Acessibilidade / Reportar erro

Synthesis, Characterization and Exchange Reactions of Layered Double Hydroxides of Copper and Aluminum, Intercalated with Sulfate

Abstract

Layered double hydroxides (LDHs) of Cu:Al in the molar ratio of 2:1, intercalated with sulfate, sulfate/(Li+, Na+, K+ or NH4+), NO3- and CO32-, were synthesized by co-precipitation with increasing pH. The materials were submitted to exchange reactions using B2SO4 (B = Li+, Na+, K+, NH4+) solutions in an attempt to replace previously intercalated cations or incorporated cations without removing intercalated sulfate. X-ray diffraction (XRD) patterns and Fourier transform infrared (FTIR) spectra were consistent with the expected intercalated species and scanning electron microscopy (SEM) images indicated submicrometric platelet-like particles, typical of LDHs. The chemical compositions of all phases were confirmed by inductively coupled plasma optical emission spectrometry (ICP OES) and thermogravimetric analyses (TGA). In the exchange reactions, only in [Cu6Al3(OH)18][Na(H2O)6(SO4)2]·6H2O the sodium cations were almost totally replaced with lithium, potassium and ammonium, without removing the intercalated sulfate.

Keywords:
layered double hydroxide; sulfate; exchange reaction; copper; aluminum


Introduction

Layered double hydroxides (LDHs) are a class of materials belonging to the hydrotalcite-like compounds. These compounds have brucite-like structure (Mg(OH)2), where each metallic cation occupies the center of an octahedron whose vertices contain hydroxide anions. The octahedra share edges to form two-dimensional layers, which are stacked along the basal axis. However, in the case of LDHs, part of the M2+ cations are replaced by M3+ cations, generating an excess of positive charges on the layer. This excess of positive charges is compensated by the insertion of anions in the space between layers.11 Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem. 1995, 34, 883.

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

3 Chubar, N.; Gilmour, R.; Gerda, V.; Mičušík, M.; Omastova, M.; Heister, K.; Man, P.; Fraissard, J.; Zaitsev, V.; Adv. Colloid Interface Sci. 2017, 245, 62.
-44 Chen, Y.; Jing, C.; Zhang, X.; Jiang, D.; Liu, X.; Dong, B.; Feng, L.; Li, S.; Zhang, Y.; J. Colloid Interface Sci. 2019, 548, 100.

LDHs are represented by the general composition M2+1xM3+xOH2Anx/n·yH2O, where M2+ and M3+ are divalent and trivalent metal cations and An- denotes intercalated anhydrous or hydrated anions with charge n-.55 Cavani, F.; Trifiro, F.; Vaccari, A.; Catal. Today 1991, 11, 173.

6 Crepaldi, E. L.; Valim, J. B.; Quim. Nova 1998, 21, 300.
-77 Chitrakar, R.; Sonoda, A.; Makita, Y.; Hirotsu, T.; Sep. Purif. Technol. 2011, 80, 652. Several combinations between M2+ and M3+ with different intercalated anions have been studied, and new phases are being synthesized and evaluated regarding the formation of compounds, their properties and applications.88 Miyata, S.; Clays Clay Miner. 1983, 31, 305.

9 Meyn, M.; Beneke, K.; Lagaly, G.; Inorg. Chem. 1990, 29, 5201.

10 Bravo-Suárez, J. J.; Páez-Mozo, E. A.; Oyama, S. T.; Quim. Nova 2004, 27, 601.

11 Chang, Z.; Evans, D. G.; Duan, X.; Vial, C.; Ghanbaja, J.; Prevot, V.; de Roy, M.; Forano, C.; J. Solid State Chem. 2005, 178, 2766.
-1212 Basu, D.; Das, A.; Stöckelhuber, K. W.; Wagenknecht, U.; Heinrich, G.; Prog. Polym. Sci. 2014, 39, 594.

Due to the presence of positively charged layers, these compounds are well known for their anion exchange capacity. However, it has recently been reported1313 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531. that some LDHs have the ability to intercalate cations and anions and exchange both simultaneously or separately. Among the anions studied, sulfate has been gaining prominence, mainly because LDHs are similar to the minerals motukoreaite (Mg/Al), natroglaucocerinite (Zn/Al) and shigaite (Mn/Al),1414 Rodgers, K. A.; Chisholm, J. E.; Davis, R. J.; Nelson, C. S.; Mineral. Mag. 1977, 41, 389.

15 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91.
-1616 Mills, S. J.; Christy, A. G.; Génin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289. in addition to other phases containing Co/Al and Ni/Al.1717 Sotiles, A. R.; Gomez, N. A. G.; dos Santos, M. P.; Grassi, M. T.; Wypych, F.; Appl. Clay Sci. 2019, 181, 105217. These have been synthesized with sulfate, and intercalation of alkali metal cations such as lithium, sodium and potassium has been observed.1313 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531.,1717 Sotiles, A. R.; Gomez, N. A. G.; dos Santos, M. P.; Grassi, M. T.; Wypych, F.; Appl. Clay Sci. 2019, 181, 105217.

In recent years, layered compounds have received great attention due to their wide applications, such as protection against UV radiation in polymers,1818 Gomez, N. A. G.; Wypych, F.; J. Polym. Res. 2019, 26, 203. emulsion stabilization,1919 Zhang, N.; Zhang, L.; Sun, D.; Langmuir 2015, 31, 4619. catalysis,2020 Nakagaki, S.; Castro, K. A. D. F.; Ucoski, G. M.; Halma, M.; Prévot, V.; Forano, C.; Wypych, F.; J. Braz. Chem. Soc. 2014, 25, 2329.,2121 Castro, K. A. D. F.; Halma, M.; Machado, G. S.; Ricci, G. P.; Ucoski, G. M.; Ciuffi, K. J.; Nakagaki, S.; J. Braz. Chem. Soc. 2010, 21, 1329. adsorption,2222 Zhou, H.; Jiang, Z.; Wei, S.; Appl. Clay Sci. 2018, 153, 29. degradation of dyes and organic compounds,2323 Wang, X.; Wu, P.; Lu, Y.; Huang, Z.; Zhu, N.; Lin, C.; Dang, Z.; Sep. Purif. Technol. 2014, 132, 195.,2424 Wang, H.; Zhang, Z.; Jing, M.; Tang, S.; Wu, Y.; Liu, W.; Appl. Clay Sci. 2020, 186, 105433. drug release,2525 Bini, M.; Monteforte, F.; Quinzeni, I.; Friuli, V.; Maggi, L.; Bruni, G.; J. Solid State Chem. 2019, 272, 131.

26 Yasaei, M.; Khakbiz, M.; Zamanian, A.; Ghasemi, E.; Mater. Sci. Eng., C 2019, 103, 109816.
-2727 Yasaei, M.; Khakbiz, M.; Ghasemi, E.; Zamania, A.; Appl. Surf. Sci. 2019, 467-468, 782. UV-radiation polymer protection2828 Gómez, N. A. G.; Silva, G. M.; Wilhelm, H. M.; Wypych, F.; J. Braz. Chem. Soc. 2020, 31, 971. and flame retardance.2929 Jaerger, S.; Wypych, F.; J. Appl. Polym. Sci. 2019, 48737. Since LDHs are materials obtained at relatively low cost, it is attractive to increase the study of these compounds, including synthesizing new phases that can have diversified properties for diverse applications. An element not often analyzed in structures of layered compounds and that can be further explored is copper. Most of the studies3030 Suzuki, J.; Ito, M.; Sugiura, T.; J. Jpn. Assoc. Mineral., Petrol. Econ. Geol. 1976, 71, 183.

31 Sarp, H.; Perroud, P.; N. Jb. Miner. Mh., Jg. 1991, 11, 481.

32 Cuchet, S.; Schweiz. Mineral. Petrogr. Mitt. 1995, 75, 283.

33 Frost, R. L.; Keeffe, E. C.; Spectrochim. Acta, Part A 2011, 81, 111.

34 Mills, S. J.; Christy, A. G.; Schnyder, C.; Favreau, G.; Price, J. R.; Mineral. Mag. 2014, 78, 1527.

35 Mills, S. J.; Christy, A. G.; Colombo, F.; Price, J. R.; Mineral. Mag. 2015, 79, 321.

36 Mills, S. J.; Christy, A. G.; Favreaud, G.; Galea-Cloluse, V.; Acta Crystallogr., Sect. B 2017, 73, 950.
-3737 Ventruti, G.; Mugnaioli, E.; Capitani, G.; Scordari, F.; Pinto, D.; Lausi, A.; Phys. Chem. Miner. 2015, 42, 651. that exist involve the characterization of minerals obtained in nature.

The general information in the literature55 Cavani, F.; Trifiro, F.; Vaccari, A.; Catal. Today 1991, 11, 173.,3838 Boclair, J. W.; Braterman, P. S.; Chem. Mater. 1998, 10, 2050.,3939 Layrac, G.; Harrisson, S.; Destarac, M.; Gérardin, C.; Tichit, D.; Appl. Clay Sci. 2020, 193, 105673. indicates that the proposed compositions investigated in the present article are unlikely to be obtained due especially to: (i) the precipitation of the isolated hydroxides like Cu(OH)2 and Al(OH)3 in the pH lower (ca. 5.0) than those used to precipitate the respective LDH (pH above 7); (ii) to the Jahn-Teller effect observed with Cu2+ when hexacoordinated with hydroxide anions, leading to poor long-range ordering of the octahedral and hindering the LDH structure formation; (iii) to the Jahn-Teller effect Cu2+ site distortion, weakening the electrostatic interactions between the positively charged layers and hydrated intercalated anions; (iv) that LDH containing Cu2+ will only be obtained when a third M2+ metal is used in higher concentration together with Cu2+ forming ternary LDH, which would overpass the Jahn-Teller effect, by diluting the Cu2+ in the brucite-like structure.

In spite of these discouraging effects we decided to investigate the Cu2Al-SO4 LDH system in an attempt to expand the knowledge of this class of compound. The main incentives were the scarce literature about the intercalation of sulfate in these phases and the information that Cu/Al LDHs are potential catalysts or catalysts precursors for application in different reactions (e.g., in photocatalysis, oxidation, hydrogenation, dehydration, isomerization, steam reforming, conversion of biomass, etc.).4040 Fan, G.; Li, F.; Evans, D. G.; Duan, X.; Chem. Soc. Rev. 2014, 43, 7040.

41 Yan, K.; Liu, Y.; Lu, Y.; Chai, J.; Sun, L.; Catal. Sci. Technol. 2017, 7, 1622.
-4242 Li, J.; Zhang, S.; Chen, Y.; Liu, T.; Liu, C.; Zhang, X.; Yi, M.; Chu, Z.; Han, X.; RSC Adv. 2017, 7, 29051.

Experimental

LDHs with Cu2+:Al3+ molar ratios of 2:1 were synthesized by coprecipitation with increasing pH using an automatic glass titration reactor operating at 50 °C, under N2 flow. The chemicals were of analytical grade and used without any treatment (LiOH, Biotec (São Paulo, Brazil), 98%; NaOH, Reatec (São Paulo, Brazil), 99%; KOH, Reatec, 98%; CuSO4, Reatec, 99%; Al2(SO4)3.16H2O, Reatec, 98-102%; Li2SO4, Reatec, 99%; Na2SO4, Neon (São Paulo, Brazil), 99.9%; K2SO4, Reatec, 98.5%; (NH4)2SO4, Reatec, 99%; NaNO3, F. Maia (Belo Horizonte, Brazil), 99%; Na2CO3, Biotec, 99%). The amounts used in the synthesis of the samples are presented in Table 1.

Table 1
Solution concentrations and pH changes during synthesis

Using the phase Cu2Al-SO4/Na as an example, CuSO4·5H2O, Al2(SO4)3 and Na2SO4 were dissolved in 100 mL of Milli-Q water (Millipore-simplicity UV, Bedford, USA). The solution was heated to 50 °C and very slowly titrated with a solution of 1 mol L-1 NaOH in a glass titration reactor, using a peristaltic pump coupled to a pHmeter, having the pH monitored all the time during the titration. After observing that solid materials were obtained, the slurries were removed from the reactor at desired pHs and aged at 90 °C for 120 h in a capped Erlenmeyer flask. All the samples were separated by centrifugation at 4000 rpm (centrifugal force of 2125 G) for 5 min. To minimize the mechanically trapped soluble species, a new portion of around 50 mL of N2 decarbonated distilled water (by boiling and bubbling N2 gas) was added to the slurry and the tube submitted to one ultrasound bath for some seconds. The centrifugation was performed, and the process repeated at least five times and finally dried at room temperature. As per our experience in the lab, filtration is not a good procedure to remove all soluble species trapped between the particles in the cake.

After defining the optimal pH for synthesis of more crystalline materials (using the procedure of observing the sharp peaks and the higher number of the basal peaks in the XRD patterns), the samples were also synthesized with sulfate salts of lithium, potassium and ammonium and precipitated with 1 mol L-1 solutions of the respective hydroxides. Samples were also intercalated with nitrate and carbonate and precipitated with a 1 mol L-1 NaOH solution. To observe the effect of the temperature in the sample crystallinity, the lithium sample was not hydrothermally treated at 90 °C, only aged at room temperature for 120 h.

Based on the sodium intercalated phase as an example of the exchange reactions, an aqueous dispersion of the solid was magnetically stirred slowly with excess Li2SO4, K2SO4 or (NH4)2SO4 (three times the concentration of the intercalated cations) for 120 h. The attempt to substitute nitrate and carbonate with sulfate occurred with an excess of Na2SO4 (five times the concentration of the intercalated cations) for 240 h.

The reactions were performed at room temperature under N2 flow to avoid contamination with carbonate. After the reactions, the materials were centrifuged at 4000 rpm, washed several times with decarbonated distilled water and dried at room temperature. For the exchange reactions in the other synthesized phases, combinations of different alkaline metal sulfates were used.

The compounds were characterized by X-ray diffraction (XRD) using a Shimadzu XRD-6000 diffractometer with Cu Kα = 1.5418 Å radiation, tension of 40 kV, current of 30 mA and dwell time of 2° min-1 (the step was of 0.02° in 2θ). For the analysis, the samples were dispersed in water after the last washing and were deposited in the glass sample holders and allowed to dry at room temperature.

Fourier-transform infrared (FTIR) spectra were obtained in the transmission mode using a Bomen MB100 spectrophotometer using KBr pellets containing around 1% (m/m) of the sample. The spectra were collected from 400-4000 cm-1, with 32 scans, using resolution of 2 cm-1.

The morphology was investigated by scanning electron microscopy (SEM). The images were acquired with a Tescan Vega3LMU microscope with AZ Tech software. The samples were deposited on carbon tapes and sputtered with a thin gold layer.

The quantitative determinations of the metals and sulfur (relative to sulfate) were quantified with a Thermo Scientific inductively coupled plasma optical emission spectrometer (ICP OES, model iCAP 6500) with the Thermo Scientific iTeVa software, version 1.2.0.30. The samples were dissolved in a solution containing 1.0% v/v of HNO3 in Milli-Q water and the data were collected in duplicate. Average values were used to obtain the LDH formulas.

Thermogravimetric analyses (TGA) were performed with a PerkinElmer TGA 4000 equipment, under synthetic air atmosphere with a flow rate of 50 mL min-1 and heating rate of 10 °C min-1.

Results and Discussion

In spite of the antecedents in the literature2929 Jaerger, S.; Wypych, F.; J. Appl. Polym. Sci. 2019, 48737.,4343 Britto, S.; Radha, A.; Ravishankar, N.; Kamath, P. V.; Solid State Sci. 2007, 9, 279. indicating the difficulty or impossibility to obtain crystalline Cu rich Al LDH phases, it was a surprise when the XRD patterns of most of samples (Figure 1A) presented the typical pattern of this class of compounds, with a series of strong basal diffraction peaks due to the natural orientation of the layered crystals in the sample holder plane. The basal distances of compounds, indicated in the figure, were determined by Bragg’s law using the peak of highest order (around 25° in 2θ).

Figure 1
(A) XRD patterns and (B) FTIR spectra of the Cu2Al-SO4/Na synthesized at different pH values: (a) 7.03; (b) 7.57; (c) 8.04; (d) 8.51; (e) 9.07; (f) 9.55 and (g) 10.03.

For all samples of Cu2Al-SO4/Na obtained in the investigated pH range, the basal distance was close to 11 Å, with slight reduction when the pH of synthesis increased. In all cases, the basal distance was typical of systems containing sulfate and alkali metals, obtained by direct synthesis or exchange reactions.1313 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531.,4444 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824. At pH = 8.51, contamination was found with basal spacing of 8.7 Å, which is consistent with the intercalation of dehydrated sulfate.11 Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem. 1995, 34, 883. After pH = 8.51, the samples lost crystallinity, tending to amorphous materials at pH higher than 10. In all samples, the (100) diffraction peaks were observed, indicating that the sample had a superstructure of the a’ = a√3 × a√3 type (a’ = 5.33 Å, a = 3.08 Å, average distance between the metals in the brucite-like layers), as expected for LDHs with 2M2+:M3+, due to the metal cations’ ordering in the two-dimensional layers.1616 Mills, S. J.; Christy, A. G.; Génin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289.,4444 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824.

45 Krivovichev, S. V.; Yakovenchuk, V. N.; Zolotarev, A. A.; Ivanyuk, G. N.; Pakhomovski, Y. A.; Chimia 2010, 64, 730.
-4646 Merlino, S.; Orlandi, P.; Am. Mineral. 2001, 86, 1293. The “a” parameter is very close to that of Cu/Al-CO3 LDH4747 Zhu, Y.; Rong, J.; Zhang, T.; Xu, J.; Dai, Y.; Qiu, F.; ACS Appl. Nano Mater. 2018, 1, 284. (a = 3.08 Å) and of the mineral woodwardite (Cu4Al2(OH)12(SO4)(H2O)2-4) (a = 3.10 Å)4848 Nickel, E.; Mineral. Mag. 1976, 43, 644. and hydrowoodwardite (Cu1-xAlx(OH)2(SO4)x/2·nH2O) (a = 3.07 Å).4949 Witzke, T.; Neues Jahrb. Mineral., Monatsh. 1999, 2, 75. For comparison purposes, the basal distances of hydrowoodwardite, shigaite, motukoreaite and natroglaucocerinite are of 10.93, 11.02, 11.17 and 11.18 Å, respectively.1414 Rodgers, K. A.; Chisholm, J. E.; Davis, R. J.; Nelson, C. S.; Mineral. Mag. 1977, 41, 389.

15 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91.
-1616 Mills, S. J.; Christy, A. G.; Génin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289.,4949 Witzke, T.; Neues Jahrb. Mineral., Monatsh. 1999, 2, 75. Woodwardite is the dehydrated analogous of hydrowoodwardite, consequently the basal distance is of 8.92 Å.4848 Nickel, E.; Mineral. Mag. 1976, 43, 644. This value is close to the impurity observed in the phase Cu2Al-SO4/Na obtained at pH = 8.51 (Figure 1A, curve d) and also in the low crystalline phase Cu2Al-SO4/Li (Figure 2A, curve a).

Figure 2
(A) XRD patterns and (B) FTIR spectra of (a) Cu2Al-SO4/Li; (b) Cu2Al-SO4/Na; (c) Cu2Al-SO4/K; (d) Cu2Al-SO4/NH4; (e) Cu2Al-NO3 and (f) Cu2Al-CO3. The pHs of the syntheses are described in Table 1.

Very similar FTIR spectra were obtained for all the samples (Figure 1B), with the characteristic bands of the O-H bond stretching (3400 cm-1 region), referring to the structure of hydroxyls and adsorbed water molecules, besides the band at 1640 cm-1, corresponding to water molecule bending.44 Chen, Y.; Jing, C.; Zhang, X.; Jiang, D.; Liu, X.; Dong, B.; Feng, L.; Li, S.; Zhang, Y.; J. Colloid Interface Sci. 2019, 548, 100.,5050 Khaldi, M.; de Roy, A.; Chaouch, M.; Besse J. P.; J. Solid State Chem. 1997, 130, 66.

51 Frost, R. L.; Kloprogge, J. T.; Spectrochim. Acta, Part A 1999, 55, 2195.

52 Frost, R. L.; Weier, M. L.; Clissold, M. E.; Williams, P. A.; Spectrochim. Acta, Part A 2003, 59, 3313.
-5353 Badreddine, M.; Khaldi, M.; Legrouri, A.; Barroug, A.; Chaouch, M.; de Roy, A.; Besse, J. P.; Mater. Chem. Phys. 1998, 52, 235. The band in the region of 1120 cm-1 was attributed to the ν3 asymmetrical bending, while the bands at 960 and 620 cm-1 were attributed to ν1 and ν4 S-O vibrations. The broadening of the bands suggests that sulfate anions are in a highly distorted environment. All compounds also presented typical bands in the region of 425-450, 533 and 735 cm-1, which can be attributed to the O-M-O deformation mode and M-O stretching vibrations.1111 Chang, Z.; Evans, D. G.; Duan, X.; Vial, C.; Ghanbaja, J.; Prevot, V.; de Roy, M.; Forano, C.; J. Solid State Chem. 2005, 178, 2766.,5252 Frost, R. L.; Weier, M. L.; Clissold, M. E.; Williams, P. A.; Spectrochim. Acta, Part A 2003, 59, 3313.

53 Badreddine, M.; Khaldi, M.; Legrouri, A.; Barroug, A.; Chaouch, M.; de Roy, A.; Besse, J. P.; Mater. Chem. Phys. 1998, 52, 235.

54 Badreddine, M.; Legrouri, A.; Barroug, A.; de Roy, A.; Besse, J. P.; Mater. Lett. 1999, 38, 391.

55 Mahjoubi, F. Z.; Khalid, A.; Abdennouri, M.; Barka, N.; J. Taibah Univ. Sci. 2017, 11, 90.
-5656 Faramawy, S.; Zaki, T.; Sakr, A. A. E.; Saber, O.; Aboul-Gheit, A. K.; Hassan, S. A.; J. Nat. Gas Sci. Eng. 2018, 54, 72. FTIR spectra are consistent with other copper-containing LDHs and minerals containing sulfate and alkali metals.3434 Mills, S. J.; Christy, A. G.; Schnyder, C.; Favreau, G.; Price, J. R.; Mineral. Mag. 2014, 78, 1527.,3535 Mills, S. J.; Christy, A. G.; Colombo, F.; Price, J. R.; Mineral. Mag. 2015, 79, 321.,3737 Ventruti, G.; Mugnaioli, E.; Capitani, G.; Scordari, F.; Pinto, D.; Lausi, A.; Phys. Chem. Miner. 2015, 42, 651.,5757 Hager, S. L.; Leverett, P.; Williams, P. A.; Can. Mineral. 2009, 47, 635.

In the LDH of Al with other metals like Mn2+, Mg2+ and Zn2+, the bands were sharper, indicating a defined sulfate environment,1313 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531. but were also broad in the LDHs of Al with Ni2+ and Co2+.1717 Sotiles, A. R.; Gomez, N. A. G.; dos Santos, M. P.; Grassi, M. T.; Wypych, F.; Appl. Clay Sci. 2019, 181, 105217. As observed for XRD patterns, after pH 8.51, some FTIR bands became very weak and even disappeared in the range below 1000 cm-1, but sulfate was still detected in all of them.

Several attempts were made to synthesize the phase of Cu2Al-SO4/NH4, but after changing the pH and temperature of synthesis, the reproducibility was very difficult. Phases with basal distances of 8.20-8.5 and 10.7 Å were obtained even under the same synthesis conditions. These phases were consistent with the structure of woodwardite (Cu4Al2(OH)12(SO4)(H2O)2-4)4949 Witzke, T.; Neues Jahrb. Mineral., Monatsh. 1999, 2, 75. and a synthetic analog with the composition Cu0.67Al0.33(OH)2(SO4)0.15(CO3)0.015·0.5H2O,5858 Park, Y.; Kuroda, K.; Kato, C.; Solid State Ionics 1990, 42, 197.,5959 Jayanthi, K.; Kamath, P. V.; Periyasamy, G.; Eur. J. Inorg. Chem. 2017, 3675. respectively with basal distances of 8.92 and 8.58 Å, close to the reported values6060 Radha, S.; Antonyraj, C. A.; Kamath, P. V.; Kannan, S.; Z. Anorg. Allg. Chem. 2010, 636, 2658. for LDH without copper, which are highly dependent on the degree of hydration.

The phase of 10.7 Å can be related to the presence of ammonium and excess sulfate as in shigaite-like structures,1414 Rodgers, K. A.; Chisholm, J. E.; Davis, R. J.; Nelson, C. S.; Mineral. Mag. 1977, 41, 389.

15 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91.
-1616 Mills, S. J.; Christy, A. G.; Génin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289. or materials containing sulfate with high degree of hydration, as observed in hydrowoodwardite or synthetic hydrated Cu0.667Al0.333(OH)2(SO4)0.1665·nH2O.4646 Merlino, S.; Orlandi, P.; Am. Mineral. 2001, 86, 1293. Due to this difficulty, the phases with the composition Cu2Al-SO4/NH4 and basal distance of 10.75 Å, more frequently observed in the synthesis, will be described.

In the XRD patterns, the phases synthesized with sodium (c = 10.72 Å) and ammonium (c = 10.75 Å) showed intense and defined peaks (Figure 2A, curves b and d), indicating greater crystallinity than the other sulfate phases. Cu2Al-SO4/K (Figure 2A, curve c) (c = 10.53 Å) presented intermediary crystallinity while Cu2Al-SO4/Li (Figure 2A, curve a) (c ca. 8.9 Å) presented the lowest crystallinity due to the absence of the ripening process. Other LDHs intercalated with sulfate or sulfate/alkali metals have been reported1313 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531.,1515 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91.,1717 Sotiles, A. R.; Gomez, N. A. G.; dos Santos, M. P.; Grassi, M. T.; Wypych, F.; Appl. Clay Sci. 2019, 181, 105217.,6161 Sotiles, A. R.; Gomez, N. A. G.; da Silva, S. C.; Wypych, F.; J. Braz. Chem. Soc. 2019, 30, 1807.

62 Huminicki, D. M. C.; Hawthorne, F. C.; Grice, J. D.; Roberts, A. C.; Jambor, J. L.; Mineral. Rec. 2003, 34, 155.

63 Wachowiak, J.; Pieczka, A.; Mineral. Mag. 2016, 80, 277.
-6464 Zamarreño, I.; Plana, F.; Vazquez, A.; Clague, D. A.; Am. Mineral. 1989, 74, 1054. with similar basal spacing, but the sulfate quickly dehydrated after exposure to dry air, which did not happen to our samples.

This indicates that our samples did not contain only sulfate, since samples containing sulfate/alkali metals are stable in air, preserving basal distances close to 11 Å. For the Cu2Al-NO3 and Cu2Al-CO3 (Figure 2A, curves e and f), the basal distances of 8.76 and 7.54 Å, respectively, are characteristic of nitrate and carbonate intercalation.88 Miyata, S.; Clays Clay Miner. 1983, 31, 305.,4747 Zhu, Y.; Rong, J.; Zhang, T.; Xu, J.; Dai, Y.; Qiu, F.; ACS Appl. Nano Mater. 2018, 1, 284.,6565 Kang, H.; Leoni, M.; He, H.; Huang, G.; Yang, X.; Eur. J. Inorg. Chem. 2012, 3859.,6666 Berner, S.; Araya, O.; Govan, J.; Palza, H.; J. Ind. Eng. Chem. 2018, 59, 134. The Cu2Al-NO3 compound showed a diffraction peak with basal distance close to 7.3 Å, which can be attributed to carbonate contamination, although the reactions were performed under N2 flow.

In compounds synthesized with sulfate (Figure 2B, curves a-d), the bands were very similar to those observed in Figure 1B. Cu2Al-NO3 and Cu2Al-CO3 (Figure 2B, curves e and f) showed the characteristic bands of the respective anions intercalated with the stretching vibration of the C-O and N-O bonds, located at 1360 and 1380 cm-1, respectively.33 Chubar, N.; Gilmour, R.; Gerda, V.; Mičušík, M.; Omastova, M.; Heister, K.; Man, P.; Fraissard, J.; Zaitsev, V.; Adv. Colloid Interface Sci. 2017, 245, 62.,2727 Yasaei, M.; Khakbiz, M.; Ghasemi, E.; Zamania, A.; Appl. Surf. Sci. 2019, 467-468, 782.,6767 Kloprogge, J. T.; Wharton, D.; Hickey, L.; Frost, R. L.; Am. Mineral. 2002, 87, 623.

68 Kloprogge, J. T.; Hickey, L.; Frost, R. L.; Mater. Chem. Phys. 2005, 89, 99.
-6969 Liu, J.; Song, J.; Xiao, H.; Zhang, L.; Qin, Y.; Liu, D.; Hou, W.; Du, N.; Powder Technol. 2014, 253, 41. The Cu/Al-NO3 sample showed broadening of the N-O band (1380 cm-1) towards the 1360 cm-1 region, attributed to carbonate contamination, as shown in the XRD pattern. Table 2 reports the chemical composition of the synthesized samples obtained in the ICP OES analysis. For the samples containing lithium and sodium, the results were close to the values used in the synthesis and the chemical composition was consistent with the ideal anhydrous formula [Cu0.667Al0.333(OH)2][B0.111(SO4)0.222] (B = Li+, Na+).

Table 2
Compositions of the synthesized phases obtained by ICP OES analysis

Since XRD patterns indicated basal spacing typical of a compound with SO42-/NH4+ and the FTIR spectra indicated absence of N-H bonds, the formula is close to Cu0.662Al0.338(OH)2(SO4)0.169 but the content of sulfate is slightly higher than predicted. In the case of Cu2Al-SO4/K, the contents of sulfate and potassium are reduced, indicating the possible composition Cu0.667Al0.333(OH)2(SO4)0.1665, although the basal distance also suggests the presence of SO42-/alkali metal (10.53 Å) and ICP OES indicates the presence of some amount of potassium, but much lower than expected (Table 2). The SEM images of all the samples (Figure 3) indicated a compact agglomeration and crystals with submicrometric size and platelet-like morphology, typical of LDHs. The orientation of the layered crystals in the sample holder explains the exposure of the basal planes to diffraction, as observed in the XRD patterns of all samples (Figures 1A and 2A).

Figure 3
SEM images of (a) Cu2Al-SO4/Li; (b) Cu2Al-SO4/Na; (c) Cu2Al-SO4/K; (d) Cu2Al-SO4/NH4; (e) Cu2Al-NO3 and (f) Cu2Al-CO3.

The TGA and corresponding derivative thermogravimetry (DTG) curves (Figure 4) showed the same profile. In general, the release of intercalated and sorbed water occurs up to 220 °C.7070 Karami, Z.; Jouyandeh, M.; Hamad, S. M.; Ganjali, M. R.; Aghazadeh, M.; Torre, L.; Puglia, D.; Saeb, M. R.; Prog. Org. Coat. 2019, 136, 105278. The second event up to 450 °C is attributed to the LDH dehydroxylation process, indicated by the DTG peak, which varied according to the intercalated alkali metal, as expected due to the different hydration energy of the intercalated cations. The third event indicated by the DTG peak in the region of 680-720 °C is attributed to the partial decomposition of sulfates.7171 Kameda, T.; Fubasami, Y.; Yoshioka, T.; J. Therm. Anal. Calorim. 2012, 110, 641. In spite of the deviation between the expected and experimental values of up to 4% due to the partial decomposition of sulfates at 1000 °C (see mass loss under way close to 1000 °C), there were relative good agreements between the experimental residual mass from the theoretical one (Table 3).

Figure 4
TGA/DTG curves of (a) Cu2Al-SO4/Li; (b) Cu2Al-SO4/Na; (c) Cu2Al-SO4/K and (d) Cu2Al-SO4/NH4.

Table 3
Proposed chemical formulas of the evaluated compounds

Although the literature4444 Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824. describes the transformation of chloride and nitrate intercalated LDH into SO42-/B+ (B+ = alkali metal), our attempts to substitute interlayer nitrate and carbonate with sulfate produced materials with very low crystallinity indicating the unsuccessful reaction (data not shown).

The XRD patterns (Figure 5A) and FTIR spectra (Figure 5B) after the alkali metal cation exchange reactions showed basal distances close to 10 Å, typical of the presence of SO42-/alkali metal or hydrated sulfate.

Figure 5
(A) XRD patterns and (B) FTIR spectra of (a) Cu2Al-SO4/Li and after exchange with (b) Na and (c) K; of (d) Cu2Al-SO4/Na and after exchange with (e) Li and (f) K; and of (g) Cu2Al-SO4/K and after exchange with (h) Li and (i) Na.

The XRD patterns of compound Cu2Al-SO4/Li (8.87 Å) (Figure 5A, curve a) showed displacement to higher basal distance after the exchange reactions Li-Na (10.27 Å) and Li-K (10.35 Å) (Figure 5A, curves b and c). The basal distances of the samples Cu2Al-SO4/Na after exchange with Li+ and K+ changed from 10.72 to 10.70 Å (Na-Li) and 10.75 Å (Na-K). In the case of Cu2Al-SO4/K, the basal distances changed from 10.53 to 10.50 Å (K-Li) and 10.61 Å (K-Na). This behavior has been observed previously1313 Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531.,1717 Sotiles, A. R.; Gomez, N. A. G.; dos Santos, M. P.; Grassi, M. T.; Wypych, F.; Appl. Clay Sci. 2019, 181, 105217. in similar compounds synthesized with different M2+ and intercalated with sulfate and alkali metals.

The FTIR spectrum of the samples after exchanges (Figure 5B) shows the same bands as the precursors, indicating the maintenance of the basic LDH structure and functional groups. Traces of carbonate were observed in some samples (band at 1380 cm-1), but the concentration was very low. Table 4 presents the ICP OES data of the samples before and after the exchange reactions.

Table 4
Compositions of the samples obtained by ICP OES analysis

It can be seen that the samples Cu2Al-SO4/Li, Cu2Al-SO4/Na and Cu2Al-SO4/NH4, before the exchange reactions present the chemical composition close to the expected ideal formula equivalent to the shigaite-like minerals ([Cu0.667Al0.333(OH)2][B0.111(SO4)0.222]; B = Li+, NH4+),1414 Rodgers, K. A.; Chisholm, J. E.; Davis, R. J.; Nelson, C. S.; Mineral. Mag. 1977, 41, 389.

15 Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91.

16 Mills, S. J.; Christy, A. G.; Génin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289.
-1717 Sotiles, A. R.; Gomez, N. A. G.; dos Santos, M. P.; Grassi, M. T.; Wypych, F.; Appl. Clay Sci. 2019, 181, 105217. while the sample Cu2Al-SO4/K had a composition equivalent to the mineral hydrowoodwardite or synthetic analogous (Cu0.667Al0.333(OH)2(SO4)0.1665).4949 Witzke, T.; Neues Jahrb. Mineral., Monatsh. 1999, 2, 75.,6060 Radha, S.; Antonyraj, C. A.; Kamath, P. V.; Kannan, S.; Z. Anorg. Allg. Chem. 2010, 636, 2658.

After the exchange reactions of Cu2Al-SO4/Li, the content of sulfate and alkali metal was reduced, and the samples tended to be transformed into hydrowoodwardite. The Cu2Al-SO4/Na sample was the only one that preserved the structure of shigaite-like minerals, with the presence of alkali metal and sulfate. Cu2Al-SO4/K after the exchange reactions, although with a small concentration of alkali metal cations, especially after the exchange of K+ with Na+, suggests the preservation of the hydrowoodwardite structure. The same was also observed for the sample Cu2Al-SO4/NH4, after the exchange reactions.

Figure 6A shows the XRD patterns of the sodium and ammonium samples after the exchange reactions.

Figure 6
(A) XRD patterns and (B) FTIR spectra of (a) Cu2Al-SO4/Na and (b) after exchange with NH4, of (c) Cu2Al-SO4/NH4 and (d) after exchange with Na.

The compounds Cu2Al-SO4/Na and Cu2Al-SO4/NH4 showed basal distances of 10.72 and 10.75 Å, respectively. After the exchange reactions, Cu2Al-SO4/Na-NH4 had basal distance of 10.94 Å, while Cu2Al-SO4/NH4-Na had basal distance of 10.73 Å. By XRD analysis, it was not possible to tell whether the cation exchange occurred, nor was this possible from the FTIR spectra (Figure 6B), which showed only broadening of some bands due to disturbance of the sulfate/Na+ or NH4+ environment. The samples with the composition [Cu0.667Al0.333(OH)2][B0.111(SO4)0.222] (B = Li+, Na+, K+ and NH4+) and Cu0.667Al0.333(OH)2(SO4)0.1665 have very similar basal distances and only precise chemical analyses would give the information about the real compositions of the evaluated materials.

The SEM images (Figure 7) of all the Cu/Al like phases after the exchange reactions showed agglomerated particles, but less compacted than in the precursors, possibly due to the magnetic stirring during the reactions.

Figure 7
SEM images of Cu2Al-SO4/Li after exchange with (a) Na and (b) K; of Cu2Al-SO4/Na after exchange with (c) Li, (d) K and (e) NH4; of Cu2Al-SO4/K after exchange with (f) Li and (g) Na; and of Cu2Al-SO4/NH4 after exchange with (h) Na.

The ideal compositions of the synthesized LDHs can be described by equations 1 to 14 (to avoid the indication of soluble product, at the right of the arrows only the solids were included).

Synthesis

Precipitation of Cu and Al sulfates in the presence of sodium and lithium sulfates (B = Li+ or Na+):

(1) 0 . 667 CuSO 4 + 0 . 1665 Al 2 SO 4 3 + B 2 SO 4 + BOH Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 B 0 . 111

Precipitation of Cu and Al sulfates in the presence of potassium and ammonium sulfates (B = K+ or NH4+):

(2) 0 . 667 CuSO 4 + 0 . 1665 Al 2 SO 4 3 + B 2 SO 4 + BOH Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 1665

Precipitation of Cu and Al nitrates in the presence of sodium nitrate:

(3) 0 . 667 Cu NO 3 2 + 0 . 333 Al NO 3 3 + NaOH Cu 0 . 667 Al 0 . 333 OH 2 NO 3 0 . 333

Precipitation of Cu and Al nitrates in the presence of sodium carbonate:

(4) 0 . 667 Cu NO 3 2 + 0 . 333 Al NO 3 3 + Na 2 CO 3 + NaOH Cu 0 . 667 Al 0 . 333 OH 2 CO 3 0 . 1665

Exchange reactions

Using Li2SO4, K2SO4, Na2SO4, (NH4)2SO4 (B = Li+, Na+, K+, NH4+):

(5) Cu 0 . 667 AL 0 . 333 OH 2 CO 3 0 . 1665 + B 2 SO 4 no apparent reaction
(6) Cu 0 . 667 Al 0 . 333 OH 2 NO 3 0 . 333 + B 2 SO 4 no apparent reaction
(7) Cu 0 . 667 Al 0 . 333 OH 2 CO 3 0 . 1665 + B 2 SO 4 no apparent reaction

Using Cu2Al-SO4/Li phase and different alkali metals or ammonium sulfate salts:

(8) Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Li 0 . 111 + Na 2 SO 4 Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 y CO 3 y Li 0 . 111 x Na x
(9) Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Li 0 . 111 + K 2 SO 4 Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 y CO 3 y Li 0 . 111 x K x
(10) Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Li 0 . 111 + NH 4 2 SO 4 Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 y CO 3 y Li 0 . 111 x NH 4 x

After extending the reaction time:

(11) Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Li 0 . 111 + B 2 SO 4 Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 1665 B = Na + , K + , NH 4 +

Using Cu2Al-SO4/Na phase and different alkali metals or ammonium sulfate salts:

(12) Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Na 0 . 111 + Li 2 SO 4 Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Na 0 . 111 x Li x Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Li 0 . 111
(13) Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Na 0 . 111 + K 2 SO 4 Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Na 0 . 111 x K x Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 K 0 . 111
(14) Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Na 0 . 111 + NH 4 2 SO 4 Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 Na 0 . 111 x NH 4 x Cu 0 . 667 Al 0 . 333 OH 2 SO 4 0 . 211 NH 4 0 . 111

Only in the case of the system Cu2Al-SO4/Na, real cation exchange reactions occurred. In the schematic representations of both possible structures adopted for Cu0.667Al0.333(OH)2(SO4)0.211Na0.111 or shigaite-like structure, and hydrated Cu0.667Al0.333(OH)2(SO4)0.1665·nH2O (Figure 8), it is difficult to detect the differences by the traditional instrumental techniques used in materials science, especially due to the very close basal distances, low concentration of alkali metals, and slight increase of sulfate amount in the first in comparison to the second.

Figure 8
Schematic representation of both possible structures adopted for Cu0.667Al0.333(OH)2(SO4)0.211Na0.111 and hydrated Cu0.667Al0.333(OH)2(SO4)0.1665.

It is common to detect alkali metal cations in the analysis when the sample is not properly washed, especially when the precipitate is submitted to filtration or centrifugation without dispersing the solid after each centrifugation step. The best technique to differentiate the two structures is quantitative analysis, especially via ICP OES.

Conclusions

Layered double hydroxides of Cu:Al in the molar ratio of 2:1 intercalated with sulfate, nitrate and carbonate were successfully synthesized by co-precipitation at 50 °C, followed by ripening at 90 °C. To investigate the quality of the structures of the precipitated materials, an optimization step was included where solid were removed during the titration process. After identifying the optimized pHs, new synthesis were performed for all compositions. The compounds synthesized with sulfate in the presence of Na+ and K+ had the greatest basal distances (around 11 Å), consistent with the intercalation of sulfate anions and hydrated cations. The same did not occur with the compound synthesized with lithium, which had a basal distance around of 8.9 Å, possibly due to the absence of coordinated molecules in the interlayer space and absence of the Ostwald ripening process.

In the case of the samples synthesized with NH4+, the basal distances varied from 8.2 to 10.7 Å, consistent with the dehydrated and hydrated sulfate, while the samples synthesized with carbonate and nitrate presented basal distances consistent with the intercalation of the respective anions.

SEM images indicated typical LDH morphology and the presence of micrometric or submicrometric particles, while FTIR spectra were similar to those of other LDHs and showed bands consistent with intercalated anions.

ICP OES and thermogravimetric analysis indicated that the phases Cu2Al-SO4/Li and Cu2Al-SO4/Na had the expected composition [Cu6Al3(OH)18][B(H2O)6(SO4)2]·6H2O (B = Li+, Na+), while the phases Cu2Al-SO4/K, in spite of having a small amount of alkali metals and the basal distance expected for the phases containing sulfate/alkali metals or ammonium, had composition close to Cu0.667Al0.333(OH)2(SO4)0.1665, where sulfate was hydrated as in hydrowoodwardite. In the case of Cu2Al-SO4/NH4, sulfate was dehydrated as in woodwardite.

All the samples were submitted to exchange reactions using B2SO4 (B = Li+, Na+, K+, NH4+) solutions in an attempt to replace previously intercalated cations or incorporated cations, without removing intercalated sulfate. The XRD patterns and FTIR spectra were consistent with the expected intercalated species and SEM images indicated submicrometric platelet-like particles, typical of LDHs.

According to ICP OES analysis, Cu2Al-SO4/Li had lower amounts of sulfate and lithium after the exchange reactions, while in Cu2Al-SO4/Na the sodium cations were almost totally replaced with lithium, potassium and ammonium, without removing the intercalated sulfate. Phases containing nitrate or carbonate could not be analyzed since the exchanged with sulfate/alkali metals or ammonium lead to the drastic reduction of crystallinity.

In spite of this contribution, structural aspects and properties of Cu/Al-SO4 LDH still need to be further investigated.

Acknowledgments

This study was financed in part by the Office to Coordinate Improvement of University Personnel (CAPES, finance code 001), National Council for Scientific and Technological Development (CNPq, projects 400117/2016-9 and 300988/2019-2) and the Financier of Studies and Projects (FINEP). A. R. S. thanks CAPES for the PhD scholarship. We also acknowledge Center for Electron Microscopy of Federal University of Paraná (CME-UFPR) for the SEM and Prof Marcos Rogério Mafra (DEQ-UFPR) for the TGA analyses.

References

  • 1
    Constantino, V. R. L.; Pinnavaia, T. J.; Inorg. Chem. 1995, 34, 883.
  • 2
    Wang, Q.; O’Hare, D.; Chem. Rev. 2012, 112, 4124.
  • 3
    Chubar, N.; Gilmour, R.; Gerda, V.; Mičušík, M.; Omastova, M.; Heister, K.; Man, P.; Fraissard, J.; Zaitsev, V.; Adv. Colloid Interface Sci. 2017, 245, 62.
  • 4
    Chen, Y.; Jing, C.; Zhang, X.; Jiang, D.; Liu, X.; Dong, B.; Feng, L.; Li, S.; Zhang, Y.; J. Colloid Interface Sci. 2019, 548, 100.
  • 5
    Cavani, F.; Trifiro, F.; Vaccari, A.; Catal. Today 1991, 11, 173.
  • 6
    Crepaldi, E. L.; Valim, J. B.; Quim. Nova 1998, 21, 300.
  • 7
    Chitrakar, R.; Sonoda, A.; Makita, Y.; Hirotsu, T.; Sep. Purif. Technol. 2011, 80, 652.
  • 8
    Miyata, S.; Clays Clay Miner. 1983, 31, 305.
  • 9
    Meyn, M.; Beneke, K.; Lagaly, G.; Inorg. Chem. 1990, 29, 5201.
  • 10
    Bravo-Suárez, J. J.; Páez-Mozo, E. A.; Oyama, S. T.; Quim. Nova 2004, 27, 601.
  • 11
    Chang, Z.; Evans, D. G.; Duan, X.; Vial, C.; Ghanbaja, J.; Prevot, V.; de Roy, M.; Forano, C.; J. Solid State Chem. 2005, 178, 2766.
  • 12
    Basu, D.; Das, A.; Stöckelhuber, K. W.; Wagenknecht, U.; Heinrich, G.; Prog. Polym. Sci. 2014, 39, 594.
  • 13
    Sotiles, A. R.; Baika, L. M.; Grassi, M. T.; Wypych, F.; J. Am. Chem. Soc. 2019, 141, 531.
  • 14
    Rodgers, K. A.; Chisholm, J. E.; Davis, R. J.; Nelson, C. S.; Mineral. Mag. 1977, 41, 389.
  • 15
    Cooper, M. A.; Hawthorne, F. C.; Can. Mineral. 1996, 34, 91.
  • 16
    Mills, S. J.; Christy, A. G.; Génin, J. M. R.; Kameda, T.; Colombo, F.; Mineral. Mag. 2012, 76, 1289.
  • 17
    Sotiles, A. R.; Gomez, N. A. G.; dos Santos, M. P.; Grassi, M. T.; Wypych, F.; Appl. Clay Sci. 2019, 181, 105217.
  • 18
    Gomez, N. A. G.; Wypych, F.; J. Polym. Res. 2019, 26, 203.
  • 19
    Zhang, N.; Zhang, L.; Sun, D.; Langmuir 2015, 31, 4619.
  • 20
    Nakagaki, S.; Castro, K. A. D. F.; Ucoski, G. M.; Halma, M.; Prévot, V.; Forano, C.; Wypych, F.; J. Braz. Chem. Soc. 2014, 25, 2329.
  • 21
    Castro, K. A. D. F.; Halma, M.; Machado, G. S.; Ricci, G. P.; Ucoski, G. M.; Ciuffi, K. J.; Nakagaki, S.; J. Braz. Chem. Soc. 2010, 21, 1329.
  • 22
    Zhou, H.; Jiang, Z.; Wei, S.; Appl. Clay Sci. 2018, 153, 29.
  • 23
    Wang, X.; Wu, P.; Lu, Y.; Huang, Z.; Zhu, N.; Lin, C.; Dang, Z.; Sep. Purif. Technol. 2014, 132, 195.
  • 24
    Wang, H.; Zhang, Z.; Jing, M.; Tang, S.; Wu, Y.; Liu, W.; Appl. Clay Sci. 2020, 186, 105433.
  • 25
    Bini, M.; Monteforte, F.; Quinzeni, I.; Friuli, V.; Maggi, L.; Bruni, G.; J. Solid State Chem. 2019, 272, 131.
  • 26
    Yasaei, M.; Khakbiz, M.; Zamanian, A.; Ghasemi, E.; Mater. Sci. Eng., C 2019, 103, 109816.
  • 27
    Yasaei, M.; Khakbiz, M.; Ghasemi, E.; Zamania, A.; Appl. Surf. Sci. 2019, 467-468, 782.
  • 28
    Gómez, N. A. G.; Silva, G. M.; Wilhelm, H. M.; Wypych, F.; J. Braz. Chem. Soc. 2020, 31, 971.
  • 29
    Jaerger, S.; Wypych, F.; J. Appl. Polym. Sci. 2019, 48737.
  • 30
    Suzuki, J.; Ito, M.; Sugiura, T.; J. Jpn. Assoc. Mineral., Petrol. Econ. Geol. 1976, 71, 183.
  • 31
    Sarp, H.; Perroud, P.; N. Jb. Miner. Mh., Jg. 1991, 11, 481.
  • 32
    Cuchet, S.; Schweiz. Mineral. Petrogr. Mitt. 1995, 75, 283.
  • 33
    Frost, R. L.; Keeffe, E. C.; Spectrochim. Acta, Part A 2011, 81, 111.
  • 34
    Mills, S. J.; Christy, A. G.; Schnyder, C.; Favreau, G.; Price, J. R.; Mineral. Mag. 2014, 78, 1527.
  • 35
    Mills, S. J.; Christy, A. G.; Colombo, F.; Price, J. R.; Mineral. Mag. 2015, 79, 321.
  • 36
    Mills, S. J.; Christy, A. G.; Favreaud, G.; Galea-Cloluse, V.; Acta Crystallogr., Sect. B 2017, 73, 950.
  • 37
    Ventruti, G.; Mugnaioli, E.; Capitani, G.; Scordari, F.; Pinto, D.; Lausi, A.; Phys. Chem. Miner. 2015, 42, 651.
  • 38
    Boclair, J. W.; Braterman, P. S.; Chem. Mater. 1998, 10, 2050.
  • 39
    Layrac, G.; Harrisson, S.; Destarac, M.; Gérardin, C.; Tichit, D.; Appl. Clay Sci. 2020, 193, 105673.
  • 40
    Fan, G.; Li, F.; Evans, D. G.; Duan, X.; Chem. Soc. Rev. 2014, 43, 7040.
  • 41
    Yan, K.; Liu, Y.; Lu, Y.; Chai, J.; Sun, L.; Catal. Sci. Technol. 2017, 7, 1622.
  • 42
    Li, J.; Zhang, S.; Chen, Y.; Liu, T.; Liu, C.; Zhang, X.; Yi, M.; Chu, Z.; Han, X.; RSC Adv. 2017, 7, 29051.
  • 43
    Britto, S.; Radha, A.; Ravishankar, N.; Kamath, P. V.; Solid State Sci. 2007, 9, 279.
  • 44
    Sotiles, A. R.; Wypych, F.; Chem. Commun. 2019, 55, 7824.
  • 45
    Krivovichev, S. V.; Yakovenchuk, V. N.; Zolotarev, A. A.; Ivanyuk, G. N.; Pakhomovski, Y. A.; Chimia 2010, 64, 730.
  • 46
    Merlino, S.; Orlandi, P.; Am. Mineral. 2001, 86, 1293.
  • 47
    Zhu, Y.; Rong, J.; Zhang, T.; Xu, J.; Dai, Y.; Qiu, F.; ACS Appl. Nano Mater. 2018, 1, 284.
  • 48
    Nickel, E.; Mineral. Mag. 1976, 43, 644.
  • 49
    Witzke, T.; Neues Jahrb. Mineral., Monatsh. 1999, 2, 75.
  • 50
    Khaldi, M.; de Roy, A.; Chaouch, M.; Besse J. P.; J. Solid State Chem. 1997, 130, 66.
  • 51
    Frost, R. L.; Kloprogge, J. T.; Spectrochim. Acta, Part A 1999, 55, 2195.
  • 52
    Frost, R. L.; Weier, M. L.; Clissold, M. E.; Williams, P. A.; Spectrochim. Acta, Part A 2003, 59, 3313.
  • 53
    Badreddine, M.; Khaldi, M.; Legrouri, A.; Barroug, A.; Chaouch, M.; de Roy, A.; Besse, J. P.; Mater. Chem. Phys. 1998, 52, 235.
  • 54
    Badreddine, M.; Legrouri, A.; Barroug, A.; de Roy, A.; Besse, J. P.; Mater. Lett. 1999, 38, 391.
  • 55
    Mahjoubi, F. Z.; Khalid, A.; Abdennouri, M.; Barka, N.; J. Taibah Univ. Sci. 2017, 11, 90.
  • 56
    Faramawy, S.; Zaki, T.; Sakr, A. A. E.; Saber, O.; Aboul-Gheit, A. K.; Hassan, S. A.; J. Nat. Gas Sci. Eng. 2018, 54, 72.
  • 57
    Hager, S. L.; Leverett, P.; Williams, P. A.; Can. Mineral. 2009, 47, 635.
  • 58
    Park, Y.; Kuroda, K.; Kato, C.; Solid State Ionics 1990, 42, 197.
  • 59
    Jayanthi, K.; Kamath, P. V.; Periyasamy, G.; Eur. J. Inorg. Chem. 2017, 3675.
  • 60
    Radha, S.; Antonyraj, C. A.; Kamath, P. V.; Kannan, S.; Z. Anorg. Allg. Chem. 2010, 636, 2658.
  • 61
    Sotiles, A. R.; Gomez, N. A. G.; da Silva, S. C.; Wypych, F.; J. Braz. Chem. Soc. 2019, 30, 1807.
  • 62
    Huminicki, D. M. C.; Hawthorne, F. C.; Grice, J. D.; Roberts, A. C.; Jambor, J. L.; Mineral. Rec. 2003, 34, 155.
  • 63
    Wachowiak, J.; Pieczka, A.; Mineral. Mag. 2016, 80, 277.
  • 64
    Zamarreño, I.; Plana, F.; Vazquez, A.; Clague, D. A.; Am. Mineral. 1989, 74, 1054.
  • 65
    Kang, H.; Leoni, M.; He, H.; Huang, G.; Yang, X.; Eur. J. Inorg. Chem. 2012, 3859.
  • 66
    Berner, S.; Araya, O.; Govan, J.; Palza, H.; J. Ind. Eng. Chem. 2018, 59, 134.
  • 67
    Kloprogge, J. T.; Wharton, D.; Hickey, L.; Frost, R. L.; Am. Mineral. 2002, 87, 623.
  • 68
    Kloprogge, J. T.; Hickey, L.; Frost, R. L.; Mater. Chem. Phys. 2005, 89, 99.
  • 69
    Liu, J.; Song, J.; Xiao, H.; Zhang, L.; Qin, Y.; Liu, D.; Hou, W.; Du, N.; Powder Technol. 2014, 253, 41.
  • 70
    Karami, Z.; Jouyandeh, M.; Hamad, S. M.; Ganjali, M. R.; Aghazadeh, M.; Torre, L.; Puglia, D.; Saeb, M. R.; Prog. Org. Coat. 2019, 136, 105278.
  • 71
    Kameda, T.; Fubasami, Y.; Yoshioka, T.; J. Therm. Anal. Calorim. 2012, 110, 641.

Publication Dates

  • Publication in this collection
    20 Jan 2021
  • Date of issue
    Jan 2021

History

  • Received
    4 Apr 2020
  • Accepted
    14 Aug 2020
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