Next Article in Journal
Laser Textured Superhydrophobic SiC Ceramic Surface and the Performance of Condensation Heat Transfer
Next Article in Special Issue
Atomic Arrangement, Hydrogen Bonding and Structural Complexity of Alunogen, Al2(SO4)3·17H2O, from Kamchatka Geothermal Field, Russia
Previous Article in Journal
Magnetic Force Microscopy of Multiferroic Bulk Ceramic Oxides
Previous Article in Special Issue
Carbonate Minerals’ Precipitation in the Presence of Background Electrolytes: Sr, Cs, and Li with Different Transporting Anions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Crystal Structure of Mg–Al–CO3 Layered Double Hydroxide

by
Elena S. Zhitova
1,*,
Rezeda M. Sheveleva
1,2,
Andrey A. Zolotarev
1,2 and
Sergey V. Krivovichev
2,3
1
Institute of Volcanology and Seismology, Far East Branch of Russian Academy of Sciences, Petropavlovsk-Kamchatsky 683006, Russia
2
Saint-Petersburg State University, Saint-Petersburg 199034, Russia
3
Kola Science Centre, Apatity 189209, Russia
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(5), 839; https://doi.org/10.3390/cryst13050839
Submission received: 3 May 2023 / Revised: 16 May 2023 / Accepted: 16 May 2023 / Published: 19 May 2023
(This article belongs to the Special Issue Mineralogical Crystallography (3rd Edition))

Abstract

:
The crystal structure of quintinite, Mg4Al2(OH)12(CO3)·3H2O, from the Jacupiranga alkaline complex (Cajati, São Paulo, Brazil), was refined for two samples (91002 and C7029) using single-crystal X-ray diffraction data. The mineral crystallizes in the P-3c1 space group, a = 5.246/5.298, c = 15.110/15.199 Å for samples 91002/C7029. The crystal structure consists of octahedral sheets with Mg and Al ordering according to a 3 × 3 superstructure. The Mg and Al atoms are coordinated by six hydroxylated oxygen atoms; the average <Mg–O> and <Al–O> bond distances are in the ranges 2.022–2.053 Å and 1.974–1.978 Å, respectively. The interlayer structures are identical (in contradiction to the previous assumptions), and consist of disordered (CO3)2− groups and (H2O)0 molecules. The samples from Jacupiranga can be identified as quintinite-2T, which is the second finding of this polytype after the Kovdor alkaline complex (Kola peninsula, Russia). The powder X-ray diffraction pattern of quintinite-2T contains weak superstructure reflection at 4.57 Å (010), indicative of Mg and Al ordering. An important crystal-chemical criterion of quintinite is the interlayer distance (d00n-value) of ~7.56 Å, which is steady among natural specimens from various findings worldwide.

1. Introduction

Hydrotalcite supergroup minerals [1] are natural representatives of industrially applied Layered Double Hydroxides (LDHs) [2]. LDHs form a class of inorganic lamellar compounds with crystal structures consisting of alternating positively charged metal-hydroxide layers (octahedral sheets) and negatively charged interlayers [3,4]. LDHs can be represented by the general formula [5,6,7] [M2+1−xM3+x(OH)2][An−]x/nmH2O, where M2+ = Mg2+, Ni2+, Fe2+, Mn2+, etc., and M3+ may be, e.g., Al3+, Fe3+, Cr3+ and other trivalent cations; x = 0.33 and 0.25 are the most common values and correspond to M2+:M3+ = 2:1 and 3:1 ratios, respectively, which are the most widespread among LDHs; An− is the anion of the n negative charge (the most common among mineral are CO32−, Cl, SO42−, Sb(OH)6); m is the number of interlayer H2O molecules. This general formula is further extended by (i) minerals of the wermlandite group [8,9,10,11] and their synthetic counterparts [12,13,14] that contain additional interlayer mono- or divalent cations (like Na, K, Ca etc.), and by (ii) Li–Al (i.e., M+M3+) LDHs, the structures of which are represented by gibbsite-like layers [15,16], with the general formula of such compounds given as [Al4Li2(OH)2][An−]x/nmH2O.
Magnesium–aluminum LDHs with carbonate interlayers are among the longest known in the LDH family; they are applied as catalytic materials [17,18] and considered efficient and inexpensive oxygen evolution reaction electrocatalysts [19]. Mg–Al LDHs are often used as model systems for LDH families of different (rare) chemical compositions (for example, the substitution of cations while maintaining stoichiometry). Mg–Al–CO3 LDHs are applied in pharmacology as a Talcid medicine. Recent pharmaceutical studies involving Mg–Al LDHs include hybrid materials with glibenclamide [20], nanocarrier materials for antimicrobial chemotherapy [21], and a porous nanocomposite of LDH with chitosan for cosmetic application [22]. LDHs, including Mg–Al varieties, are considered sorbents of undesirable anions and cations; recent research includes methyl orange [23,24], PbII [25], CuII [26], and Congo red dye [27] removal from aqueous solutions, as well as harmful anion (such as Cl, Br, I, AsO43−, etc.) sorption from fertile soil [28]. The uptake of Cl and CO32− anions by Mg–Al and Ca–Al LDHs from pore solutions was simulated for cementitious materials [29]. The possibility of anion uptake by LDHs also makes them advanced materials for CO2 capture [30]. The high industrial and material science interest towards LDHs requires a detailed understanding of the internal structure of these materials. Highly crystalline and extremely stable prototypes of such materials are natural compounds [31]—hydrotalcite supergroup minerals [1].
Despite the wide chemical variability of LDHs, their crystal structures show a certain structural commonality, in which one structural type extends to compounds of different compositions. For example, quintinite is isotypic to the Li–Al gibbsite-based mineral akopovaite, Al4Li2(OH)12CO3∙3H2O [32], following the substitution scheme [Mg4Al2(OH)12]2+ → [Al4Li2(OH)12]2+. Both structures show cation (Mg/Al or Al/Li) ordering according to 3 × 3 superstructures and identical schemes of interlayer species arrangement despite a rather principal chemical variability. The same applies to their Cl-analogues: chlormagaluminite, Mg4Al2(OH)2Cl2∙3H2O, [33], and dritsite, Al4Li2(OH)12Cl2∙3H2O [34]. Comparison of the crystal structures of minerals shows that they are isotypic to their synthetic analogues, which is well shown by Li–Al LDHs [35,36,37]. Therefore, the crystal structures obtained for minerals through single-crystal X-ray diffraction analysis (the technique that makes it possible to obtain models of crystal structures with greater accuracy) can be extended to industrially applied synthetic LDH materials.
In this study, we provide crystal chemical characterization of quintinite from Jacupiranga (São Paulo, Brazil). Previously, the crystal structure of quintinite from that locality was described as stratified with two types of interlayers (see details below), which contradicts the more recent studies of quintinite from the Kovdor complex [38,39,40,41,42]. The goal of the present study is to resolve this contradiction, i.e., to check whether quintinites from different localities are isotypic to each other or not. In addition, we provide data that should be helpful for the identification of cation-ordered polytypes with hexagonal stacking sequences using a powder X-ray diffraction method.

Previous Crystal Structure Studies of Quintinite

The first crystal structure study of quintinite (at that time—“manasseite”, i.e., hydrotalcite) was carried out on a sample from Jacupiranga [43]. The study demonstrated the presence of Mg and Al ordering according to a 3 × 3 superstructure in the metal-hydroxide layer (or octahedral sheet), hexagonal layer stacking sequence and stratified interlayer that consisted of two types: (i) (H2O) molecules and (ii) (CO3)2−, which alternated along the z direction [43] (Figure 1).
Almost simultaneously, quintinite was approved as a separate (from hydrotalcite/manasseite) mineral species on samples from the alkaline complex of Mont Saint-Hilaire, Quebec, Canada [44]. The crystal structure was not refined, but unit-cell parameters determined were indicative of polytypes with hexagonal and rhombohedral layer stacking sequences doubled in comparison to previous data [43] with lattice parameter a according to a 2 3 × 2 3 superstructure (which, however, has not been identified elsewhere else) (Figure 2).
We studied the polytypism of quintinite from the Kovdor alkaline complex (Kola peninsula, Russia), where the polytype character depends upon the following factors: (i) disorder or order of cations within a metal-hydroxide layer and (ii) type of layer stacking [38,39,40,41,42]. In total, five polytypes of quintinite were described in Kovdor; among them, 2H, 2T, 3R and 1M polytypes appear commonly (Figure 3). In addition, it was shown that the sequence of polytype formation can be described as 2H → 2T → 1M. The polytype 2T-3c (or 2H-3c in the original version) can be characterized as a cell of 2T polytype shifted along the z direction, producing tripling of the number of layers within the unit cell; this polytype was observed in only one sample. Later, the study of “hydrotalcite” from two localities in the Ural Mountains (Russia), Ural emerald mines (Malyshevskoe or former Mariinskoe deposit) [45] and Bazhenovskoe chrysotile–asbestos deposit [46], demonstrated that Ural samples are in fact quintinite-1M (previously identified as hydrotalcite) isotypic to the Kovdor quintinite-1M. Recently, quintinite-1M was described from Mount Mather Creek, British Columbia, together with the 3T(?) polytype, which possesses some residual reflections [47].

2. Materials and Methods

2.1. Materials

Two samples from the Jacupiranga alkaline complex (Cajati, São Paulo, Brazil) labelled as “manasseite” were investigated from the following sources: (a) the systematic collection of the Fersman Mineralogical Museum of the Russian Academy of Science (Moscow, Russia) stored under catalogue number 91002 and (b) the collection of the Smithsonian Institution (National Museum of Natural History in Washington, DC, USA) stored under catalogue number C7029.
Originally, quintinite was described in Jacupiranga as “manasseite” (a hexagonal form of hydrotalcite that is currently discredited) [48]. Quintinite and manasseite/hydrotalcite have a confusing description history {see hydrotalcite’s and quintinite’s descriptions in Snarum (Norway) by [49,50,51]} because both are Mg–Al–CO3 minerals with different Mg:Al ratios: 2:1 (quintinite) and 3:1 (hydrotalcite and formerly “manasseite”). The crystal structure of quintinite from the Jacupiranga mine was studied by [43], and the mineral was stated to be a potentially new mineral species from the “hydrotalcite-manasseite group” that had previously been described as “manasseite”. Paying attention to the synchronism of this work on structure refinement (1996) [43] with the description of a new mineral—quintinite (1997) [44] and the subsequent citation of this structural work in the first description of quintinite, it is clear that the mineral from Jacupiranga studied by [43] was in fact quintinite. At the time of the description of its crystal structure, the name of quintinite had not yet been approved by the Commission on New Minerals, Nomenclature and Classification, and therefore is not mentioned in the text. The mineral is described as quintinite in the most recent study of the Jacupiranga minerals [52]. In Jacupiranga, quintinite was found in carbonatite rocks [48] and is considered a late-stage hydrothermal mineral.

2.2. Single-Crystal X-ray Diffraction

Single-crystal X-ray diffraction analysis was carried out for quintinite samples 91002 and C7029 using a Bruker Smart Apex diffractometer (X-ray diffraction Resource Center, St. Petersburg State University), MoKα radiation, operated at 50 kV and 40 mA, equipped with a CCD area detector; more details on data collection are given in Table 1. The unit-cell parameters were refined using least-square methods. The data were processed using the P-3c1 space group. During data processing, we first successfully processed single-crystal X-ray diffraction data and refined the crystal structure in the P63/mcm space group as the most probable option suggested by CrysAlisPro software package [53] (this space group was used for the refinement of chlormagaluminite structure [33]). However, this led to a large number of unaccounted-for reflections, as a result of which we decided to use the P-3c1 space group previously used for quintinite [42].
The structure was solved and refined using the ShelX program package [54] incorporated into the Olex2 software shell [55], to R1 = 0.053/0.066 for 320/424 independent reflections with I ≥ 2σ(I) for samples 91002 and C7029.
The positions of Mg, Al and oxygen in the metal-hydroxide layer were located by direct methods and refined anisotropically (Table S2). The interlayer species (O and C) and hydrogen atoms in the metal-hydroxide layer were located using the difference-Fourier maps and refined isotropically. The Mg, Al and O site occupancies (in the metal-hydroxide layer) were found to be close to 100% and fixed. Site occupancies of the interlayer atoms were refined for O atoms and fixed for C atoms. The main crystallographic characteristics and structure refinement parameters are listed in Table 1.

2.3. Powder X-ray Diffraction

Powder X-ray diffraction (XRD) data were collected for sample 91002 using a Rigaku R-Axis Rapid II single-crystal diffractometer using Debye–Scherrer geometry (d = 127.4 mm). The diffractometer was equipped with a rotating anode (CoKα, λ = 1.79026, voltage = 40 kV and current = 15 mA), microfocus optics and a cylindrical image plate detector. The data were converted using the Osc2xrd program [56].
The unit-cell parameters were refined via the Pawley method using Topas 4.2 [57], with the hexagonal structure model of space group P-3c1 and with the starting unit-cell parameters of sample 91002 reported herein. Refinement was based on the reflections in the 2θ region 10–60°. The indexing of the XRD pattern and refinement of the unit-cell parameters were performed with fixed atom coordinates, site scattering and isotropic displacement parameters. Neutral scattering factors were used for all atoms. The background was modelled using a Chebyshev polynomial approximation of the 14th order. The peak profile was described using the fundamental parameters approach. Refinement of preferred orientation parameters confirmed the presence of a significantly preferred orientation along the [001] direction.

3. Results

3.1. Crystal Structure Solution and Refinement

The collected single-crystal X-ray diffraction data were processed in the P-3c1 space group, a = 5.246/5.298 Å, c = 15.110/15.199 Å for samples 91002/C7029. Eight atomic positions were located from the structure solution and refinement for each quintinite sample (sites labelled): Mg, Al, O1 and H1 are part of the metal-hydroxide layer (or octahedral sheet) and fully occupied (Table 2).
The Mg and Al sites are ordered according to a 3 × 3 superstructure, and each of the metal sites is coordinated by six protonated oxygen atoms. The average <Mg–O> and <Al–O> bond distances are in the ranges 2.022–2.053 Å and 1.974–1.978 Å, respectively (Table 3). The other four sites, C1, C2, O2 and O3, are at the interlayer level and correspond to the low-occupied sites of statistically disordered (CO3)2− groups and (H2O)0 molecules. All interlayers are identical and consist of both (CO3)2− groups and (H2O)0 molecules in one layer.
The O1–H1 bond is nearly perpendicular to the plane of the octahedral sheet; the bond distance is 0.82/0.91 Å for samples 91002/C7029 (Table 4). The O1 atom acts as a donor (D), while the interlayer O2 and O3 atoms are acceptors (A) of hydrogen bonds. The HA bond distance is almost two times longer than the DH distance (Table 4).

3.2. Powder X-ray Diffraction

The main set of reflections of the powder X-ray diffraction pattern (Figure 4) correspond to the classic (for LDHs) 2H polytype (Table S1). Low-intensity superstructure reflection indicative of Mg and Al ordering according to a 3 × 3 superstructure is observed at d = 4.57 Å and indexed (hkl) as 010. Another characteristic feature of the quintinite powder X-ray diffraction pattern is the d-value of 7.56 Å, which agrees well with the previous crystal-chemical studies of quintinite [58].

4. Discussion

In this work, using X-ray diffraction analysis, it is shown that quintinite from Jacupiranga is characterized by a hexagonal type of layer stacking, a superstructure based on Mg and Al ordering in a 3 × 3 pattern, and identical interlayers composed of disordered (H2O)0 molecules and (CO3)2− groups at the same level in one gallery. This polytype is designated 2T according to the Ramsdell [59] notation. Topologically, quintinite-2T from Jacupiranga is identical to quintinite-2T from the Kovdor alkaline complex [42] and is the second confirmed finding of quintinite-2T in the world. It should be noted that, due to the high degree of disorder in the interlayer components, the positions of localized interlayer C and O atoms may differ from sample to sample. Despite that, the electron density distribution maps at the interlayer level are identical. The same is true of the space group selection, i.e., potential polytypes with structures refined in space groups P-3c1 or P63/mcm with identical crystal-chemical features should not be taken as separate polytypes (see details in Section 2.2).
Previously, we assumed [42] that the structure obtained for quintinite from Jacupiranga [43] with stratified interlayers is not entirely correct from a crystal-chemical point of view, since, probably, the “carbonate” and “water” interlayers should have had different heights of the interlayer gallery, which was not observed. It should also be noted that such a division into two types of interlayers in the previous work was most likely because not all electron density peaks at the interlayer level were detected as their intensity could be very low, on the order of 1 ē. A similar problem is typical for earlier refinements of many LDH minerals and is a consequence of the (apparently) insufficient accuracy of equipment at that time. Although we had no opportunity to examine the same quintinite crystal as earlier researchers [43], we suggest that quintinite does not form structures with stratified interlayers. We also suggest that quintinite is characterized by four main polytypes (Figure 3), one of which is 2T, which is described in this work. We believe that this is a rather important assumption, since experimentally obtained crystal structures often become the basis for Rietveld refinements or computer simulation. Unfortunately, sometimes in the literature, it is possible to observe crystal structures that are unrealistic from a crystal-chemical point of view, in which, for example, H2O molecules and carbonate groups are located in the same interlayer gallery at different levels.
The presence of M2+ and M3+ cation ordering within metal-hydroxide layers that form a 3 × 3 superstructure is manifested in the powder X-ray diffraction pattern by the presence of a superstructure reflection with d ~ 4.57 Å (indexed as 010 in the case of 2T polytype). The same superstructure reflection was observed for chlormagaluminite [33]. The powder X-ray diffraction patterns for both minerals, quintinite and chlormagaluminite, were recorded using the Gandolfi technique, reducing the effect of the sample’s preferred orientation along [001]. The presence of such a reflection is indicative of M2+ and M3+ cation ordering that should further be combined with the layer stacking sequence. However, the absence of this reflection, especially considering its low intensity (i.e., difficulty of detection), should not be taken as an indication of the absence of cation ordering.
An indirect criterion for the topological identity of quintinites from different natural environments and places can be the value of the interlayer distance (d-value) of quintinite, which is ~7.56 Å (Table 5) (or slightly above if the Mg:Al ratio is higher than the ideal of 2:1 [58]). Quintinite findings from different localities around the globe and different genetic types are characterized by a steady d00n-value, despite some differences in the physicochemical and thermodynamic conditions of the formation of the minerals (Table 5). Thus, we believe that the internal structure of the mineral indicates certain principles for the construction of metal-hydroxyl layers and interlayers, as well as their interactions within the crystal structure, which can be transferred to artificially obtained compounds. In particular, we emphasize that, under room conditions, we do not observe fundamental changes in the content of interlayer H2O molecules and related structural transformations of quintinite.
Finally, we would like to emphasize that, according to the available structural data for quintinites from different findings and differing in geological position, that is, different conditions of formation, the crystal structures are built in the same way: (i) the structure of the interlayers is the same, (ii) the structure of the octahedral sheet differs only in the recorded ordering of Mg and Al or absence of recorded ordering, i.e., “disordered” (although disorder may be apparent due to the incorporation of impurities) and (iii) different types of stacking of layers.

5. Conclusions

In this work, we refined the crystal structure of Jacupiranga quintinite and showed that, from a structural point of view, it is identical to the 2T polytype of Kovdor quintinite and consists of metal hydroxide layers [Mg4Al2(OH)12]2+ and [(CO3)(H2O)3]2− interlayers. The Mg and Al ordering according to a 3 × 3 superstructure is detected via single-crystal X-ray diffraction and powder X-ray diffraction through the appearance of additional (to 2H pattern) low-intensity superstructure reflections, including the reflection with d010~4.57 Å in the powder pattern. Combining our data and data from the literature, as well as information on the d00n-values of quintinites (d00n~7.56 Å) of various genetic types and localities, we conclude that the structure of quintinite obeys certain crystal chemical principles and is sustained in the light of layer–interlayer interactions and construction of octahedral sheets and interlayers. We consider it highly probable that quintinite polytypes already known from the Kovdor alkaline complex can be found in other localities, as shown in the example of quintinite from the Jacupiranga alkaline complex. In addition, we also hope that a more detailed history of the description of quintinite from Jacupiranga will lead to a correct definition of this mineral, which will help streamline the literature data on the description of hydrotalcite and quintinite.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst13050839/s1, Table S1: Powder X-ray diffraction data for quintinite sample 91002; Table S2: Anisotropic displacement parameters (Å2) for quintinite samples 91002 and C7029.

Author Contributions

Conceptualization, E.S.Z. and S.V.K.; methodology, E.S.Z., A.A.Z. and S.V.K.; software, E.S.Z., R.M.S., A.A.Z. and S.V.K.; validation, E.S.Z., R.M.S., A.A.Z. and S.V.K.; formal analysis, E.S.Z. and R.M.S.; investigation, E.S.Z., R.M.S., A.A.Z. and S.V.K.; resources, S.V.K.; data curation, E.S.Z. and A.A.Z.; writing—original draft preparation, E.S.Z., R.M.S. and S.V.K.; writing—review and editing, E.S.Z., R.M.S., A.A.Z. and S.V.K.; visualization, E.S.Z.; project administration, E.S.Z.; funding acquisition, E.S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation, grant number 22-77-10036.

Data Availability Statement

Crystallographic information files (cif) have been deposited via the joint Cambridge Crystal Data Centre CCDC/FIZ Karlsruhe deposition service; the deposition numbers are CSD 2260275 and 2260276 for samples 91002 and C7029, respectively.

Acknowledgments

The technical support of St. Petersburg State University Resource Centers “XRD” is gratefully acknowledged. We thank the Fersman Mineralogical Museum (Dmitry Belakovsky) and the Smithsonian National Museum of Natural History (Jeffrey Post) for providing the samples for research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mills, S.J.; Christy, A.G.; Génin, J.M.; Kameda, T.; Colombo, F. Nomenclature of the hydrotalcite supergroup: Natural layered double hydroxides. Mineral. Mag. 2012, 76, 1289–1336. [Google Scholar] [CrossRef]
  2. Rives, V. Layered Double Hydroxides: Present and Future; Nova Science Publishers: New York, NY, USA, 2001; pp. 1–439. [Google Scholar]
  3. Duan, X.; Evans, D.G. Layered Double Hydroxides; Springer: Berlin, Germany, 2006; pp. 1–87. [Google Scholar] [CrossRef]
  4. Aminoff, G.; Broome, B. Contributions to the mineralogy of Långban. III. Contributions to the knowledge of the mineral pyroaurite. K. Sven. Vetensk. Akad. Handl. 1931, 9, 23–48. [Google Scholar]
  5. Bukhtiyarova, M.V. A review on effect of synthesis conditions on the formation of layered double hydroxides. J. Solid State Chem. 2019, 269, 494–506. [Google Scholar] [CrossRef]
  6. Yu, J.; Wang, Q.; O’Hare, D.; Sun, L. Preparation of two dimensional layered double hydroxide nanosheets and their applications. Chem. Soc. Rev. 2017, 46, 5950–5974. [Google Scholar] [CrossRef] [PubMed]
  7. Kameliya, J.; Verma, A.; Dutta, P.; Arora, C.; Vyas, S.; Varma, R.S. Layered Double Hydroxide Materials: A Review on Their Preparation, Characterization, and Applications. Inorganics 2023, 11, 121. [Google Scholar] [CrossRef]
  8. Rius, J.; Allmann, R. The superstructure of the double layer mineral wermlandite [Mg7(Al0.57,Fe3+0.43)2(OH)18]2+[(Ca0.6,Mg0.4)(SO4)2(H2O)12]2−. Z. fur Krist. 1984, 168, 133–144. [Google Scholar] [CrossRef]
  9. Zhitova, E.S.; Chukanov, N.V.; Jonsson, E.; Pekov, I.V.; Belakovskiy, D.I.; Vigasina, M.F.; Zubkova, N.V.; Van, K.V.; Britvin, S.N. Erssonite, CaMg7Fe3+2(OH)18(SO4)2⋅12H2O, a new hydrotalcite-supergroup mineral from Långban, Sweden. Mineral. Mag. 2021, 85, 817–826. [Google Scholar] [CrossRef]
  10. Huminicki, D.M.C.; Hawthorne, F.C. The crystal structure of nikischerite, NaFe62+Al3(SO4)2(OH)18(H2O)12, a mineral of the shigaite group. Can. Mineral. 2003, 41, 79–82. [Google Scholar] [CrossRef]
  11. Cooper, M.A.; Hawthorne, F.C. The crystal structure of shigaite, [AlMn22+(OH)6]3(SO4)2Na(H2O)6(H2O)6, a hydrotalcite group mineral. Can. Mineral. 1996, 34, 91–97. [Google Scholar]
  12. Sotiles, A.R.; Gomez, N.A.G.; Santos, M.P.; Grassi, M.T.; Wypych, F. Synthesis, characterization, thermal behavior and exchange reactions of new phases of layered double hydroxides with the chemical composition [M+26Al3(OH)18(SO4)2]. (A(H2O)6).6H2O (M+2 = Co, Ni; A = Li+, Na+, K+). Appl. Clay Sci. 2019, 181, 105217. [Google Scholar] [CrossRef]
  13. Sotiles, A.R.; Baika, L.M.; Grassi, M.T.; Wypych, F. Cation exchange reactions in layered double hydroxides intercalated with sulfate and alkali cations (A(H2O)6)[M+26Al3(OH)18(SO4)2].6H2O (M+2 = Mn, Mg, Zn, A+ = Li, Na, K). J. Am. Chem. Soc. 2019, 141, 531–540. [Google Scholar] [CrossRef] [PubMed]
  14. Sotiles, A.R.; Wypych, F. Synthesis and topotactic exchange reactions of new layered double hydroxides intercalated with ammonium/sulfate. Solid State Sci. 2020, 106, 106304. [Google Scholar] [CrossRef]
  15. Serna, C.J.; Rendon, J.L.; Iglesias, J.E. Crystal-chemical study of layered [Al2Li(OH)6]+X·nH2O. Clay Clay Miner. 1982, 30, 180–184. [Google Scholar] [CrossRef]
  16. Sissoko, I.; Iyagba, E.T.; Sahai, I.; Biloen, P. Anion intercalation and exchange in Al(OH)3– derived compounds. J. Solid State Chem. 1985, 60, 283–288. [Google Scholar] [CrossRef]
  17. Karcz, R.; Napruszewska, B.D.; Walczyk, A.; Kryściak-Czerwenka, J.; Duraczyńska, D.; Płaziński, W.; Serwicka, E.M. Comparative Physicochemical and Catalytic Study of Nanocrystalline Mg-Al Hydrotalcites Precipitated with Inorganic and Organic Bases. Nanomaterials 2022, 12, 2775. [Google Scholar] [CrossRef] [PubMed]
  18. Karcz, R.; Napruszewska, B.D.; Michalik, A.; Kryściak-Czerwenka, J.; Duraczyńska, D.; Serwicka, E.M. Fine Crystalline Mg-Al Hydrotalcites as Catalysts for Baeyer-Villiger Oxidation of Cyclohexanone with H2O2. Catalysts 2021, 11, 1493. [Google Scholar] [CrossRef]
  19. Yin, X.; Hua, Y.; Gao, Z. Two-Dimensional Materials for High-Performance Oxygen Evolution Reaction: Fundamentals, Recent Progress, and Improving Strategies. Renewables 2023, 1, 190–226. [Google Scholar] [CrossRef]
  20. Leão, A.D.; Oliveira, V.V.; Marinho, F.A.; Wanderley, A.G.; Aguiar, J.S.; Silva, T.G.; Soare, M.F.R.; Soares-Sobrinho, J.L. Hybrid systems of glibenclamide and layered double hydroxides for solubility enhancement for the treatment of diabetes mellitus II. Appl. Clay Sci. 2019, 181, 105218. [Google Scholar] [CrossRef]
  21. Francius, G.; André, E.; Soulé, S.; Merlin, C.; Carteret, C. Layered Double Hydroxides (LDH) as nanocarriers for antimicrobial chemotherapy: From formulation to targeted applications. Mater. Chem. Phys. 2023, 293, 126965. [Google Scholar] [CrossRef]
  22. Kim, J.; Kim, T.H.; Lee, J.H.; Park, Y.A.; Kang, Y.J.; Ji, H.G. Porous nanocomposite of layered double hydroxide nanosheet and chitosan biopolymer for cosmetic carrier application. Appl. Clay Sci. 2021, 205, 106067. [Google Scholar] [CrossRef]
  23. Zaghloul, A.; Benhiti, R.; Ichou, A.A.; Carja, G.; Soudani, A.; Zerbet, M.; Sinan, F.; Chiban, M. Characterization and application of MgAl layered double hydroxide for methyl orange removal from aqueous solution. Mater. Today 2021, 37, 3793–3797. [Google Scholar] [CrossRef]
  24. Grégoire, B.; Bantignies, J.L.; Le-Parc, R.; Prélot, B.; Zajac, J.; Layrac, G.; Tichit, D.; Martin-Gassin, G. Multiscale mechanistic study of the adsorption of methyl orange on the external surface of layered double hydroxide. J. Phys. Chem. C 2019, 123, 22212–22220. [Google Scholar] [CrossRef]
  25. Zhao, D.; Sheng, G.; Hu, J.; Chen, C.; Wang, X. The adsorption of Pb (II) on Mg2Al layered double hydroxide. Chem. Eng. J. 2011, 171, 167–174. [Google Scholar] [CrossRef]
  26. Yang, X.; Kameda, T.; Saito, Y.; Kumagai, S.; Yoshioka, T. Investigation of the mechanism of Cu (II) removal using Mg-Al layered double hydroxide intercalated with carbonate: Equilibrium and pH studies and solid-state analyses. Inorg. Chem. Commun. 2021, 132, 108839. [Google Scholar] [CrossRef]
  27. Nestroinaia, O.V.; Ryltsova, I.G.; Yaprintsev, M.N.; Nakisko, E.Y.; Seliverstov, E.S.; Lebedeva, O.E. Sorption of Congo Red anionic dye on natural hydrotalcite and stichtite: Kinetics and equilibrium. Clay Miner. 2022, 57, 105–113. [Google Scholar] [CrossRef]
  28. Guo, J.Y.; Zhang, W.; Zhao, X.J.; Jie, Y.; Song, H.T.; Xu, X.Y.; Lu, H.; Yan, H. Theoretical Study on the Mechanism of Removing Harmful Anions from Soil via Mg2al Layered Double Hydroxide as Stabilizer. SSRN 2022. [Google Scholar] [CrossRef]
  29. Ke, X.; Bernal, S.A.; Provis, J.L. Uptake of chloride and carbonate by Mg-Al and Ca-Al layered double hydroxides in simulated pore solutions of alkali-activated slag cement. Cem. Concr. Res. 2017, 100, 1–13. [Google Scholar] [CrossRef]
  30. Sakai, M.; Imagawa, H.; Baba, N. Layered-double-hydroxide-based Ni catalyst for CO2 capture and methanation. Appl. Catal. A-Gen. 2022, 647, 118904. [Google Scholar] [CrossRef]
  31. Zhitova, E.S.; Sheveleva, R.M.; Kasatkin, A.V.; Zolotarev, A.A.; Bocharov, V.N.; Kupchinenko, A.N.; Belakovsky, D.I. Crystal structure of hydrotalcite group mineral—Desautelsite, Mg6MnIII2(OH)16(CO3)·4H2O, and relationship between cation size and in-plane unit cell parameter. Symmetry 2023, 15, 1029. [Google Scholar] [CrossRef]
  32. Karpenko, V.Y.; Zhitova, E.S.; Pautov, L.A.; Agakhanov, A.A.; Siidra, O.I.; Krzhizhanovskaya, M.G.; Rassulov, V.A.; Bocharov, V.N. Akopovaite, Li2Al4(OH)12(CO3)(H2O)3, a new Li-member of the hydrotalcite supergroup from Turkestan Range, Kyrgyzstan. Mineral. Mag. 2020, 84, 301–311. [Google Scholar] [CrossRef]
  33. Zhitova, E.S.; Krivovichev, S.V.; Pekov, I.V.; Yapaskurt, V.O. Crystal chemistry of chlormagaluminite, Mg4Al2(OH)12Cl2(H2O)2, a natural layered double hydroxide. Minerals 2019, 9, 221. [Google Scholar] [CrossRef]
  34. Zhitova, E.S.; Pekov, I.V.; Chaikovskiy, I.I.; Chirkova, E.P.; Yapaskurt, V.O.; Bychkova, Y.V.; Belakovsky, D.I.; Chukanov, N.V.; Zubkova, N.V.; Krivovichev, S.V.; et al. Dritsite, Li2Al4(OH)12Cl2·3H2O, a new gibbsite-based hydrotalcite supergroup mineral. Minerals 2019, 9, 492. [Google Scholar] [CrossRef]
  35. Britto, S.; Thomas, G.S.; Kamath, P.V.; Kannan, S. Polymorphism and structural disorder in the carbonate containing layered double hydroxide of Li with Al. J. Phys. Chem. 2008, 112, 9510–9515. [Google Scholar] [CrossRef]
  36. Britto, S.; Kamath, P.V. Structure of Bayerite-Based Lithium–Aluminum Layered Double Hydroxides (LDHs): Observation of Monoclinic Symmetry. Inorg. Chem. 2009, 48, 11646–11654. [Google Scholar] [CrossRef]
  37. Britto, S.; Kamath, P.V. Polytypism in the lithium–aluminum layered double hydroxides: The [LiAl2(OH)6]+ layer as a structural synthon. Inorg. Chem. 2011, 50, 5619–5627. [Google Scholar] [CrossRef] [PubMed]
  38. Krivovichev, S.V.; Yakovenchuk, V.N.; Zhitova, E.S.; Zolotarev, A.A.; Pakhomovsky, Y.A.; Ivanyuk, G.Y. Crystal chemistry of natural layered double hydroxides. I. Quintinite-2H-3 c from the Kovdor alkaline massif, Kola peninsula, Russia. Mineral. Mag. 2010, 74, 821–832. [Google Scholar] [CrossRef]
  39. Krivovichev, S.V.; Yakovenchuk, V.N.; Zhitova, E.S.; Zolotarev, A.A.; Pakhomovsky, Y.A.; Ivanyuk, G.Y. Crystal chemistry of natural layered double hydroxides. 2. Quintinite-1M: First evidence of a monoclinic polytype in M2+-M3+ layered double hydroxides. Mineral. Mag. 2010, 74, 833–840. [Google Scholar] [CrossRef]
  40. Krivovichev, S.V.; Yakovenchuk, V.N.; Zhitova, E.S. Natural double layered hydroxides: Structure, chemistry, and information storage capacity. In Minerals as Advanced Materials II; Krivovichev, S.V., Ed.; Springer: Berlin/Heidelberg, Germany, 2012; pp. 87–102. [Google Scholar] [CrossRef]
  41. Zhitova, E.S.; Yakovenchuk, V.N.; Krivovichev, S.V.; Zolotarev, A.A.; Pakhomovsky, Y.A.; Ivanyuk, G.Y. Crystal chemistry of natural layered double hydroxides. 3. The crystal structure of Mg, Al-disordered quintinite-2H. Mineral. Mag. 2010, 74, 841–848. [Google Scholar] [CrossRef]
  42. Zhitova, E.S.; Krivovichev, S.V.; Yakovenchuk, V.N.; Ivanyuk, G.Y.; Pakhomovsky, Y.A.; Mikhailova, J.A. Crystal chemistry of natural layered double hydroxides: 4. Crystal structures and evolution of structural complexity of quintinite polytypes from the Kovdor alkaline-ultrabasic massif, Kola peninsula, Russia. Mineral. Mag. 2018, 82, 329–346. [Google Scholar] [CrossRef]
  43. Arakcheeva, A.V.; Pushcharovskii, D.Y.; Rastsvetaeva, R.K.; Atencio, D.; Lubman, G.U. Crystal structure and comparative crystal chemistry of Al2Mg4(OH)12(CO3)·3H2O, a new mineral from the hydrotalcite-manasseite group. Crystallogr. Rep. 1996, 41, 972–981. [Google Scholar]
  44. Chao, G.Y.; Gault, R.A. Quintinite-2H, quintinite-3T, charmarite-2H, charmarite-3T and caresite-3T, a new group of carbonate minerals related to the hydrotalcite-manasseite group. Can. Mineral. 1997, 35, 1541–1549. [Google Scholar]
  45. Zhitova, E.S.; Popov, M.P.; Krivovichev, S.V.; Zaitsev, A.N.; Vlasenko, N.S. Quintinite-1M from the Mariinsky Deposit, Ural Emerald Mines, Central Urals, Russia. Geol. Ore Depos. 2018, 59, 745–751. [Google Scholar] [CrossRef]
  46. Krivovichev, S.V.; Antonov, A.A.; Zhitova, E.S.; Zolotarev, A.A.; Krivovichev, V.G.; Yakovenchuk, V.N. Quintinite-1M from Bazhenovskoe deposit (Middle Ural, Russia): Crystal structure and properties. Vestnik Saint Petersburg Univ. Earth Sci. 2012, 7, 3–9. (In Russian) [Google Scholar]
  47. Piilonen, P.; Poirier, G.; Rowe, R.; Mitchell, R.; Robak, C. Mount Mather Creek, British Columbia—A new sodalite-bearing carbohydrothermal breccia deposit including a new Canadian occurrence for the rare minerals edingtonite and quintinite. Mineral. Mag. 2022, 86, 282–306. [Google Scholar] [CrossRef]
  48. Menezes, L.A.D.; Martins, J.M. The Jacupiranga mine Sao-Paulo, Brazil. Mineral. Rec. 1984, 15, 261–270. [Google Scholar]
  49. Alker, A.; Golob, P.; Postl, W.; Waltinger, H. Hydrotalkit, Nordstrandit und Motukoreait vom Stradner Kogel, südlich Gleichenberg, Steiermark. Mitt.-Bl. Abt. Mineral. Landesmuseum Joanneum 1981, 49, 1–13. [Google Scholar]
  50. Mills, S.J.; Christy, A.G.; Schmitt, R. The creation of neotypes for hydrotalcite. Mineral. Mag. 2016, 80, 1023–1029. [Google Scholar] [CrossRef]
  51. Raade, G. Hydrotalcite and quintinite from Dypingdal, snarum, Buskerud, Norway. Norsk Bergverksmuseum Skrifter 2013, 50, 55–57. [Google Scholar]
  52. Oliveira, S.B.D.; Sant’Agostino, L.M. Lithogeochemistry and 3D geological modeling of the apatite-bearing Mesquita Sampaio beforsite, Jacupiranga alkaline complex, Brazil. Brazil. J. Geol. 2020, 50, e20190071. [Google Scholar] [CrossRef]
  53. CrysAlisPro Software System, Version 1.171.38.46; Rigaku Oxford Diffraction: Oxford, UK, 2015.
  54. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. A 2015, 71, 3–8. [Google Scholar] [CrossRef]
  55. Dolomanov, O.V.; Bourhis, L.J.; Gildea, R.J.; Howard, J.A.; Puschmann, H. OLEX2: A complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339–341. [Google Scholar] [CrossRef]
  56. Britvin, S.N.; Dolivo-Dobrovolsky, D.V.; Krzhizhanovskaya, M.G. Software for processing of X-ray powder diffraction data obtained from the curved image plate detector of Rigaku RAXIS Rapid II diffractometer. Zapiski RMO 2017, 146, 104–107. (In Russian) [Google Scholar]
  57. TOPAS V4.2: General Profile and Structure Analysis Software for Powder Diffraction Data; Bruker-AXS: Karlsruhe, Germany, 2009.
  58. Zhitova, E.S.; Krivovichev, S.V.; Pekov, I.V.; Yakovenchuk, V.N.; Pakhomovsky, Y.A. Correlation between the d-value and the M2+: M3+ cation ratio in Mg–Al–CO3 layered double hydroxides. Appl. Clay Sci. 2016, 130, 2–11. [Google Scholar] [CrossRef]
  59. Guinier, A.; Bokij, G.B.; Boll-Dornberger, K.; Cowley, J.M.; Ďurovič, S.; Jagodzinski, H.; Krishna, P.; de Woff, P.M.; Zvyagin, B.B.; Cox, D.E.; et al. Nomenclature of polytype structures. Report of the International Union of Crystallography Ad hoc Committee on the nomenclature of disordered, modulated and polytype structures. Acta Crystallogr. A 1984, 40, 399–404. [Google Scholar] [CrossRef]
  60. Černý, P. Hydrotalcite z Véžná na zapadni Morave. Acta Musei Moraviae 1963, XLVII, 23–30. [Google Scholar]
  61. Allmann, R.; Jespen, H.P. Die struktur des hydrotalkits. Neues Jahrbuch Für Mineralogie Monatshefte 1969, 1969, 544–551. [Google Scholar]
  62. Lisitsina, V.A.; Drits, V.A.; Sokolova, T.N. New complex of secondary minerals-products of low temperature transformations of rocks, covering basalts of Atlantic Ocean underwater mountains. Lithol. Miner. Resour. 1985, 6, 20–39. [Google Scholar]
  63. Drits, V.A.; Sokolova, T.N.; Sokolova, G.V.; Cherkashin, V.I. New members of the hydrotalcite—Manasseite group. Clays Clay Miner. 1987, 35, 401–417. [Google Scholar] [CrossRef]
Figure 1. The crystal structure of quintinite from the Jacupiranga alkaline complex (Brazil) with stratified interlayers obtained previously (visualized using public cif-file) [43].
Figure 1. The crystal structure of quintinite from the Jacupiranga alkaline complex (Brazil) with stratified interlayers obtained previously (visualized using public cif-file) [43].
Crystals 13 00839 g001
Figure 2. The superstructures within the octahedral sheet of LDHs, suggested for quintinite (see text for details). Different colours correspond to different in-plane structures: 1 × 1 (brown); 3 × 3 (green); 2 3 × 2 3 (dark blue)—not yet confirmed.
Figure 2. The superstructures within the octahedral sheet of LDHs, suggested for quintinite (see text for details). Different colours correspond to different in-plane structures: 1 × 1 (brown); 3 × 3 (green); 2 3 × 2 3 (dark blue)—not yet confirmed.
Crystals 13 00839 g002
Figure 3. Schematic representation of quintinite, Mg4Al2(OH)12(CO3) × 3(H2O), polytypes 3R, 2H, 1M and 2T.
Figure 3. Schematic representation of quintinite, Mg4Al2(OH)12(CO3) × 3(H2O), polytypes 3R, 2H, 1M and 2T.
Crystals 13 00839 g003
Figure 4. Indexed powder X-ray diffraction pattern of quintinite from Jacupiranga, sample 91002 {space group P-3c1; lattice parameters from powder X-ray diffraction data: a = 5.2783 (3) Å, c = 15.1171 (16) Å, V = 364.74 (6) Å3}.
Figure 4. Indexed powder X-ray diffraction pattern of quintinite from Jacupiranga, sample 91002 {space group P-3c1; lattice parameters from powder X-ray diffraction data: a = 5.2783 (3) Å, c = 15.1171 (16) Å, V = 364.74 (6) Å3}.
Crystals 13 00839 g004
Table 1. Crystal data, data collection, and structure refinement details for quintinite samples 91002 and C7029.
Table 1. Crystal data, data collection, and structure refinement details for quintinite samples 91002 and C7029.
Sample91002C7029
Crystal chemical data
Crystal systemTrigonalTrigonal
Space groupP-3c1P-3c1
a (Å)5.2459 (14)5.2978 (5)
c (Å)15.110 (5)15.1991 (19)
V3)360.1 (2)369.44 (8)
Z11
Calculated density (g/cm3)2.1982.127
Absorption coefficient0.4860.472
Data collection
DiffractometerBruker Smart APEX II
Temperature (K)293
Radiation, wavelength (Å)MoKα, 0.71073
Range of data collection, 2θ (°)5.392–67.1745.36–72.626
h, k, l ranges−7→6, −7→7, −21→21−8→8, −8→8, −25→23
Total reflections collected55496783
Unique reflections (Rint)406 (0.0324)596 (0.0356)
Number of unique reflections F > 2σ(F)320424
Data completeness (%)100100
Structure refinement
Refinement methodFull-matrix least-squares on F2
Weighting coefficients a, b0.0410, 1.10220.0330, 1.5179
Data/restrain/parameters406/2/34596/2/34
R1 [F > 2σ(F)], wR2 [F > 2σ(F)]0.0541, 0.12880.0678, 0.1384
R1 all, wR2 all0.0685, 0.13780.0890, 0.1488
Goodness-of-fit on F21.1831.117
Largest diff. peak and hole (ēÅ−3)0.38/−0.400.63/−0.56
Table 2. Atom coordinates, equivalent isotropic displacement parameters (Å2), site occupancies and assigned site populations for quintinite samples 91002 and C7029.
Table 2. Atom coordinates, equivalent isotropic displacement parameters (Å2), site occupancies and assigned site populations for quintinite samples 91002 and C7029.
AtomxyZUeqs.o.f.W.P.Assigned Site
Populations
Sample 91002
Octahedral sheet
Mg0.33330.66670.4995 (1)0.0069 (4)0.3333 *4d
Al000.50.0107 (5)0.1667 *2bMg4Al2(OH)12
O10.3257 (5)0.3262 (5)0.5652 (1)0.0145 (4)1 *12g
H10.341 (9)0.315 (9)0.619 (3)0.036 (12)1 *12g
Interlayer gallery
C10.33330.66670.745 (3)0.010 (7)0.033 *4d(CO3)1.0(H2O)3.0 **
C20.360 (4)00.75000.028 (9)0.05 *6f
O20.290 (3)0.209 (2)0.7514 (8)0.033 (3)0.280 (11)12g
O30.362 (3)0.440 (2)0.751 (1)0.038 (4)0.291 (11)12g
Sample C7029
Octahedral sheet
Mg0.33330.66670.4999 (2)0.0100 (3)0.3333 *4d
Al000.50000.0124 (4)0.1667 *2bMg4Al2(OH)12
O10.3228 (4)0.3227 (5)0.5655 (1)0.0176 (4)1 *12g
H10.341 (10)0.314 (10)0.625 (2)0.034 (11)1 *12g
Interlayer gallery
C10.33330.66670.748 (6)0.025 (9)0.033 *4d
C20.363 (4)00.75000.021 (6)0.050 *6f(CO3)1.0(H2O)2.9 **
O20.291 (3)0.210 (3)0.7497 (12)0.038 (3)0.272 (11)12g
O30.360 (3)0.440 (3)0.7500 (14)0.042 (3)0.279 (11)12g
* fixed during refinement; ** electron per formula unit (epfu) for O2 and O3 sites is recalculated as (H2O)0. W.P.—Wyckoff position.
Table 3. Selected bond distances (Å) for quintinite samples 91002 and C7029.
Table 3. Selected bond distances (Å) for quintinite samples 91002 and C7029.
AtomAtomBond DistanceAtomAtomBond Distance
91002C702991002C7029
MgO12.027 (3)2.053 (3)C1O31.275 (10)1.277 (11)
MgO1 12.022 (2)2.051 (2)C2O21.317 (16)1.348 (16)
AlO11.974 (2)1.978 (2)C2O3 11.304 (14)1.308 (14)
1 Oxygen atoms, equivalent to O1 and O3, respectively, multiplied via symmetry elements.
Table 4. Hydrogen bonding scheme for quintinite samples 91002 and C7029.
Table 4. Hydrogen bonding scheme for quintinite samples 91002 and C7029.
SampleD–Hd (D–H)d (HA)<DHAd (DA)A
91002O1–H10.82 (4)2.06(5)168(4)2.866(13)O2
2.06(5)155(5)2.823(13)O2 1
2.08(4)157(5)2.857(15)O3
2.01(5)170(4)2.826(15)O3 1
C7029O1–H10.91 (4)1.96(4)166(4)2.850(19)O2
2.01(4)154(5)2.859(19)O2 1
2.00(4)156(5)2.86(2)O3
1.96(4)170(4)2.86(2)O3 1
1 Oxygen atoms, equivalent to O1 and O3, respectively, multiplied via symmetry elements.
Table 5. Quintinite worldwide findings and their unit-cell parameters.
Table 5. Quintinite worldwide findings and their unit-cell parameters.
LocalityKovdor,
Russia
Mont
Saint-Hilaire,
Canada
Jacupiranga,
Brazil
Bazhenovskoe
Deposit,
Russia
Ural Emerald
Mines
(Malyshevskoe),
Russia
OccurrenceLate-stage hydrothermal mineralMetosomatic alteration
d00n-value (Å) 17.53–7.597.577.587.567.51
a` (Å) 23.03–3.0653.053.053.0503.021
Reference[38,39,41,42][44][43]; this work[46][45]
LocalityWestern Moravia,
Czech Republic
Ampere and Josephine seamounts
of the North Atlantic
Caspian lowland, RussiaStrandel Kogel,
Austria
Occurrencequartz-oligoclase pegmatites of spentinite rocksecondary in basalt voidsevaporites (saline deposits)cavities of hauyn-nephelinite
d00n-value (Å)7.607.567.567.57
a` (Å)3.05?3.0423.035
Reference[60,61][62][63][49]
1 d00n-value is the distance between two neighboring metal-hydroxide layers. 2 a` is the distance between two neighboring metal cations in one metal-hydroxide layer.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhitova, E.S.; Sheveleva, R.M.; Zolotarev, A.A.; Krivovichev, S.V. The Crystal Structure of Mg–Al–CO3 Layered Double Hydroxide. Crystals 2023, 13, 839. https://doi.org/10.3390/cryst13050839

AMA Style

Zhitova ES, Sheveleva RM, Zolotarev AA, Krivovichev SV. The Crystal Structure of Mg–Al–CO3 Layered Double Hydroxide. Crystals. 2023; 13(5):839. https://doi.org/10.3390/cryst13050839

Chicago/Turabian Style

Zhitova, Elena S., Rezeda M. Sheveleva, Andrey A. Zolotarev, and Sergey V. Krivovichev. 2023. "The Crystal Structure of Mg–Al–CO3 Layered Double Hydroxide" Crystals 13, no. 5: 839. https://doi.org/10.3390/cryst13050839

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop