Acessibilidade / Reportar erro

New Attempts to Synthesize Layered Double Hydroxides Intercalated with SO42-/Cs+ Using Co-Precipitation and Exchange Reactions

Abstract

Layered double hydroxides (LDHs) with the compositions (Cs+/NH4+)0.111[M2+0.667Al0.333(OH)2.0 (SO4)0.222] (M2+ = Mn, Zn) and basal distance of ca. 11 Å were obtained by co-precipitating Mn/Al and Zn/Al sulfate salts with aqueous NH3, using excess of Cs2SO4. [Mn0.667Al3+0.333(OH)2]Cl0.333.nH2O and [M2+0.667Al3+0.333(OH)2](NO3)0.333.nH2O (M2+ = Mn, Zn) were also synthesized by co-precipitation, presenting respective basal distances of 8.92 and 7.92 Å. After applying exchange reactions with excess of Cs2SO4, materials with basal distances of ca. 11 Å were obtained, indicating the exchange of chloride and nitrate with sulfate, without incorporation of Cs+. When Na+0.111[M2+0.667Al0.333(OH)2.0(SO4)0.222](M2+ = Mn, Zn) obtained by co-precipitation and having basal distances of ca. 11 Å was exchanged with excess of Cs2SO4, the content of sulfate remained constant and Na+ was partially replaced with Cs+, but the amount was lower, indicating the probable composition (Na+/Cs+)y[M2+0.667Al0.333(OH)2-y(SO4)y/2(SO4)0.222-(y/2)](M2+ = Mn, Zn), where some of the hydroxide anions were replaced with grafted SO42-.

Keywords:
layered double hydroxide; intercalation; exchange reactions; sulfate; cesium


Introduction

The removal of radioactive nuclides from contaminated waters is an important topic of research, especially after the Fukushima Daiichi nuclear power plant accident in 2011, which released large amounts of radioactive nuclides into the environment, especially the dangerous radioactive nuclide137Cs. Several methods of cesium intercalation in layered materials have been reported in the literature, such in graphite,11 Cajipe, V. B.; Heiney, P. A.; Fischer, J. E.; Phys. Rev. B 1989, 39, 4374. 2H-NbS2,22 Paulus, W.; Katzke, H.; Schöllhorn, R.; J. Solid State Chem. 1992, 96, 162. C-60,33 Messaoudi, A.; Conard, J.; Setton, R.; Beguin, F.; Chem. Phys. Lett. 1993, 202, 506. 1T-TaSe2,44 Tonti, D.; Pettenkofer, C.; Jaegermann, W.; Papageorgopoulos, D. C.; Kamaratos, M.; Papageorgopoulos, C. A.; Ionics 1998, 4, 93. manganese thiophosphate,55 Silipigni, L.; Acacia, N.; Quattrone, T.; de Luca, G.; Scolaro, L. M.; Salvato, G.; J. Appl. Phys. 2007, 102, 113713. clay minerals,66 Park, C. W.; Kim, B. H.; Yang, H. M.; Seo, B. K.; Moon, J. K.; Lee, K. W.; Chemosphere 2017, 168, 1068. 1T-MoS2,77 Mao, Y. L.; Fang, Y. Q.; Pan, J.; Wang, D.; Bu, K. J.; Che, X. L.; Zhao, W.; Huang, F. Q.; J. Solid State Chem. 2019, 279, 120937. sulfate green rusts (GRSO4) with the formula NaFeII6FeIII3(SO4)2(OH)18·12H2O88 Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem. 2014, 53, 8887. clay minerals from the smectite group,66 Park, C. W.; Kim, B. H.; Yang, H. M.; Seo, B. K.; Moon, J. K.; Lee, K. W.; Chemosphere 2017, 168, 1068. and other clay minerals,99 Mukai, H.; Hirose, A.; Motai, S.; Kiruchi, R.; Tanoi, K.; Nakanishi, T. M.; Yaiti, T.; Yaita, T.; Kogure, T.; Sci. Rep. 2016, 21543.

10 Naamen, S.; Jâafar, N.; Rhaiem, H. B.; Amara, A. B. H.; Plançon, A.; Muller, F.; Clay Miner. 2016, 51, 29.
-1111 Park, C. W.; Kim, S. M.; Kim, I.; Yoon, I. H.; Hwang, J.; Kim, J. H.; Yang, H. M.; Seo, B. K.; J. Environ. Radioact. 2021, 233, 106592. but to the best of our knowledge, none of them have reported the intercalation of cations in layered double hydroxides (LDH), especially due to the fact that LDHs are typically anion exchangers.

Traditional LDH with the chemical composition [M2+1-xM3+x(OH)2](An-)x/n.nH2O are natural minerals and synthetic materials derived from the brucite-like structure (Mg(OH)2), in which M2+ octahedrally coordinated to six OH- anions share edges to form two-dimensional layers that are stacked in the basal direction. In the LDH structure, M2+ cations in the brucite-like structure are partially replaced by M3+ and the excess positive charges of the layers are compensated by the intercalation of normally hydrated An-anions.1212 Clay Surfaces - Fundamentals and Applications, vol. 1; Wypych, F.; Satyanarayana, K. G., eds.; Academic Press: Amsterdam, 2004.

13 Layered Double Hydroxides; Duan, X.; Evans, D. G., eds.; Springer Verlag: Berlin, 2006.

14 Wang, Q.; O’Hare, D.; Chem. Rev. 2012, 112, 4124.

15 Rives, V.; del Arco, M.; Martín, C.; Appl. Clay Sci. 2014, 88-89, 239.

16 Leroux, F.; Moujahid, E.; Taviot-Gueho, C.; Besse, J. P.; Solid State Sci. 2001, 3, 81.

17 Radha, A. V.; Kamath, P. V.; Shivakumara, C.; Solid State Sci. 2005, 7, 1180.

18 Williams, G. R.; Rees, N. H.; O’Hare, D.; Solid State Sci. 2009, 11, 1229.

19 Zhang, F. R.; Du, N.; Li, H. P.; Liu, J. Q.; Hou, W. G.; Solid State Sci. 2014, 32, 41.
-2020 Delmas, C.; Borthomieu, Y.; J. Solid State Chem. 1993, 104, 345.

Recently, it has also been reported that LDHs can be obtained with the chemical composition (D(H2O)6)[M2+6Al3(OH)18(SO4)2].6H2O (D+ = Li, Na or K and M2+ = Mn, Zn, Mg, Co, Ni, Cu). These compounds were intercalated with hydrated sulfate anions, alkali metal cations and ammonium.2121 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824.

22 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531.

23 Sotiles, A. R.; Grassi, M. T.; dos Santos, M. P.; Wypych, F.; J. Braz. Chem. Soc. 2021, 32, 170.
-2424 Sotiles, A. R.; Wypych, F.; Solid State Sci. 2020, 106, 106304. Although the synthetic phases involved the intercalation of Li+, Na+, K+ and NH4+, the minerals containing M2+/Al3+ in the molar ratio of 2:1 were reported only with the intercalation of Na+ (natroglaucocerinite-M2+ = Zn; shigaite-M2+ = Mn and motukoreaite-M2+ = Mg).2525 Rodgers, K. A.; Chisholm, J. E.; Davis, R. J.; Nelson, C. S.; Mineral. Mag. 1977, 41, 389.

26 Zamareno, I.; Plana, F.; Vazques, A.; Clague, D. A.; Am. Mineral. 1989, 74, 1054.

27 Witzke, T.; Pöllmann, H.; Vogel, A.; Z. Kristallogr. 1995, 9, S.252.
-2828 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91. Other phases have also been reported with the compositions Fe2+/Fe3+, like nikischerite and sodium sulfate green rusts.88 Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem. 2014, 53, 8887.,2929 Huminicki, D. M. C.; Hawthorne, F. C.; Can. Mineral. 2003, 41, 79.,3030 Christiansen, B. C.; Balic-Zunic, T.; Petit, P. O.; Frandsen, C.; Mørup, S.; Geckeis, H.; Katerinopoulou, A.; Stipp, S. L. S.; Geochim. Cosmochim. Acta 2009, 73, 3579. In fact, nikischerite and sulfate green rusts have the same composition [NaFe2+6Fe3+3(SO4)2(OH)18.(H2O)12] and also similar lattice parameters (nikischerite: a = b = 9.347 Å, c = 33.00 Å and sulfate green rust: a = b = 9.258 Å, c = 10.968 Å), but they occur in different polytypes, which differ only in the number of stacked layers in the unit cell.2929 Huminicki, D. M. C.; Hawthorne, F. C.; Can. Mineral. 2003, 41, 79.

30 Christiansen, B. C.; Balic-Zunic, T.; Petit, P. O.; Frandsen, C.; Mørup, S.; Geckeis, H.; Katerinopoulou, A.; Stipp, S. L. S.; Geochim. Cosmochim. Acta 2009, 73, 3579.
-3131 Mills, S. J.; Christy, A. G.; Genin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289.

Only sulfate green rusts have Cs+ and Rb+ intercalated together with sulfate anions,88 Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem. 2014, 53, 8887. but washing of the materials with water indicated that Na+ and K+ were not affected, while Rb+ and Cs+ were leached out of the structure, with the corresponding reduction of the amount of intercalated sulfate and the basal distance. The reaction can be probably formulated as described in equation 1, when the cesium sulfate green rusts are converted into regular LDH.

(1)(Cs+)Fe62+Fe33+(SO4)2(OH)18Fe62+Fe33+(SO4)1.5(OH)18+1/2Cs2SO4

Due to the possibility to extend the broad range of applications of LDH, especially related to the possibility to remove radioactive Cs+ from contaminated solutions and the lack of data about the intercalation of Cs+ in LDH/SO4, the objective of the present work is to describe attempts to synthesize and characterize Zn/Al and Mn/Al LDH containing SO42-/Cs+, obtained by co-precipitation at increasing pH and exchange reactions.

Experimental

The synthesis of Mn2Al:Cl, Mn2Al:NO3 and Zn2Al:NO3 phases was performed as recently reported.2121 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824. Briefly, LDH containing M2+:Al with molar ratios of 2:1 were synthesized by co-precipitation with increasing pH using an automatic glass titration reactor operating at 90 °C, under N2 flow, to avoid contamination with carbonate, where the pH was controlled by an internal pHmeter. A 100 mL solution of MnCl2/AlCl3, Mn(NO3)2/Al(NO3)3 or Zn(NO3)2/Al(NO3)3 were prepared with Milli-Q water and slowly titrated with a solution of NaOH 1 mol L-1, until attaining the desired pH. In the absence of CsOH to perform the co-precipitation, the M2+2Al:SO4/Cs (M2+ = Mn, Zn) samples were also synthesized by adding aqueous NH3 1 mol L-1 solutions to 200 mL solutions of M2+SO4 (M2+ = Mn, Zn), Al2(SO4)3 and Cs2SO4 prepared with Milli-Q water, with M2+:Al metal molar ratios of 2:1 using an excess of Cs2SO4 (Table 1).

Table 1
Amount of chemicals used and pH control during the LDH syntheses

All the chemicals were of analytical grade and used without any treatment: Al2(SO4)3.16H2O 98-102%, NaOH 99%, Al(NO3)3∙9H2O 98.50%, Mn(NO3)2∙4H2O 98%, Li2SO4∙H2O 99%, ZnSO4∙7H2O 99% and NaCl 99.85% were from Reatec (São Paulo, Brazil); AlCl3∙6H2O 99.50%, MnCl2∙4H2O 98-101%, ZnCl2 Vetec 97% and Zn(NO3)2∙6H2O 99% were from Vetec (Rio de Janeiro, Brazil); LiOH∙H2O Biotec 98% (São Paulo, Brazil); Na2SO4 Neon 99.9% (São Paulo, Brazil); NaNO3 F. Maia 99%; (São Paulo, Brazil); Aqueous NH3 Quimex 28-30% (Minas Gerais, Brazil); MnSO4∙H2O Alphatec 98-101% (Rio de Janeiro, Brazil); Cs2SO4 Merck 99.9% (Darmstadt, Germany).

After precipitation, the resulting slurries were ripened at 90 °C for 5 days in closed Erlenmeyer flasks, and separated by centrifugation at 4,000 rpm for 5 min (centrifugal force of 2,125 g), with the process being repeated after redispersing the slurry with an ultrasound bath for several seconds.2222 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531. The samples were dried at room temperature. The M2+2Al:SO4/Na, M2+2Al:NO3 (M2+ = Mn, Zn) and Mn2Al:Cl samples were synthesized as previously reported.2121 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824.,2222 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531.

In the exchange reactions, the phases Mn2Al:SO4/Na, Zn2Al:SO4/Na, Mn2Al:NO3, Zn2Al:NO3 and Mn2Al:Cl were dispersed in Milli-Q water containing excess of Cs2SO4 (three times the sulfate in relation of nitrate and chloride and three times in relation to sodium), and the mixtures were gently stirred for 7 days under N2 flow at room temperature. The same procedure was used in the centrifugation washing and drying process of the other samples. In all exchanged samples, the pH was kept almost constant and close to neutral (in the range of 6.7 to 7.7).

The synthesized compounds were characterized by X-ray diffraction (XRD) using a Shimadzu XRD-6000 diffractometer (Kyoto, Japan). After aging and the last centrifugation step, drops of the slurry were deposited on glass sample holders and dried at room temperature. The analyses were performed using Cu Kα = 1.5418 Å radiation, tension of 40 KV and current of 30 mA, with a dwell time of 2° min-1.

The samples were also characterized by Fourier transform infrared (FTIR) spectroscopy using a Bruker Vertex 70 spectrophotometer (Karlsruhe, Germany). KBr pellets containing around 1% (m/m) of LDH were gently mixed and pressed at 10 tons and the spectra were collected in transmission mode by accumulating 32 scans in the region of 400-4000 cm-1, using resolution of 2 cm-1.

Scanning electron microscopy (SEM) images and energy dispersive X-ray spectroscopy (EDS) data of some precursors were obtained with a Tescan Vega3LMU microscope (Brno-Kohoutovice, Czech Republic) with an AZ Tech software. The sample dispersions in water were dripped on copper tapes and after EDS measurements, the samples were sputtered with a thin gold layer to obtain the SEM images.

The quantitative analyses of the metals and sulfur (relative to SO42-) used to formulate the samples’ chemical composition were performed with a Thermo Scientific model iCAP 6500 inductively coupled plasma optical emission spectrometry (ICP-OES) (Waltham, United States) after dissolving the samples in 1.0% v/v of HNO3 in Milli-Q water. The data were collected in triplicate, treated with the Thermo Scientific iTeVa software version 1.2.0.30 and average values were used to obtain the LDH compositions.

Results and Discussion

In general, during exchange reactions, the layer lattice parameters remain almost constant while the basal parameter is normally dependent on the size of the intercalated cations and/or anions. During these reactions, it is also common for interpolytype transitions to occur due to the re-ordering of the layer stacking sequence. However, this seems not to be the case of exchange reactions of (D1+)[M2+6Al3(OH)18(SO4)2] to (D2+)[M2+6Al3(OH)18(SO4)2] (D1+ and D2+ = alkali metal cations), since the basal distance is obtained by the combination of different factors: the hydrated sulfate size in the form of a double layer; the size of the hydrated alkali metal cations with variable numbers of water molecules in the first hydration shell; and also to the interactions of sulfate with the alkali metal, water molecules and both with the layers having different compositions.

The samples Mn2Al:SO4/Cs (Figure 1Aa) and Zn2Al:SO4/Cs (Figure 1Ab) obtained by co-precipitation presented, respectively, basal distances of 11.36 and 10.91 Å. Zn2Al:SO4/Cs (Figure 1Ab) also was slightly contaminated with compounds having a basal distance of 8.9 Å, exactly the same impurity observed when Zn2Al:SO4/NH4 was prepared,2424 Sotiles, A. R.; Wypych, F.; Solid State Sci. 2020, 106, 106304. attributed probably to the intercalation of dehydrated sulfate.3232 Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem. 1995, 34, 883.,3333 Radha, S.; Jayanthi, K.; Breu, J.; Kamath, P. V.; Clays Clay Miner. 2014, 62, 53.

Figure 1
XRD patterns (A) and FTIR spectra (B) of Mn2Al:SO4/Cs (a) and Zn2Al:SO4/Cs (b) obtained by co-precipitation with increasing pH.

The values of the prevalent compounds are slightly bigger than those obtained for the Mn2Al:SO4 phases intercalated with sodium (11.02-11.03 Å) and potassium (11.27-11.28 Å) and slightly smaller than those observed for Zn2Al:SO4 phases intercalated with sodium (11.22-11.14 Å) and potassium (11.40 Å). Both values are consistent with the intercalation of sulfate in double layer arrangement and a single layer of hydrated Cs+ and NH4+, as observed for other alkali metal cations.2424 Sotiles, A. R.; Wypych, F.; Solid State Sci. 2020, 106, 106304.

An expansion of the XRD pattern of Mn2Al:SO4/Cs (Figure 1Aa, insert) presents the (100; d = 4.80 Å) and (101; d = 4.42 Å) diffraction peaks, which correspond to an a = a’ 3 superlattice, common for LDH with 2M2+:M3+ molar ratios intercalated with different anions and also observed in LDH intercalated with sulfate and alkali metal cations,2121 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824.,2222 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531. attributed to the ordering of metal cations in the layers. The cell parameters were calculated and found to be a = b = 5.54 Å and c = 11.36 Å with average distance between the metals of a’ = 3.20 Å, very close to a’ = 3.171 Å observed in the sodium shigaite structure.2828 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91. The same was not observed for Zn2Al:SO4/Cs due to the overlapping of the second peak of the 8.9 Å phase, in the same region.

Although the a = a’ 3 × a3’ superstructure is also observed in synthetic LDHs intercalated with sulfate and alkali metal cations,2121 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824.,2222 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531. natural LDHs intercalated with sulfate and alkali metal cations present the superstructure a = 3a’ × 3a’, which correspond to the ordering of the metal cations and/or sulfate anions between the layers.2828 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91.

29 Huminicki, D. M. C.; Hawthorne, F. C.; Can. Mineral. 2003, 41, 79.

30 Christiansen, B. C.; Balic-Zunic, T.; Petit, P. O.; Frandsen, C.; Mørup, S.; Geckeis, H.; Katerinopoulou, A.; Stipp, S. L. S.; Geochim. Cosmochim. Acta 2009, 73, 3579.

31 Mills, S. J.; Christy, A. G.; Genin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289.
-3232 Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem. 1995, 34, 883. These superstructure differences are probably associated with the sizes of the crystals and corresponding long and short-range order/disorder in synthetic and natural LDH. Since the basal distances are similar, the hydration shell of the intercalated alkali metal cations and ammonium should be different from those observed in solution, being respectively tetrahedral and octahedral for Li+, octahedral for Na+, square antiprismatic for K+ and Rb+, probably 12-coordinated for Cs+3434 Persson, I.; Pure Appl. Chem. 2010, 82, 1901. and not well established for NH4+, for which the number of water molecules in the hydration shell ranges from 4 to 6.3535 Brugé, F.; Bernasconi, M.; Parrinello, M.; J. Am. Chem. Soc. 1999, 121, 10883.,3636 Guo, J.; Zhou, L.; Zen, A.; Michaelides, A.; Wu, X.; Wang, E.; Xu, L.; Chen, J.; Phys. Rev. Lett. 2020, 125, 106001.

The average thicknesses of the particles along the basal axis and the respective number of stacked layers were calculated by the Scherrer equation using the first basal peak. The data correspond to 13.5 and 41 nm, respectively, for Mn2Al:SO4/Cs and Zn2Al:SO4/Cs, which corresponds to 12 and 36 stacked layers, respectively.

The presence of ammonium was inferred since the precipitation was conducted with aqueous NH3 instead of CsOH, so no other cations were available in the alkaline solution to explain the higher content of sulfate anions. Although hydrated SO42- and NH4+ were intercalated between the layers, the basal distance was only slightly affected in comparison to the other sulfate/alkali metal cations.

The FTIR spectra of Mn2Al:SO4/Na (Figure 1Ba) and Zn2Al:SO4/Na (Figure 1Bb) indicated the presence of typical stretching vibration bands of hydroxyl groups and water molecules in the region of 3430 cm-1, sulfate bands and M-O and M-OH bands below 600 cm-1 and a band in the region of 770-790 cm-1. The absorption band in the region of 1620 cm-1 is attributed to the bending vibration of water molecules.2121 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824.

22 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531.

23 Sotiles, A. R.; Grassi, M. T.; dos Santos, M. P.; Wypych, F.; J. Braz. Chem. Soc. 2021, 32, 170.
-2424 Sotiles, A. R.; Wypych, F.; Solid State Sci. 2020, 106, 106304.,3737 Khaldi, M.; de Roy, A.; Chaouch, M.; Besse, J. P.; J. Solid State Chem. 1997, 130, 66.

Due to different compositions and metals involved in the structure, the FTIR bands attributed to M-O and M-OH bonds were observed at 419, 534 and 767 cm-1 for Mn2Al:SO4/Cs (Figure 1Ba), while only two bands, at 427 and 617 cm-1, were observed for Zn2Al:SO4/Cs (Figure 1Bb). This absence can be attributed to the mixture of phases and low crystallinity of Zn2Al:SO4/Cs in comparison with Mn2Al:SO4/Cs. The sulfate bands in distorted tetrahedral symmetry also changed positions, being observed as a broad band at 1116 and 617 cm-1 for Zn2Al:SO4/Cs (Figure 1Bb) and at 1145, 1105, 981, 617 and 606 cm-1 for Mn2Al:SO4/Cs (Figure 1Ba). The presence of typical N-H bending vibration bands with low-symmetry ammonium was detected by FTIR in both samples at 1401 and 1443 cm-1, along with a low-intensity O-H overlapping band in the region of 3000 3700 cm1.2424 Sotiles, A. R.; Wypych, F.; Solid State Sci. 2020, 106, 106304.,3535 Brugé, F.; Bernasconi, M.; Parrinello, M.; J. Am. Chem. Soc. 1999, 121, 10883.,3838 Pironon, J.; Pelletier, M.; de Donato, P.; Mosser-Ruck, R.; Clay Miner. 2003, 38, 201. The band splitting in the region of 1400 cm-1 is probably related to hydrogen bonding occurring between NH4+ ions and H2O molecules and interactions with the negatively charged layers [M2+6Al3(OH)18(SO4)2]-. In solid NH4Cl, the band in the region of 1400 cm-1 is attributed to symmetric deformation mode ν4 of tetrahedral ammonium symmetry,3939 Aranda, P.; Ruiz-Hitzky, E.; Appl. Clay Sci. 1999, 15, 119.,4040 Nakamoto, K.; Infrared and Raman Spectra of Inorganic and Coordination Compounds, Part A: Theory and Applications in Inorganic Chemistry, 5th ed.; Wiley: New York, 1997. and this band was also observed at 1450 cm-1 when NH4+ was in solution.

The composition of the synthesized sample (Table 2) indicated that the observed values for the M2+ and M3+ metals in Zn2Al:NO3, Mn2Al:NO3 and Mn2Al:Cl were almost identical to the expected ideal formulas [M2+0.667Al3+0.333(OH)2](A-)0.333.nH2O, according to the proportions used during the synthesis procedures (Table 1), with the exception of Zn2Al:SO4/Cs, which indicated a slightly higher content of Al3+, due to the presence of impurities (Figure 1Ab) (0.372 instead of 0.333).

Table 2
Composition of the samples obtained by ICP-OES analyses

The same observation occurred after the exchange reaction of Zn2Al:SO4/Na with Cs2SO4 (0.367 instead of 0.333). The chemical composition of the phases Zn2Al:SO4/Cs and Mn2Al:SO4/Cs obtained by direct synthesis also indicated the correct proportions of M2+ and M3+ and sulfur from sulfate for the ideal composition Cs+0.111[M2+0.667Al0.333(OH)2](SO4)0.222, but the content of cesium was lower than predicted.

Although not quantified, this difference can be attributed to NH4+, as indicated by FTIR (Figure 1Ba,b), since the precipitation was conducted in the presence of aqueous NH3 and NH4+/SO42-. Intercalated LDH have already been reported in the literature, having similar basal distances.2424 Sotiles, A. R.; Wypych, F.; Solid State Sci. 2020, 106, 106304. These compositions are proposed to be (Cs0.111-xNH4+x)[M2+0.667Al0.333(OH)2](SO4)0.222.

The reduced content of intercalated Cs+ and Rb+ have already been reported for sulfate green rusts (GRSO4), having chemical composition close to D+FeII6FeIII3(SO4)2(OH)18·12H2O (D+ = Na, K, Rb, Cs),88 Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem. 2014, 53, 8887. while rinsed samples of intercalated cesium were observed to have smaller basal distance (10.85 Å) than sodium (10.96 Å) and potassium (11.20 Å), in spite of the bigger ionic radius of the first. A single washing step was sufficient to remove all intercalated Cs+ from GRSO4, and only a double layer of sulfate was retained.88 Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem. 2014, 53, 8887. This behavior was not observed in our samples, in which cesium was still observed despite being washed twice, indicating a stronger interaction with the LDH layers.

In the attempt to exchange Na+ from Mn2Al:SO4/Na and Zn2Al:SO4/Na with Cs+, again the content of Cs+ plus Na+ did not have the expected value, and the content of sulfate was again higher, suggesting the maintenance of the original composition D+0.111[M2+0.667Al0.333(OH)2](SO4)0.222. HSO4- is not expected at neutral and slightly alkaline pH,4141 Iftekhar, S.; Küçük, M. E.; Srivastava, V.; Repo, E.; Sillanpa, M.; Chemosphere 2018, 209, 470. since this would increase the amount of sulfur in the samples and the absence of other cations in solution, an alternative is to consider that part of the alkali metal cations was replaced with hydrated protons. However, this hypothesis is unlikely since the pH of the exchange solution was neutral or slightly alkaline.

The second possibility is the adsorption of sulfate on the protonated hydroxide anions of the particles’ surface through -OH2+(SO42-)0.5.4141 Iftekhar, S.; Küçük, M. E.; Srivastava, V.; Repo, E.; Sillanpa, M.; Chemosphere 2018, 209, 470. However, it is highly unlikely that the exact amount proposed in the ideal formula D+0.111[M2+0.667Al0.333(OH)2](SO4)0.222 would be obtained.

A third and more feasible hypothesis is the grafting of sulfate, as observed in layered hydroxide salts like sodium gordaite (NaZn4(OH)6(SO4)Cl.6H2O),4242 Adiwidjaja, G.; Friese, K.; Klaska, K. H.; Schlüter, J.; Z. Kristallogr. 1997, 212, 704.,4343 Maruyama, S. A.; Krause, F.; Tavares, S. R.; Leitão, A. A.; Wypych, F.; Appl. Clay Sci. 2017, 146, 100. where part of the structural hydroxide anions is replaced and grafted with chloride and sulfate, generating negatively charged layers [Zn4(OH)6(SO4)Cl]-, whose charges are compensated by the intercalation of Na+ cations. Although bereft of alkali metal cations, layered double hydroxides with the proposed grafting mechanism have similar structures to spangolite (Cu6Al(OH)12(SO4)Cl·3H2O)4444 Hawthorne, F. C.; Kimata, M.; Eby, R. K.; Am. Mineral. 1993, 78, 649. and jamborite (Ni2+1-xCo3+x(OH)2-x(SO4)x·nH2O),4545 Bindi, L.; Christy, A. G.; Mills, S. J.; Ciriotti, M. E.; Bittarello, E.; Can. Mineral. 2015, 53, 791. where the SO42- tetrahedra are grafted to the layer, partially replacing surface OH-. Using the structures of jamborite (where sulfate is grafted to the layers) and zincowoodwardite ([Zn1-xAlx(OH)2][(SO4)x/2.nH2O)4646 Witzke, T.; Raade, G.; Neues Jahrb. Mineral., Monatsh. 2000, 455. (where sulfate is only intercalated) as examples, and with x = 0.333, the formulas would be Ni2+0.667Co3+0.333(OH)1.667(SO4)0.333·nH2O and Zn0.667Al3+0.333(OH)2(SO4)0.167·nH2O.

In jamborite, the content of sulfate would double in relation to zincowoodwardite while the M2+:M3+ ratio would be the same (2:1). Grafting of sulfate and carbonate has even been observed in single hydroxides, as is in the case of paraotwayite (Ni(OH)2-x(SO4,CO3)x/2)4747 Nickel, E. H.; Graham, J.; Can. Mineral. 1987, 25, 409. and was also in LDH intercalated with dehydrated sulfate.3232 Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem. 1995, 34, 883.,3333 Radha, S.; Jayanthi, K.; Breu, J.; Kamath, P. V.; Clays Clay Miner. 2014, 62, 53. Hence, it would not be surprising to find this pattern in LDH intercalated with hydrated sulfate.

Using x = 0.333 and y = 0.50 in the formula D+y[M2+1-xAlx(OH)2-y(SO4)y/2(SO4)0.222-(y/2)] as an example (reduction of the content of alkali metal cations of 50%) in relation to the shigaite-like formula Cs[M2+6Al3(OH)18(SO4)2.0] and keeping the metal molar ratio of 2:1, LDHs with the composition (Cs)0.50[M2+6Al3(OH)17.5(SO4)0.25*(SO4)1.75] would be obtained, which in a reduced way would be formulated as (Cs)0.056[M2+0.667Al0.333(OH)1.944(SO4)0.028*(SO4)0.194] (total SO42- = 0.222 and *= grafted SO42-) or even (Cs)0.056[M2+0.667Al0.333(OH)1.944(SO4)0.222].

These propositions where some of hydroxide anions from the layers are replaced and grafted with SO42- are in relatively good agreement with the formulas suggested by the ICP-OES analyses (Table 2).

The XRD patterns and FTIR spectra before and after exchange reactions are shown in Figure 2. When the samples Mn2Al:SO4/Na (Figure 2Aa) and Zn2Al:SO4/Na (Figure 2Ac) were exchanged with Cs2SO4, the basal distances remained almost constant in both cases (11.04 to 11.15 Å for Zn2Al:SO4/Na (Figure 2Aa,c) and 11.15 to 11.14 Å for Mn2Al:SO4/Na (Figure 2Ab,d). This indicates that the exchange reaction was unsuccessful, since the basal distances observed for the samples of Mn2Al:SO4/Cs (Figure 1Aa) and Zn2Al:SO4/Cs (Figure 1Ab) obtained by co-precipitation were also in the same range (11.36 Å for Mn2Al:SO4/Cs and 10.91 Å for Zn2Al:SO4/Cs).

Figure 2
XRD patterns (A) and FTIR spectra (B, C) of Mn2Al:SO4/Na (a) and Zn2Al:SO4/Na (c) and after respective exchange reactions with Cs2SO4: Mn2Al:SO4/Na-Cs (b) and Zn2Al:SO4/Na-Cs (d).

When the FTIR spectra were evaluated, the same band positions were observed in the samples before and after exchange reactions, with Cs2SO4, at 1192, 1145, 1108, 953, 773, 618, 604, 533 and 421 cm-1 for Mn2Al:SO4/Cs (Figure 2Ca,b) and at 1192, 1156, 1110, 962, 792, 619, 550, 427 cm-1 for Zn2Al:SO4/Cs (Figure 2Cc,d). This observation is different than that of washed samples of green rusts intercalated with cesium and rubidium, when the removal of the cations was attributed to shifting of the v3 bands of sulfate at about 15 cm-1 to lower wavenumbers.88 Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem. 2014, 53, 8887.

The presence of the split band in the region of 1100 cm-1 in all samples suggests that the sulfate environment is highly distorted,3737 Khaldi, M.; de Roy, A.; Chaouch, M.; Besse, J. P.; J. Solid State Chem. 1997, 130, 66. since in the absence of sulfate/alkali metal/layer interaction, sulfate’s undistorted tetrahedral symmetry would be characterized by a single band in the region of 1100 cm-1.4848 Zegeye, A.; Ona-Nguema, G.; Carteret, C.; Huguet, L.; Abdelmoula, M.; Jorand, F.; Geomicrobiol. J. 2005, 22, 389. Grafting of SO42- would only contribute to very low intensity bands in the same region,3232 Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem. 1995, 34, 883. which were overlapped by the other sulfate bands. More details of the band attributions can be found in the literature.4848 Zegeye, A.; Ona-Nguema, G.; Carteret, C.; Huguet, L.; Abdelmoula, M.; Jorand, F.; Geomicrobiol. J. 2005, 22, 389. Figure 3A shows the results of our attempt to produce the phases intercalated with sulfate and cesium, using the precursor Mn2Al:NO3 (Figure 3Aa), Zn2Al:NO3 (Figure 3Ac) and Mn2Al:Cl (Figure 3Ae).

Figure 3
XRD patterns (A) and FTIR spectra (B) of Mn2Al:NO3 (a), Zn2Al:NO3 (c) and Mn2Al:Cl (e) and after respective exchange with Cs2SO4:Mn2Al:NO3/Cs (b), Zn2Al:NO3/Cs (d) and Mn2Al:Cl/Cs (f).

As expected, in all cases after the attempts to exchange the pristine anions with SO42-/Cs+, we found that the basal distances increased from 8.92 Å in Mn2Al:NO3 (Figure 3Aa) and Zn2Al:NO3 (Figure 3Ac) to around 11 Å, and from 7.82 Å in Mn2Al:Cl (Figure 3Ae)3232 Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem. 1995, 34, 883.,4949 Marappa, S.; Radha, S.; Kamath, P. V.; Eur. J. Inorg. Chem. 2013, 12, 2122. also to around 11 Å (Figure 3Ab,d,f), indicating that the former’s intercalated anions were replaced with SO422 Paulus, W.; Katzke, H.; Schöllhorn, R.; J. Solid State Chem. 1992, 96, 162.

3 Messaoudi, A.; Conard, J.; Setton, R.; Beguin, F.; Chem. Phys. Lett. 1993, 202, 506.

4 Tonti, D.; Pettenkofer, C.; Jaegermann, W.; Papageorgopoulos, D. C.; Kamaratos, M.; Papageorgopoulos, C. A.; Ionics 1998, 4, 93.

5 Silipigni, L.; Acacia, N.; Quattrone, T.; de Luca, G.; Scolaro, L. M.; Salvato, G.; J. Appl. Phys. 2007, 102, 113713.

6 Park, C. W.; Kim, B. H.; Yang, H. M.; Seo, B. K.; Moon, J. K.; Lee, K. W.; Chemosphere 2017, 168, 1068.

7 Mao, Y. L.; Fang, Y. Q.; Pan, J.; Wang, D.; Bu, K. J.; Che, X. L.; Zhao, W.; Huang, F. Q.; J. Solid State Chem. 2019, 279, 120937.

8 Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem. 2014, 53, 8887.

9 Mukai, H.; Hirose, A.; Motai, S.; Kiruchi, R.; Tanoi, K.; Nakanishi, T. M.; Yaiti, T.; Yaita, T.; Kogure, T.; Sci. Rep. 2016, 21543.

10 Naamen, S.; Jâafar, N.; Rhaiem, H. B.; Amara, A. B. H.; Plançon, A.; Muller, F.; Clay Miner. 2016, 51, 29.

11 Park, C. W.; Kim, S. M.; Kim, I.; Yoon, I. H.; Hwang, J.; Kim, J. H.; Yang, H. M.; Seo, B. K.; J. Environ. Radioact. 2021, 233, 106592.

12 Clay Surfaces - Fundamentals and Applications, vol. 1; Wypych, F.; Satyanarayana, K. G., eds.; Academic Press: Amsterdam, 2004.

13 Layered Double Hydroxides; Duan, X.; Evans, D. G., eds.; Springer Verlag: Berlin, 2006.

14 Wang, Q.; O’Hare, D.; Chem. Rev. 2012, 112, 4124.

15 Rives, V.; del Arco, M.; Martín, C.; Appl. Clay Sci. 2014, 88-89, 239.

16 Leroux, F.; Moujahid, E.; Taviot-Gueho, C.; Besse, J. P.; Solid State Sci. 2001, 3, 81.

17 Radha, A. V.; Kamath, P. V.; Shivakumara, C.; Solid State Sci. 2005, 7, 1180.

18 Williams, G. R.; Rees, N. H.; O’Hare, D.; Solid State Sci. 2009, 11, 1229.

19 Zhang, F. R.; Du, N.; Li, H. P.; Liu, J. Q.; Hou, W. G.; Solid State Sci. 2014, 32, 41.

20 Delmas, C.; Borthomieu, Y.; J. Solid State Chem. 1993, 104, 345.
- 2121 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824.,3737 Khaldi, M.; de Roy, A.; Chaouch, M.; Besse, J. P.; J. Solid State Chem. 1997, 130, 66. but not with SO42-/Cs+, as already indicated by the ICP-OES analyses (Table 2).

This is evidence that Cs+/SO42- are not stable phases that can be obtained by exchange reactions, as already reported for sulfate green rusts,88 Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem. 2014, 53, 8887. but this is not true in the case of exchanging the same LDH with Li+, Na+ and K+, where the exchange reactions were successful.2121 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824. More studies are still necessary to explain this instability also for LDHs intercalated with Rb+/SO42-.

It has also been reported in the literature that sulfate contains several degrees of hydration and those with 11 Å correspond to double hydration. Lower basal distances and interpolytypic transitions can be obtained by heating or submitting the samples to dry air or vacuum.3333 Radha, S.; Jayanthi, K.; Breu, J.; Kamath, P. V.; Clays Clay Miner. 2014, 62, 53.,5050 Bookin, A. S.; Drits, V. A.; Clays Clay Miner. 1993, 41, 551.

51 Pachayappan, L.; Kamath, P. V.; Clays Clay Miner. 2019, 67, 154.
-5252 Radha, S.; Antonyraj, C. A.; Kamath, P. V.; Kannan, S.; Z. Anorg. Allg. Chem. 2010, 636, 2658. However, our samples were dried at room temperature and no changes of the basal distances were observed (they remained close to 11 Å). The obtainment of exchanged compounds having the basal distance of 10.94 Å for Mn2Al:NO3/Cs (Figure 3Ab), 11.00 Å for Zn2Al:NO3/Cs (Figure 3Ad) and 10.90 Å for Mn2Al:Cl/Cs (Figure 3Af) is an indication that the compounds are similar to those obtained with the direct synthesis (11.36 Å for Mn2Al:SO4/Cs and 10.91 Å for the Zn2Al:SO4/Cs). Nevertheless, as indicated by the ICP-OES analyses (Table 2), although traces of cesium were found, the content of sulfate was lower, indicating that the formulas are attributed to regular LDHs ([M2+0.667Al3+0.333(OH)2](SO4)0.167).3737 Khaldi, M.; de Roy, A.; Chaouch, M.; Besse, J. P.; J. Solid State Chem. 1997, 130, 66.,5252 Radha, S.; Antonyraj, C. A.; Kamath, P. V.; Kannan, S.; Z. Anorg. Allg. Chem. 2010, 636, 2658.

The confirmation of the exchange reactions NO3-/SO42- can also be clearly seen by the replacement of the sharp bands at 1384 cm-1 (Figure 3Ba,c), attributed to nitrate, and appearance of bands in the region of 1100 cm-1, attributed to sulfate (Figure 3Bb,d). This band split also indicated distorted sulfate symmetry. After the exchange reaction of Mn2Al:NO3 with Cs2SO4, (Figure 3Ba,b), extra bands were also observed at 968 and 1026 cm-1, which could be attributed to O-H bending modes, similar to those observed in gibbsite.5353 Balan, E.; Lazzeri, M.; Morin, G.; Mauri, F.; Am. Mineral. 2006, 91, 115. The same bands (although that at 968 cm-1 can be attributed to sulfate) were also observed for Mn2Al:NO3 and Mn2Al:Cl/Cs (indicated by *). The origin of these bands was not clearly understood. A small contamination of carbonate (band at 1362 cm-1) was also observed in the sample Zn2Al:NO3/Cs (Figure 3Bd),2828 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91. already present as a shoulder in the sample Zn2Al:NO3before the exchange reactions (Figure 3Bc). The presence of carbonate can explain the slightly lower content of sulfate (0.146 instead of 0.167) (Table 2).

After the exchange of chloride (Figure 3Be), the expected band appeared in the region of 1100 cm-1, attributed to sulfate (Figure 3Bf). All the other bands remained constant, indicating the maintenance of the LDH lattice and suggesting exchange reactions of chloride with sulfate but without the incorporation of Cs+ (Table 2). SEM images of some key synthesized samples (Figure 4) indicated the expected morphology of LDH, with platelet-like particles, in which the diameter varied according to the compositions, being smaller than 1 mm in Mn2Al:SO4/Na (Figure 4a) and Zn2Al:SO4/Na (Figure 4b).

Figure 4
SEM images of Mn2Al:SO4/Na (a), Zn2Al:SO4/Na (b), Mn2Al:NO3 (c), Mn2Al:Cl (d) and the corresponding EDS spectra (e-h) (*Cu from the sample holder).

In the case of Mn2Al:NO3 (Figure 4c) and Mn2Al:Cl (Figure 4d), bigger particles were observed, reaching several micrometers with some powdered particles at the top of the crystals. These powdered particles are not attributed to crystalline impurities, since the XRD patterns indicated only basal peaks from the LDH, and chemical analysis also indicated the predicted composition (Table 2). The EDS spectra (Figures 4e-4h) also indicated qualitatively the presence of the expected elements according to the chemical compounds used during the synthesis.

As copper tapes were used to hold the samples, a small contamination with this element was observed in some of the spectra (indicated with asterisks).

All the compounds intercalated with hydrated sulfate in the presence of alkali metal cations, hydrated sulfate or even with grafted hydrated sulfate presented basal distances close to 11 Å (Figure 5), making them difficult to distinguish through XRD, but quantitative analysis by ICP-OES helped us to give some information about these complex systems.

Figure 5
Schematic representation of the 2M2+:M3+LDH structures intercalated with hydrated Na+/SO42-, such as in shigaite (a),2828 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91. hydrated SO42- as in [Zn2Cr(OH)6](SO4)0.5·4H2O (b)5454 Radha, S.; Kamath, P. V.; Inorg. Chem. 2013, 52, 4834. and intercalated/grafted with SO42-, as proposed in the present work (c). Water molecules were removed to facilitate visualization.

Conclusions

LDH intercalated with Cs+ and NH4+ together with SO42-obtained by co-precipitation of sulfate salts with aqueous NH3 in the presence of excess Cs2SO4, presented the composition (Cs/NH4)0.111[M2+0.667Al0.333(OH)2.0(SO4)0.222](M2+ = Mn or Zn) (equation 2), attested by ICP-OES analyses, basal distances close to 11 Å in the XRD patterns and typical bands of SO42- and NH4+ in the FTIR spectra.

(2)0.667M2+SO4+0.167Al2(SO4)3+Cs2SO4( excess )+NH4OH(Cs+/NH4+)0.111[M2+0.667Al0.333(OH)2.0(SO4)0.222](M2+=Mn,Zn)

LDH synthesized by co-precipitation with increasing pH of sulfate salts with NaOH and excess of Na2SO4 presented the composition Na0.111[M2+0.667Al0.333(OH)2.0(SO4)0.222] (M2+ = Mn, Zn) (equation 3), basal distances close to 11 Å and typical bands of SO42-.

(3)0.667M2+SO4+0.167Al2(SO4)3+Na2SO4( excess )+NaOHNa0.111+[M2+0.667Al0.333(OH)2.0(SO4)0.222](M2+=Mn,Zn)

After putting Na0.111[M2+0.667Al0.333(OH)2.0(SO4)0.222] in contact with excess Cs2SO4 in an attempt to exchange Na+ with Cs+, the LDH containing SO42-/Cs+ obtained were compared to those obtained by direct co-precipitation with increasing pH. In these samples, the basal distance was also close to 11 Å, but Na+ were partially exchanged with Cs+, while the amount of sulfate was constant, as attested by ICP-OES analyses and typical bands in the FTIR spectra, suggesting the formulas D+y[M2+0.667Al0.333(OH)2-y (SO4)y/2(SO4)0.222-(y/2)] (equation 4) (D+= Na+ or Cs+), where (SO4)y/2 represents grafted SO42-, partially replacing the OH- in the layers: ((OH)2-y).

(4)D0.111+[M0.6672+Al0.333(OH)2.0(SO4)0.222]+Cs2SO4( excess )Dy+[M0.6672+Al0.333(OH)2-y(SO4)y/2(SO4)0.222-(y/2)](M2+=Mn,Zn)

Co-precipitating nitrate or chloride salts with NaOH, compounds with the composition [M2+0.667Al3+0.333(OH)2](NO3)0.333.nH2O (M2+ = Mn, Zn) were obtained (equations 5,6), having respective basal distances of 8.92 and 7.92 Å and presenting typical bands in the FTIR spectra.

(5)0.667M2+(NO3)2+0.333Al(NO)3+NaOH[M0.6672+Al0.3333+(OH)2](NO3)0.333nH2O(M2+=Mn,Zn)
(6)0.667MnCl2+0.333AlCl3+NaOH[Mn0.6672+Al0.3333+(OH)2](Cl)0.333nH2O

After exchanging the samples intercalated with nitrate and chloride with Cs2SO4 in excess, the basal distance changed to around 11 Å and typical FTIR bands of SO42-were observed, indicating the intercalation of SO42-. But the content of cesium was very low, suggesting the maintenance of the formula [M2+0.667Al3+0.333(OH)2](SO4)0.167.nH2O (equation 7), as also proposed by equation 1.

(7)[M2+0.667Al3+0.333(OH)2](A-)0.333nH2O+Cs2SO4 (excess) [M2+0.667Al3+0.333(OH)2](SO4)0.167nH2O(A=NO3-,Cl-;M2+=Zn,Mn)

This is the first report in the literature of synthesis and characterization of LDHs intercalated with SO42-/Cs+/Na+ or SO42-/Cs+/NH4+, opening new alternatives to remove radioactive nuclides from contaminated waters using LDHs obtained by co-precipitation syntheses with increasing pH.

Acknowledgments

We acknowledge funding from the Office to Coordinate Improvement of University Personnel (CAPES) - Finance Code 001 for the PhD scholarship of ARS; the National Council for Scientific and Technological Development (CNPq, grant 300988/2019-2); Financier of Studies and Projects (FINEP). We are also grateful to Prof Marco Tadeu Grassi and MSc Mayara Padovan dos Santos for the ICP-OES analyses.

References

  • 1
    Cajipe, V. B.; Heiney, P. A.; Fischer, J. E.; Phys. Rev. B 1989, 39, 4374.
  • 2
    Paulus, W.; Katzke, H.; Schöllhorn, R.; J. Solid State Chem 1992, 96, 162.
  • 3
    Messaoudi, A.; Conard, J.; Setton, R.; Beguin, F.; Chem. Phys. Lett 1993, 202, 506.
  • 4
    Tonti, D.; Pettenkofer, C.; Jaegermann, W.; Papageorgopoulos, D. C.; Kamaratos, M.; Papageorgopoulos, C. A.; Ionics 1998, 4, 93.
  • 5
    Silipigni, L.; Acacia, N.; Quattrone, T.; de Luca, G.; Scolaro, L. M.; Salvato, G.; J. Appl. Phys 2007, 102, 113713.
  • 6
    Park, C. W.; Kim, B. H.; Yang, H. M.; Seo, B. K.; Moon, J. K.; Lee, K. W.; Chemosphere 2017, 168, 1068.
  • 7
    Mao, Y. L.; Fang, Y. Q.; Pan, J.; Wang, D.; Bu, K. J.; Che, X. L.; Zhao, W.; Huang, F. Q.; J. Solid State Chem 2019, 279, 120937.
  • 8
    Christiansen, B. C.; Dideriksen, K.; Katz, A.; Nedel, S.; Bovet, N.; Sørensen, H. O.; Frandsen, C.; Gundlach, C.; Andersson, M. P.; Stipp, S. L. S.; Inorg. Chem 2014, 53, 8887.
  • 9
    Mukai, H.; Hirose, A.; Motai, S.; Kiruchi, R.; Tanoi, K.; Nakanishi, T. M.; Yaiti, T.; Yaita, T.; Kogure, T.; Sci. Rep 2016, 21543.
  • 10
    Naamen, S.; Jâafar, N.; Rhaiem, H. B.; Amara, A. B. H.; Plançon, A.; Muller, F.; Clay Miner 2016, 51, 29.
  • 11
    Park, C. W.; Kim, S. M.; Kim, I.; Yoon, I. H.; Hwang, J.; Kim, J. H.; Yang, H. M.; Seo, B. K.; J. Environ. Radioact. 2021, 233, 106592.
  • 12
    Clay Surfaces - Fundamentals and Applications, vol. 1; Wypych, F.; Satyanarayana, K. G., eds.; Academic Press: Amsterdam, 2004.
  • 13
    Layered Double Hydroxides; Duan, X.; Evans, D. G., eds.; Springer Verlag: Berlin, 2006.
  • 14
    Wang, Q.; O’Hare, D.; Chem. Rev. 2012, 112, 4124.
  • 15
    Rives, V.; del Arco, M.; Martín, C.; Appl. Clay Sci 2014, 88-89, 239.
  • 16
    Leroux, F.; Moujahid, E.; Taviot-Gueho, C.; Besse, J. P.; Solid State Sci 2001, 3, 81.
  • 17
    Radha, A. V.; Kamath, P. V.; Shivakumara, C.; Solid State Sci. 2005, 7, 1180.
  • 18
    Williams, G. R.; Rees, N. H.; O’Hare, D.; Solid State Sci 2009, 11, 1229.
  • 19
    Zhang, F. R.; Du, N.; Li, H. P.; Liu, J. Q.; Hou, W. G.; Solid State Sci 2014, 32, 41.
  • 20
    Delmas, C.; Borthomieu, Y.; J. Solid State Chem 1993, 104, 345.
  • 21
    Sotiles, A. R.; Wypych, F.; Chem. Commun 2019, 55, 7824.
  • 22
    Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc 2019, 141, 531.
  • 23
    Sotiles, A. R.; Grassi, M. T.; dos Santos, M. P.; Wypych, F.; J. Braz. Chem. Soc 2021, 32, 170.
  • 24
    Sotiles, A. R.; Wypych, F.; Solid State Sci 2020, 106, 106304.
  • 25
    Rodgers, K. A.; Chisholm, J. E.; Davis, R. J.; Nelson, C. S.; Mineral. Mag. 1977, 41, 389.
  • 26
    Zamareno, I.; Plana, F.; Vazques, A.; Clague, D. A.; Am. Mineral. 1989, 74, 1054.
  • 27
    Witzke, T.; Pöllmann, H.; Vogel, A.; Z. Kristallogr 1995, 9, S.252.
  • 28
    Cooper, M. A.; Hawthorne, F. C.; Can. Mineral 1996, 34, 91.
  • 29
    Huminicki, D. M. C.; Hawthorne, F. C.; Can. Mineral 2003, 41, 79.
  • 30
    Christiansen, B. C.; Balic-Zunic, T.; Petit, P. O.; Frandsen, C.; Mørup, S.; Geckeis, H.; Katerinopoulou, A.; Stipp, S. L. S.; Geochim. Cosmochim. Acta 2009, 73, 3579.
  • 31
    Mills, S. J.; Christy, A. G.; Genin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289.
  • 32
    Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem 1995, 34, 883.
  • 33
    Radha, S.; Jayanthi, K.; Breu, J.; Kamath, P. V.; Clays Clay Miner 2014, 62, 53.
  • 34
    Persson, I.; Pure Appl. Chem 2010, 82, 1901.
  • 35
    Brugé, F.; Bernasconi, M.; Parrinello, M.; J. Am. Chem. Soc 1999, 121, 10883.
  • 36
    Guo, J.; Zhou, L.; Zen, A.; Michaelides, A.; Wu, X.; Wang, E.; Xu, L.; Chen, J.; Phys. Rev. Lett 2020, 125, 106001.
  • 37
    Khaldi, M.; de Roy, A.; Chaouch, M.; Besse, J. P.; J. Solid State Chem. 1997, 130, 66.
  • 38
    Pironon, J.; Pelletier, M.; de Donato, P.; Mosser-Ruck, R.; Clay Miner 2003, 38, 201.
  • 39
    Aranda, P.; Ruiz-Hitzky, E.; Appl. Clay Sci 1999, 15, 119.
  • 40
    Nakamoto, K.; Infrared and Raman Spectra of Inorganic and Coordination Compounds, Part A: Theory and Applications in Inorganic Chemistry, 5th ed.; Wiley: New York, 1997.
  • 41
    Iftekhar, S.; Küçük, M. E.; Srivastava, V.; Repo, E.; Sillanpa, M.; Chemosphere 2018, 209, 470.
  • 42
    Adiwidjaja, G.; Friese, K.; Klaska, K. H.; Schlüter, J.; Z. Kristallogr 1997, 212, 704.
  • 43
    Maruyama, S. A.; Krause, F.; Tavares, S. R.; Leitão, A. A.; Wypych, F.; Appl. Clay Sci 2017, 146, 100.
  • 44
    Hawthorne, F. C.; Kimata, M.; Eby, R. K.; Am. Mineral 1993, 78, 649.
  • 45
    Bindi, L.; Christy, A. G.; Mills, S. J.; Ciriotti, M. E.; Bittarello, E.; Can. Mineral 2015, 53, 791.
  • 46
    Witzke, T.; Raade, G.; Neues Jahrb. Mineral., Monatsh. 2000, 455.
  • 47
    Nickel, E. H.; Graham, J.; Can. Mineral 1987, 25, 409.
  • 48
    Zegeye, A.; Ona-Nguema, G.; Carteret, C.; Huguet, L.; Abdelmoula, M.; Jorand, F.; Geomicrobiol. J. 2005, 22, 389.
  • 49
    Marappa, S.; Radha, S.; Kamath, P. V.; Eur. J. Inorg. Chem 2013, 12, 2122.
  • 50
    Bookin, A. S.; Drits, V. A.; Clays Clay Miner 1993, 41, 551.
  • 51
    Pachayappan, L.; Kamath, P. V.; Clays Clay Miner 2019, 67, 154.
  • 52
    Radha, S.; Antonyraj, C. A.; Kamath, P. V.; Kannan, S.; Z. Anorg. Allg. Chem. 2010, 636, 2658.
  • 53
    Balan, E.; Lazzeri, M.; Morin, G.; Mauri, F.; Am. Mineral 2006, 91, 115.
  • 54
    Radha, S.; Kamath, P. V.; Inorg. Chem. 2013, 52, 4834.

Edited by

Editors handled this article: Jaísa Fernandes Soares and Pedro H. C. Camargo (Associate)

Publication Dates

  • Publication in this collection
    10 Jan 2022
  • Date of issue
    Jan 2022

History

  • Received
    06 July 2021
  • Accepted
    09 Sept 2021
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br