Kirchain, R. E., Environ. Sci. Technol. 2012, 46, 3406-3414. Zaimes, G. G.; Hubler, B. J.; Wang, S.; Khanna, V., Acs Sustainable Chem. Eng. 2015, 3, 237-244. Sprecher, B.; Kleijn, R.; Kramer, G. J., Environ. Sci. Technol. 2014, 48, 9506-9513. Binnemans, K.; Jones, P. T.; Blanpain, B.; Van Gerven, T.; Yang, Y. X.; Walton, A.; Buchert, M., J. Clean. Prod. 2013, 51, 1-22. Tan, Q.; Li, J.; Zeng, X., Crit. Rev. Env. Sci. Tec. 2015, 45, 749-776.

22. 23. 24. 25. 26. 27.

Page 4 of 5

Chi, R.; Hu, Y.; Zhu, G.; Xu, S.; Zhou, Z.; Xu, Z., Metall. Mater. Trans. B 2003, 34, 611-617. Tan, X. L.; Fan, Q. H.; Wang, X. K.; Grambow, B., Environ. Sci. Technol. 2009, 43, 3115-3121. Tan, X. L.; Wang, X. K.; Geckeis, H.; Rabung, T., Environ. Sci. Technol. 2008, 42, 6532-6537. Sun, Y. B.; Chen, C. L.; Tan, X. L.; Shao, D. D.; Li, J. X.; Zhao, G. X.; Yang, S. B.; Wang, Q.; Wang, X. K., Dalton Trans. 2012, 41, 13388-13394. Gannaz, B.; Antonio, M. R.; Chiarizia, R.; Hill, C.; Cote, G., Dalton Trans. 2006, 4553-4562. Yaftian, M. R.; Burgard, M.; Dieleman, C. B.; Matt, D., J. Membr. Sci. 1998, 144, 57-64. Guo, L.; Chen, J.; Shen, L.; Zhang, J.; Zhang, D.; Deng, Y., Acs Sustainable Chem. Eng. 2014, 2, 1968-1975. Vijayaraghavan, K.; Sathishkumar, M.; Balasubramanian, R., Ind. Eng. Chem. Res. 2010, 49, 4405-4411. Li, C.; Zhuang, Z.; Huang, F.; Wu, Z.; Hong, Y.; Lin, Z., ACS Appl. Mater. Interfaces 2013, 5, 9719-9725. Florek, J.; Chalifour, F.; Bilodeau, F.; Larivière, D.; Kleitz, F., Adv. Funct. Mater. 2014, 24, 2668-2676. Sui, N.; Huang, K.; Zheng, H.; Lin, J.; Wang, X.; Xiao, C.; Liu, H., Ind. Eng. Chem. Res. 2014, 53, 16033-16043. Shao, D. D.; Fan, Q. H.; Li, J. X.; Niu, Z. W.; Wu, W. S.; Chen, Y. X.; Wang, X. K., Micropor. Mesopor. Mat. 2009, 123, 1-9. Gad, H. M. H.; Awwad, N. S., Sep. Sci. Technol. 2007, 42, 36573680. Sun, Y.; Li, J.; Wang, X., Geochim. Cosmochim. Acta 2014, 140, 621-643. Li, C. R.; Huang, Y.; Lin, Z., J. Mater. Chem. A 2014, 2, 1497914985. Xiong, C.; Chen, X.; Yao, C., J. Rare Earth. 2011, 29, 979-985. Ding, N.; Kanatzidis, M. G., Nat. Chem. 2010, 2, 187-191. Manos, M. J.; Kanatzidis, M. G., Chem.- Eur. J. 2009, 15, 4779-4784. Sengupta, P.; Dudwadkar, N. L.; Vishwanadh, B.; Pulhani, V.; Rao, R.; Tripathi, S. C.; Dey, G. K., J. Hazard. Mater. 2014, 266, 94-101.

28. Manos, M. J.; Kanatzidis, M. G., Chem. Sci. 2016, 7, 4804-4824. 29. Sarma, D.; Malliakas, C. D.; Subrahmanyam, K. S.; Islam, S. M.; Kanatzidis, M. G., Chem. Sci. 2016, 7, 1121-1132. 30. Manos, M. J.; Iyer, R. G.; Quarez, E.; Liao, J. H.; Kanatzidis, M. G., Angew. Chem. Int. Ed. 2005, 44, 3552-3555. 31. Qi, X. H.; Du, K. Z.; Feng, M. L.; Li, J. R.; Du, C. F.; Zhang, B.; Huang, X. Y., J. Mater. Chem. A 2015, 3, 5665-5673. 32. Feng, M. L.; Sarma, D.; Qi, X. H.; Du, K. Z.; Huang, X. Y.; Kanatzidis, M. G., J. Am. Chem. Soc. 2016, 138 , 12578-12585. 33. Marking, G. A.; Kanatzidis, M. G., Chem. Mater. 1995, 7, 1915-1921. 34. Parise, J. B.; Ko, Y. H.; Rijssenbeek, J.; Nellis, D. M.; Tan, K. M.; Koch, S., Chem. Commun. 1994, 527-527. 35. Tan, K. M.; Ko, Y. G.; Parise, J. B., Acta Crystallogr. C 1995, 51, 398-401. 36. Jiang, T.; Lough, A.; Ozin, G. A., Adv. Mater. 1998, 10, 42-43. 37. Bowes, C. L.; Petrov, S.; Vovk, G.; Young, D.; Ozin, G. A.; Bedard, R. L., J. Mater. Chem. 1998, 8, 711-720. 38. Jiang, T.; Lough, A.; Ozin, G. A.; Bedard, R. L.; Broach, R., J. Mater. Chem. 1998, 8, 721-732. 39. Awual, M. R.; Kobayashi, T.; Miyazaki, Y.; Motokawa, R.; Shiwaku, H.; Suzuki, S.; Okamoto, Y.; Yaita, T., J. Hazard. Mater. 2013, 252, 313-320. 40. Lei, C.; Tian, X.; Ma, B., J. Radioanal. Nucl. Chem. 2013, 298, 1127-1135. 41. Yang, S. T.; Zong, P. F.; Ren, X. M.; Wang, Q.; Wang, X. K., ACS Appl. Mater. Interfaces 2012, 4, 6890-6899. 42. Tompkins, E. R.; Khym, J. X.; Cohn, W. E., J. Am. Chem. Soc. 1947, 69, 2769-2777. 43. Do, D. D. Adsorption Analysis: Equilibria and Kinetics, Imperial College Press, 1998. 44. Nightingale, E. R., J. Phys. Chem. 1959, 63, 1381-1387. 45. Fryxell, G. E.; Wu, H.; Lin, Y. H.; Shaw, W. J.; Birnbaum, J. C.; Linehan, J. C.; Nie, Z. M.; Kemner, K.; Kelly, S., J. Mater. Chem. 2004, 14, 3356-3363.

4

ACS Paragon Plus Environment

Page 5 of 5

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

TOC graphic

5

ACS Paragon Plus Environment