Next Article in Journal
Parametric Optimization in Rougher Flotation Performance of a Sulfidized Mixed Copper Ore
Next Article in Special Issue
Fedorite from Murun Alkaline Complex (Russia): Spectroscopy and Crystal Chemical Features
Previous Article in Journal
Vitreous Tesserae from the Four Seasons Mosaic of the S. Aloe Quarter in Vibo Valentia–Calabria, Italy: A Chemical Characterization
Previous Article in Special Issue
Synthesis of Zn-Saponite Using a Microwave Circulating Reflux Method under Atmospheric Pressure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Organically Templated Layered Uranyl Molybdate [C3H9NH+]4[(UO2)3(MoO4)5] Structurally Based on Mineral-Related Modular Units

1
Department of Crystallography, Saint-Petersburg State University, University emb. 7/9, 199034 St. Petersburg, Russia
2
Department of Chemistry, Moscow State University, GSP-1, 119991 Moscow, Russia
3
Kola Science Center, Russian Academy of Sciences, 184200 Apatity, Russia
*
Author to whom correspondence should be addressed.
Minerals 2020, 10(8), 659; https://doi.org/10.3390/min10080659
Submission received: 5 July 2020 / Revised: 22 July 2020 / Accepted: 24 July 2020 / Published: 25 July 2020

Abstract

:
A new organically templated uranyl molybdate [C3H9NH+]4[(UO2)3(MoO4)5] was prepared by a hydrothermal method at 220 °C. The compound is monoclinic, Сс, a = 16.768(6), b = 20.553(8), c = 11.897(4) Å, β = 108.195(7), V = 3895(2) Å3, R1 = 0.05. The crystal structure is based upon [(UO2)3(MoO4)5]4− uranyl molybdate layers. The isopropylammonium cations are located in the interlayer. The layers in the structure of [C3H9NH+]4[(UO2)3(MoO4)5] are considered as modular architectures. Topological analysis of layers with UO2:TO4 ratio of 3:5 (TVI = S, Cr, Se, Mo) was performed. Modular description is employed to elucidate the relationships between different structural topologies of [(UO2)3(TO4)5]4− layers and inorganic uranyl-based nanotubules. The possible existence of uranyl molybdate nanotubules is discussed.

1. Introduction

Investigations of natural and synthetic compounds of hexavalent uranium are important both for materials science and mineralogy. Uranium minerals are important constituents of oxidation zones of uranium deposits [1,2]. The development of the nuclear industry also requires further investigations in the area of safe nuclear waste management [3]. The latter requires proper understanding of phase formation processes, relationships between chemical compositions, crystal structures, and properties of the compounds formed. Of particular interest are both the chemical nature of the possible secondary phases and their interactions with the geological environment [4].
In the last two decades, particular attention has been paid to the hexavalent uranium (uranyl) compounds containing organic moieties, which are important for lanthanide/actinide separation. Common among these are the hybrid materials wherein the organic molecules are directly coordinated to the uranyl cation as ligands [5,6]. The templated structures based on weakly bonded organic and inorganic parts are less numerous [7] but exhibit essentially larger structural diversity of inorganic motifs.
The uranium compounds are formed during various steps of the nuclear cycle and present in the nuclear waste [3]. They accumulate various nuclides, thus preventing their migration into the environment [4]. The composition of nuclear waste is extremely complex and can contain both newly formed phases, as well as products of interaction between the fuel and container material, as well as with geological environment. The molybdenum isotopes (95Mo, 97Mo, 98Mo and 100Mo), formed during the fission process, are relatively stable and accumulate in the waste fuel. Topologically similar and even isostructural compounds are also known in the chemistry of pentavalent neptunium [7].
Though the natural and synthetic uranyl compounds are generally formed under essentially different conditions, their crystal structures demonstrate topological relationships [8]. Of more than 300 primary and secondary uranium minerals known to date (http://rruff.info/ima), the latter demonstrate the expected richer structural chemistry wherein uranium most commonly forms a linear uranyl cation coordinated by four, five, or six ligands in the equatorial plane. The coordination polyhedra can share only edges, only corners, or both edges and corners between each other or also with tetrahedral polyhedra of oxyanions. The group of compounds structurally based on exclusively corner sharing between UOn polyhedra and TO4 (TVI = S, Cr, Se, Mo) tetrahedra is represented in the secondary minerals. Their structures exhibit two types of 0D, four 1D, and four 2D complexes (Figure 1).
The 0D complexes are observed in the sulfate minerals belakovskiite, Na7(UO2)(SO4)4(SO3OH)∙3H2O [9] and bluelizardite, Na7(UO2)(SO4)4Cl∙2H2O [10], which are the secondary products of uraninite weathering; both species have been found in the Blue Lizard mine, San Juan, UT, USA. Though a large number of natural uranyl sulfates has been described [11], the complexes observed in belakovskiite (Figure 1а) and bluelizardite (Figure 1b) are yet unique.
The same locality also bears bobcookite, NaAl(UO2)2(SO4)4∙18H2O [12], fermiite, Na4(UO2)(SO4)3∙3H2O, and oppenheimerite, Na2(UO2)(SO4)2∙3H2O [13], whose structures are comprised of [(UO2)(SO4)2(H2O)]2− chains (Figure 1с). This topology is common for synthetic uranyl sulfates, selenates, and chromates [14]. They are also present in the structures of rietveldite, Fe(UO2)(SO4)2∙5H2O [15] and svornostite, K2Mg[(UO2)(SO4)2]2∙8H2O [16]. The chains observed in the structure of meisserite, Na5(UO2)(SO4)3(SO3OH)·H2O [17] can be derived from [(UO2)(SO4)2(H2O)]2− chains [18] via the addition of a protonated sulfate tetrahedron (Figure 1d). The structure of shumwayite, ((UO2)(SO4)(H2O)2)2∙H2O [19] is formed by association of [(UO2)(SO4)(H2O)2] chains (Figure 1e) and water via hydrogen bonding. In the structures of deloryite, Cu4(UO2)(MoO4)2(OH)6 [20], walpurgite, (BiO)4(UO2)(AsO4)2∙2H2O [21], and orthowalpurgite, (BiO)4(UO2)(AsO4)2∙2H2O [22], the [(UO2)(TO4)2]2− chains are comprised of UO6 tetragonal bipyramids and TO4 tetrahedra sharing vertices (Figure 1f). Their association leads to the formation of autunite-type layers (Figure 1g) [23].
A large number of uranium minerals with a common formula An+[(UO2)(TO4)](H2O)mXk adopt the structures based on autunite-related [(UO2)(TVO4)] layers. Therein, A may stand for a single, double, or even triple-charged cation while T is P or As, and X is generally F or OH. The layers observed in the structures of magnesioleydetite, Mg(UO2)(SO4)2∙11H2O, straßmannite, Al(UO2)(SO4)2F∙16H2O [24], geschieberite, K2(UO2)(SO4)2∙2H2O [25], and leydetite, Fe(UO2)(SO4)2∙11H2O [26] are topologically identical (Figure 1h). Therein, four out of five equatorial vertices of uranium polyhedra are occupied by oxygen atoms of sulfate tetrahedra while the fifth is occupied by a water molecule; all such vertices are similarly oriented in the layer plane. The layers in the structure of wetherillite, Na2Mg(UO2)2(SO4)4∙18H2O [12], differ in that the vertices occupied by water molecules exhibit two positions (Figure 1i). Yet another layered arrangement is formed by vertex-sharing UO6(H2O) and SO4 moieties (Figure 1j) in the structure of beshtauite, (NH4)2(UO2)(SO4)2∙2H2O [27]. The topology of this layer is similar to that of goldichite, KFe(SO4)2∙4H2O [28].
The structural diversity of natural and synthetic compounds is mostly governed by their formation conditions, including the acidity/basicity and the mechanism of crystallization. The majority of the minerals discussed above have been found on the walls of abandoned mines [10]. They crystallize upon evaporation under almost invariable temperature but decreasing pH. These conditions are expected to affect the structural outcome. The oxidation zones are formed under conditions changing from basic to weakly acidic [1]. Interestingly, in most structures of these minerals, UOn bipyramids and TO4 tetrahedra preferably involve corner sharing only.
Essentially richer structural chemistry with similar architecture is observed for synthetic uranyl compounds, particularly for those that are organically templated and formed in acidic environments. Hydrogen-bonded motifs are particularly common in organically templated synthetic uranyl compounds. Most synthetic protocols, using evaporation techniques, for uranyl compounds with tetrahedral anions employ essentially acidic media (pH = 2–4 and below) [29,30,31,32].
Hence, the minerals formed upon solution evaporation on the mine walls, as well as synthetic compounds, also formed by evaporation of acidic solutions, demonstrate topologically similar crystal structures. This indicated that studies of synthetic compounds, their topological variability and structural peculiarities are of essential importance.
Among both natural and synthetic uranium compounds, the largest topological diversity is exhibited by the structures of compounds containing tetrahedral TO42− anions (TVI = S, Cr, Se, Mo) [14]. Despite chemical differences, they commonly demonstrate topologically identical structural motifs. Note that crystal chemistry of MoVI compounds is more complex due to its flexible coordination environments.
In this paper, we report the synthesis and structure of new uranyl molybdate [C3H9NH+]4[(UO2)3(MoO4)5] (1). Its structure is based upon [(UO2)3(MoO4)5]4− uranyl molybdate layers, which may be considered as a precursor for nanotubule formation.

2. Materials and Methods

Synthesis. Crystals of [C3H9NH+]4[(UO2)3(MoO4)5] (1) were synthesized from a solution of 0.1 g of (UO2)(NO3)2·6H2O (Vekton, 99.7%), 0.0576 g of MoO3 (Vekton, 99.7%), 0.01 mL of isopropylamine (Aldrich, 99.5%) and 0.053 g of HCl in 5 mL of H2O (with an approximate U:Mo: isopropylamine:HCl:H2O molar gel ratio of 3:5:8:15:1550). The solution was placed in a Teflon-lined Parr reaction vessel and heated to 220 °C for 36 h (6 °C/h), followed by slow cooling to ambient temperature. The cooling rate was 2 °C/h. The crystals occur as aggregates of greenish-yellow transparent plates up to 0.2 mm in maximum dimension.
Single-crystal studies. A single crystal was attached to glass fiber using an epoxy resin and mounted on a Bruker SMART APEX II DUO diffractometer (Bruker, Karlsruhe, Germany) equipped with a micro-focus X-ray tube utilizing Mo radiation. The experimental data set was collected at 100 K. Unit cell parameters were calculated using least-squares fits. Structure factors were derived using APEX 2 after introducing the required corrections. The structure was solved using direct methods. Further details are collected in Table 1. The final model includes site coordinates and anisotropic thermal parameters for all atoms except hydrogens, which were localized using AFIX command in calculated positions (d(C-H, N-H) = 1.00 Å). The data are deposited in CCDC under Entry No. 2014144. Selected interatomic distances are given in Table 2.

3. Results

The structure of 1 (Figure 2a) contains three symmetrically independent uranium atoms each forming the uranyl cation (<U-O> = 1.780, 1.745, 1.760 Å, for U1, U2, U3, respectively). (UO2)2+ cations (Ur) are each coordinated by five oxygen atoms in the equatorial plane (<Ur-Oeq> = 2.376, 2.391, 2.375 Å, respectively). Five symmetrically unique molybdenum atoms form (MoO4)2− tetrahedra (<Mo-O> = 1.742, 1.749, 1.760, 1.755, 1.725 Å, respectively).
The bond valence sums for U and Mo were calculated using parameters given in [33]: 5.98, 6.17 and 6.14 for U1, U2 and U3, and 6.12, 6.02, 5.86, 5.93 and 6.39 for Mo1, Mo2, Mo3, Mo4 and Mo5, respectively. These values are in agreement with the reference data except for the somewhat overbonded Mo5 centering a strongly distorted tetrahedron (Mo-O distances to O24, O13, O26 and O25 are 1.765(13) Å, 1.767(12) Å, 1.702(14) Å, and 1.666(16) Å). This distortion is likely due to a strong N-H···O hydrogen bond (d (N1–O25) = 2.150 Å) (Figure 2b).
The UO7 and MoO4 polyhedra share corners to form the [(UO2)3(MoO4)5]4− layers (Figure 3a). Protonated molecules of isopropylamine are located in the interlayer and connected to the layers via hydrogen bonding (Figure 2b). The [(UO2)3(MoO4)5]4− layer topology has been described earlier in the structures of uranyl molybdates [34], selenates [35] and chromates [36]. The orientation matrix for the terminal oxygen atoms of the molybdate tetrahedra in 1 is given in Figure 3b. A compound with the same chemical composition, same graph of the layers (Figure 3c), but with the different orientation matrix for the terminal oxygen atoms of the molybdenum tetrahedra was found in the structure of [C3N2H12](H3O)2[(UO2)3(MoO4)5] [34]. According to the definition given in [37], such layers are orientation geometric isomers. Thus, the orientation matrix for the [(UO2)3(MoO4)5]4− layer in 1 has not been observed before.
There are three common approaches to the description of the structural topology of uranium minerals and synthetic compounds: anionic topologies, graphs, and modular description. In [38], the latter approach was applied to demonstrate that structures of some uranyl selenates may be obtained via assembly of [(UO2)2(SeO4)4(H2O)4]4− modules. The layers in the structure of [Co(H2O)6]3(UO2)5(SO4)8(H2O)·H2O can also be viewed as a modular arrangement formed by the assembly of chains [39]. A modular approach was also applied to the description of some uranyl chromates, sulfates and selenates [40,41]. It was shown [41] that layers with the ratio UO2:TO4 = 2:3 can be obtained by the association of two types of fundamental chains.
The [(UO2)3(MoO4)5]4− layer in 1 (Figure 4) can be also described as consisting of several modules, i.e., fundamental chains С1, С2, С`1. The C1 and С`1 chains are related by a mirror plane. In C1 chains, (U1O2)O5 and (U2O2)O5 bipyramids share vertices with Mo1O4 and Mo2O4 tetrahedra, to form the [(UO2)2(MoO4)4]4− complexes linked via additional Mo3O4 species. С2 chains are formed from two (U3O2)O5 bipyramids and Mo4O5 and Mo5O5 species. The chains are stacked in the following sequence ...C1C2C`1C1C2C`1…
Three topologies of UO2:TO4 = 3:5 layers [14] are known to date. Two of them are represented in Figure 5. In all cases, the layers can be described as alternations of chains.
The [(UO2)3(TO4)5]4− layers in the structure of Zn2((UO2)3(SeO4)5)·17(H2O) (Figure 5a) are constructed in a way similar to the layers in 1. They are formed by combination of C2 and C`1 chains (Figure 5b) in a …C2C`1C`1C2C`1C`1… sequence. The layers in the structure of Rb2((UO2)2(SeO4)3(H2O)2)⋅4(H2O) (Figure 5с) exhibit a different topology and are formed from three types of chains: С1, С3, and С`1 (Figure 5d). The С3 chain can be obtained from С2 by the removal of one link between the nodes in every second four-membered cycle.

4. Discussion

4.1. Structural Topology of Nanotubules Is Uranyl Oxysalts

Four uranyl compounds with nanotubular (NT) motifs formed by corner sharing between UO7 bipyramids and TO4 tetrahedra have been described to date: three uranyl-selenates [42,43] and one uranyl-sulfate [44]. Nanotubules differ both in the diameter and structural topology of their walls. The compounds K5(UO2)3(SeO4)5(NO3)·3.5(H2O) and (H3O)2K[(H3O)@([18]crown-6)][(UO2)3(SeO4)5]·4H2O are based on identical nanotubules [(UO2)3(SeO4)5]4− (Figure 6a). Their outer diameter is 17 Å and the inner one (defined as the distance between the opposite closest oxygen atoms) is 7.4 Å. In the structure of (C4H12N)14[(UO2)10(SeO4)17(H2O)] the [(UO2)10(SeO4)17(H2O)]14− nanotubules are elliptical with dimension 25 × 23 Å (inner diameter is 15.3 Å) (Figure 6c). The size of [(UO2)6(SO4)10]8− nanotubules (Figure 6b) in the structure of Na(C8H10NO2)7[(UO2)6(SO4)10]·3.5H2O [44] is intermediate (14 × 22 Å, inner diameter is 13.5 Å).
The approach to the description of nanotubules as structural entities formed by layer scrolling is classic, and was applied for carbon NTs [45], transition-metal chalcogenides (MoS2, MoSe2, WS2, WSe2, NbS2 and NbSe2) [46], as well as oxides (TiO2, VOx, CuO, Al2O3, SiO2, etc.) [47]. The topology of the smallest nanotubules (Figure 6a) corresponds to that of [(UO2)3(CrO4)5]4− layers found in the structures of Mg2[(UO2)3(CrO4)5]·17(H2O) and Ca2[(UO2)3(CrO4)5]·19(H2O) [48]. The [(UO2)6(SO4)10]8− nanotubules (Figure 6b) are produced from a similar layered archetype [44]. However, there are no layered “ancestors” for the [(UO2)10(SeO4)17(H2O)]14− nanotubules (Figure 6c), yet their local topology is similar to those of 3:5 layers, which were found, e.g., in the structure of Rb4[(UO2)3(SeO4)5(H2O)] [49].
If we dissect the [(UO2)3(SeO4)5]4− tube, as shown in Figure 6а, the planar projection of the layer (Figure 6d) corresponds to the sequence of chains …С1С2С1С1С2С1… Its repetition in a planar fashion ultimately leads to a layer with UO2:TO4 = 3:5.
The NT in Na(C8H10NO2)7[(UO2)6(SO4)10]·3.5H2O [44] can be split (Figure 6b) into fundamental chains with the sequence of …С`1С2С`1С`1С2С`1С`1C2C`1 (Figure 6e). It is interesting to note that the sequence of chains is similar to those in K5(UO2)3(SeO4)5(NO3)·3.5(H2O), but nine chains are required to build a tube.
The NTs in the structure of (C4H12N)14[(UO2)10(SeO4)17(H2O)] [42] can be dissected into ten chains (Figure 6i). In this case, the mutual arrangement of the polyhedra differs somewhat from the above discussed cases. Three chains (C`1, C`1, C2) form a ribbon cut from the UO2:TO4 = 3:5 layer.
In the topology of [(UO2)10(SeO4)17(H2O)]14− nanotubules, the C`1C`1C2 combination is repeated twice. Overall, the NT is formed from ten chains. Of the remaining four, two chains correspond to the С3 described earlier, and two, to С4. The stacking sequence therefore is …С4С3С`1C`1C2C4C3C`1C`1C2… The С4 are produced from С1 if one link is broken between the tetrahedron and bipyramid in the (UO2)2(TO4)2 cycle. Note the essential disorder [43] of some oxygen positions in the structure of (C4H12N)14[(UO2)10(SeO4)17(H2O)]. One may assume that in the idealized ordered version of the [(UO2)10(SeO4)17(H2O)]14− NT, the topology of the С4 chains would correspond to С1.

4.2. The Flexibility of U-O-T Bridges

As follows from above, the inorganic uranyl NTs can be described as comprised of fundamental chains and, at a higher hierarchical level, derived from 3:5 layers or their fragments. The chain propagation axes are collinear to the axis of the NTs themselves. The NT in K5[(UO2)3(SeO4)5](NO3) 3.5H2O is comprised of six fundamental chains oriented parallel to the NT axis (Figure 6g). The scrolling of the precursor layer into the NT is probably enhanced by the flexibility of the U-O-T bridges. This essentially resembles a needle bearing wherein the rollers are replaced by the fundamental chains. The other NTs can be constructed in a similar way: the 7.4 Å-NT requires six chains (Figure 6g), 13.5 Å-NT, nine chains (Figure 6h) while ten are necessary for the 15.3 Å-NT (Figure 6i). To date, all known NTs differ merely by the number of fundamental chains they are comprised of.
It was noted earlier [14] that vertex-sharing between the UOn and TO4 polyhedra results in the formation of a flexible “hinge” as the corresponding U-O-T angles can vary in a relatively broad range, providing the flexibility of the overall polyhedral backbone sufficient to compensate, for instance, the overall thermal expansion [50,51]. Yet, this mechanism seems not to contribute essentially to the formation of the NTs. The distribution of the U-O-T angles in the UO2:TO4 = 3:5 layers (Figure 7a) covers a broad range of 120°–165°; it has a nearly normal character with a maximum at 135°–140°. Almost the same is observed for the four NTs known to date (Figure 7b), the [(UO2)6(SO4)10]8− NTs are the most flexible (U-O-S angles range from 125° to 165°). The distribution is almost the same for [(UO2)3(SeO4)5]4− and [(UO2)10(SeO4)17(H2O)]14− NTs, the smallest one being evidently the most rigid.

5. Conclusions

We described the synthesis and structure of the new uranyl molybdate templated by protonated isopropylamine molecules. The structure of modular [(UO2)3(MoO4)5]4− layers in 1 represents a new isomer, not observed previously in uranyl molybdates.
Four different fundamental chains (C1, C`1, C2, C3) are sufficient for constructing any known layered topology with UO2:TO4 = 3:5, whereas five chains (C1, C`1, C2, C3, C4) are required for the known architectures of uranyl-based inorganic NTs (Figure 8).
The layers in the structure of 1 are topologically similar to the layers forming NTs in uranyl sulfates and selenates. In addition, the [(UO2)3(MoO4)5]4− layers in 1 are identical to those in uranyl chromates. We can assume the formation of NTs with a similar topology in the structures of uranyl molybdates and uranyl chromates. The rarity of uranyl-based inorganic NTs, known to date, stems probably from the very narrow range of conditions which favor layer scrolling and which are only poorly understood to date.
While the majority of uranyl compounds adopt layered structures where the 1:2 and 2:3 layers are most common and exhibit rich topological diversity, these layers could not yet be scrolled. The formation of these layers requires only two types of fundamental chains [41], which is probably insufficient to form a NT. Out of ca. 1000 known uranyl compounds, only 10 contain the UO2:TO4 = 3:5 layers, which can be scrolled, and just a handful of 4:7 and 5:8 representatives are known. Hence, formation conditions are very specific already for the “parental” layers, to say nothing about those for the scrolling and formation of nanotubules.
The results of the topological analysis of layered UO2:TO4 = 3:5 architectures and derived nanotubules suggest that the latter may also form in uranyl molybdate systems.

Author Contributions

E.N. and O.S. designed the study, performed and interpreted single crystal X-ray diffraction experiments; D.C. performed synthesis; E.N., O.S. and D.C. wrote the paper; S.K. provided work in the radiochemical laboratory. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Russian Science Foundation through the grant 16-17-10085.

Acknowledgments

Technical support by the SPbSU X-ray Diffraction and Microscopy and Microanalysis Resource Centers is gratefully acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Belova, L.N. Zones of Oxidation of Hydrothermal Deposits of Uranium; Nedra: Moscow, Russia, 1975; p. 158. [Google Scholar]
  2. Hazen, R.M.; Ewing, R.C.; Sverjensky, D.A. Evolution of uranium and thorium minerals. Am. Mineral. 2009, 94, 1293–1311. [Google Scholar] [CrossRef]
  3. Finch, R.J.; Murakami, T. Systematics and paragenesis of uranium minerals. In Uranium: Mineralogy, Geochemistry and the Environment; Burns, P.C., Finch, R.J., Eds.; Mineralogical Society of America: Chantilly, VA, USA, 1999; Reviews in Mineralogy; Volume 38, pp. 91–180. [Google Scholar]
  4. Baker, R.J. Uranium minerals and their relevance to long term storage of nuclear fuels. Coord. Chem. Rev. 2014, 266–267, 123–136. [Google Scholar] [CrossRef]
  5. Andrews, M.B.; Cahill, C.L. Uranyl bearing hybridmaterials: Synthesis, speciation, and solid-state structures. Chem. Rev. 2013, 113, 1121–1136. [Google Scholar] [CrossRef] [PubMed]
  6. Loiseau, T.; Mihalcea, I.; Henry, N.; Volkringer, C. The crystal chemistry of uranium carboxylates. Coord. Chem. Rev. 2014, 266–267, 69–109. [Google Scholar] [CrossRef]
  7. Krivovichev, S.V. Structural crystallography of inorganic oxysalts. Crystallogr. Rev. 2009, 15, 279–281. [Google Scholar]
  8. Krivovichev, S.V.; Plášil, J. Mineralogy and crystallography of uranium. In Uranium: Cradle to Grave; Mineralogical Association of Canada Short Course Series; Burns, P.C., Sigmon, G.E., Eds.; Mineralogical Association of Canada: Quebec, QC, Canada, 2013; pp. 15–120. [Google Scholar]
  9. Kampf, A.R.; Plášil, J.; Kasatkin, A.V.; Marty, J. Belakovskiite, Na7(UO2)(SO4)4(SO3OH)(H2O)3, a new uranyl sulfate mineral from the Blue Lizard mine, San Juan County, Utah, USA. Mineral. Mag. 2014, 78, 639–649. [Google Scholar] [CrossRef]
  10. Plášil, J.; Kampf, A.R.; Kasatkin, A.V.; Marty, J. Bluelizardite, Na7(UO2)(SO4)4Cl(H2O)2, a new uranyl sulfate mineral from the Blue Lizard mine, San Juan County, Utah, USA. J. Geosci. Czech. 2014, 59, 145–158. [Google Scholar] [CrossRef]
  11. Gurzhiy, V.V.; Plášil, J. Structural complexity of natural uranyl sulfates. Acta Crystallogr. Sect. B 2018, 75, 39–48. [Google Scholar] [CrossRef] [Green Version]
  12. Kampf, A.R.; Plášil, J.; Kasatkin, A.V.; Marty, J. Bobcookite, NaAl(UO2)2(SO4)4·18H2O and wetherillite, Na2Mg(UO2)2(SO4)4×18H2O, two new uranyl sulfate minerals from the Blue Lizard mine, San Juan County, Utah, USA. Mineral. Mag. 2015, 79, 695–714. [Google Scholar] [CrossRef]
  13. Kampf, A.R.; Plášil, J.; Kasatkin, A.V.; Marty, J.; Čejka, J. Fermiite, Na4(UO2)(SO4)3∙3H2O and oppeneimerite, Na2(UO2)(SO4)2·3H2O, two new uranyl sulfate minerals from the Blue Lizard mine, San Juan County, Utah, USA. Mineral. Mag. 2015, 79, 1123–1142. [Google Scholar] [CrossRef]
  14. Krivovichev, S.V.; Burns, P.C.; Tananaev, I.G. (Eds.) Structural Chemistry of Inorganic Actinide Compounds; Elsevier: Amsterdam, The Netherlands, 2007; p. 494. [Google Scholar]
  15. Kampf, A.R.; Sejkora, J.; Witzke, T.; Plášil, J.; Čejka, J.; Nash, B.P.; Marty, J. Rietveldite, Fe(UO2)(SO4)2(H2O)5, a new uranyl sulfate mineral from Giveaway-Simplot mine (Utah, USA), Willi Agatz mine (Saxony, Germany) and Jáchymov (Czech Republic). J. Geosci. Czech. 2017, 62, 107–120. [Google Scholar] [CrossRef] [Green Version]
  16. Plášil, J.; Hloušek, J.; Kasatkin, A.V.; Novak, M.; Čejka, J.; Lapčák, L. Svornostite, K2Mg[(UO2)(SO4)2]2 8H2O, a new uranyl sulfate mineral from Jáchymov, Czech Republic. J. Geosci. Czech. 2015, 60, 113–121. [Google Scholar] [CrossRef] [Green Version]
  17. Plášil, J.; Kampf, A.R.; Kasatkin, A.V.; Marty, J.; Skoda, R.; Silva, S.; Cejka, K. Meisserite, Na5(UO2)(SO4)3(SO3OH)(H2O), a new uranyl sulfate mineral from the Blue Lizard mine, San Juan County, Utah, USA. Mineral. Mag. 2013, 77, 2975–2988. [Google Scholar] [CrossRef]
  18. Nazarchuk, E.V.; Charkin, D.O.; Siidra, O.I.; Gurzhiy, V.V. Crystal-chemical features of U(VI) compounds with inorganic complexes derived from [(UO2)(TO4)(H2O)n], T = S, Cr, Se: Synthesis and crystal structures of two new uranyl sulfates. Radiochemistry 2018, 60, 345–351. [Google Scholar] [CrossRef]
  19. Kampf, A.R.; Plášil, J.; Kasatkin, A.V.; Marty, J.; Čejka, J.; Lapčák, L. Shumwayite, [(UO2)(SO4)(H2O)2]2 H2O, a new uranyl sulfate mineral from Red Canyon, San Juan County, Utah, USA. Mineral. Mag. 2017, 81, 273–285. [Google Scholar] [CrossRef]
  20. Pushcharovskii, D.Y.; Rastsvetaeva, R.K.; Sarp, H. Crystal structure of deloryite, Cu4(UO2)(Mo2O8)(OH)6. J. Alloys Compd. 1996, 239, 23–26. [Google Scholar] [CrossRef]
  21. Mereiter, K. The crystal structure of walpurgite, (UO2)Bi4O4(AsO4)2·2H2O. Tschermaks Mineral. Petrogr. Mitt. 1982, 30, 129–139. [Google Scholar] [CrossRef]
  22. Krause, W.; Effenberger, H.; Brandstaetter, F. Orthowalpurgite, (UO2)Bi4O4(AsO4)2·2(H2O), a new mineral from the Black Forest, Germany. Eur. J. Inorg. Mineral. 1995, 7, 1313–1324. [Google Scholar]
  23. Locock, A.J.; Burns, P.C. The crystal structure of synthetic autunite, Ca((UO2)(PO4))2(H2O)11. Am. Mineral. 2003, 88, 240–244. [Google Scholar] [CrossRef]
  24. Kampf, A.R.; Plášil, J.; Kasatkin, A.V.; Nash, B.P.; Marty, J. Magnesioleydetite and straβmannite, two new uranyl sulfate minerals with sheet structures from Red Canyon, Utah. Mineral. Mag. 2019, 83, 349–360. [Google Scholar] [CrossRef]
  25. Alekseev, E.V.; Suleimanov, E.V.; Chuprunov, E.V.; Marychev, M.O.; Ivanov, V.; Fukin, G.K. Crystal structures and nonlinear optical properties of K2UO2(SO4)2∙2H2O compound al 293K. Crystallogr. Rep. 2006, 51, 29–33. [Google Scholar] [CrossRef]
  26. Plášil, J.; Kasatkin, A.V.; Skoda, R.; Novak, M.; Kallistova, A.; Dusek, M.; Skala, R.; Fejfarova, K.; Cejka, J.; Meisser, N.; et al. Leydetite, Fe(UO2)(SO4)2(H2O)11, a new uranyl sulfate mineral from Mas d’Alary, Lodeve, France. Mineral. Mag. 2013, 77, 429–441. [Google Scholar] [CrossRef]
  27. Pekov, I.V.; Krivovichev, S.V.; Yapaskurt, V.O.; Chukanov, N.V.; Belakovskiy, D.I. Beshtauite, (NH4)2(UO2)(SO4)2·2(H2O), a new mineral from Mount Beshtau, Northern Caucasus, Russia. Am. Mineral. 2014, 99, 1783–1787. [Google Scholar] [CrossRef] [Green Version]
  28. Márquez-Zavalía, M.F.; Galliski, M.A. Goldichite of fumarolic origin from the Santa 1062 Bárbara mine, Jujuy, Northwestern Argentina. Can. Mineral. 1995, 33, 1059–1062. [Google Scholar]
  29. Ling, J.; Sigmon, G.E.; Ward, M.; Roback, N.; Burns, P.C. Syntheses, structures, and IR spectroscopic characterization of new uranyl sulfate/selenate 1D-chain, 2D-sheet and 3D-framework. Z. Kristallogr. 2010, 225, 230–239. [Google Scholar] [CrossRef]
  30. Doran, M.B.; Norquist, A.J.; O’Hare, D. Exploration of composition space in templated uranium sulfates. Inorg. Chem. 2003, 42, 6989–6995. [Google Scholar] [CrossRef]
  31. Norquist, A.J.; Doran, M.B.; Thomas, P.M.; O’Hare, D. Structural diversity in organically templated uranium sulfates. Dalton Trans. 2003, 6, 1168–1175. [Google Scholar] [CrossRef]
  32. Siidra, O.I.; Nazarchuk, E.V.; Suknotova, A.N.; Kayukov, R.A.; Krivovichev, S.V. Cr(VI) trioxide as a starting material for the synthesis of novel zero-, one-, and two-dimensional uranyl dichromates and chromate-dichromates. Inorg. Chem. 2013, 52, 4729–4735. [Google Scholar] [CrossRef]
  33. Gagné, O.C.; Hawthorne, F.C. Comprehensive derivation of bond-valence parameters for ion pairs involving oxygen. Acta Cryst. 2015, B71, 561–578. [Google Scholar] [CrossRef] [Green Version]
  34. Halasyamani, P.S.; Francis, R.J.; Walker, S.M.; O’Hare, D. New layered uranium(VI) molybdates:  syntheses and structures of (NH3(CH2)3NH3)(H3O)2(UO2)3(MoO4)5, C(NH2)3(UO2)(OH)(MoO4), (C4H12N2)(UO2)(MoO4)2 and (C5H14N2)(UO2)(MoO4)2H2O. Inorg. Chem. 1999, 38, 271–279. [Google Scholar] [CrossRef]
  35. Krivovichev, S.V.; Tananaev, I.G.; Myasoedov, B.F. Geometric isomerism of layered complexes of uranyl selenates: Synthesis and structure of (H3O)[C5H14N]2[(UO2)3(SeO4)4(HSeO4)(H2O)] and (H3O)[C5H14N]2[(UO2)3(SeO4)4(HSeO4)(H2O)](H2O). Radiochemistry 2006, 48, 552–560. [Google Scholar] [CrossRef]
  36. Krivovichev, S.V.; Burns, P.C. Crystal chemistry of K uranyl chromates: Crystal structures of K5[(UO2)(CrO4)3](NO3)(H2O)3, K4[(UO2)3(CrO4)5](H2O)7 and K2[(UO2)2(CrO4)3](H2O)6. Z. Kristallogr. Cryst. Mater. 2003, 218, 725–730. [Google Scholar]
  37. Moore, P.B. Laueite, pseudolaueite, stewartite and metavauxite: A study in combinatorial polymorphism. Neu. Jb. Mineral. Abh. 1975, 123, 148–159. [Google Scholar]
  38. Krivovichev, S.V.; Tananaev, I.G. Microscopic model of crystal genesis from aqueous solution of uranyl selenates. J. Sib. Fed. Uni. Chem. 2009, 2, 133–149. [Google Scholar]
  39. Krivovichev, S.V.; Burns, P.C. The modular structure of the novel uranyl sulfate sheet in [Co(H2O)6]3[(UO2)5(SO4)8(H2O)](H2O)5. J. Geosci. Czech. 2014, 59, 135–143. [Google Scholar] [CrossRef] [Green Version]
  40. Nazarchuk, E.V.; Siidra, O.I.; Charkin, D.O. Specific features of the crystal chemistry of layered uranyl compounds with the ratio UO2:TO4 = 5:8 (T = S6+, Cr6+, Se6+, Mo6+). Radiochemistry 2018, 60, 352–361. [Google Scholar] [CrossRef]
  41. Nazarchuk, E.V.; Charkin, D.O.; Kozlov, D.V.; Siidra, O.I.; Kalmykov, S.N. Topological analysis of the layered uranyl compounds bearing slabs with UO2:TO4 ratio of 2:3. Radiochim. Acta. 2020, 108, 249–260. [Google Scholar] [CrossRef]
  42. Krivovichev, S.V.; Tananaev, I.G.; Kahlenberg, V.; Kaindl, R.; Myasoedov, B.F. Synthesis, structure, and properties of inorganic nanotubes based on uranyl selenates. Radiochemistry 2005, 47, 525–536. [Google Scholar] [CrossRef]
  43. Alekseev, E.V.; Krivovichev, S.V.; Depmeier, W. A crown ether as template for microporous and nanostructured uranium compounds. Angew. Chem. Int. Ed. 2008, 47, 549–551. [Google Scholar] [CrossRef]
  44. Siidra, O.I.; Nazarchuk, E.V.; Charkin, D.O.; Chukanov, N.V.; Depmeier, W.; Bocharov, S.N.; Sharikov, M.I. Uranyl sulfate nanotubules templated by N-phenylglycine. Nanomaterials 2018, 8, 216. [Google Scholar] [CrossRef] [Green Version]
  45. Iijima, S. Helical microtubules of graphitic carbon. Nature 1991, 354, 56–58. [Google Scholar] [CrossRef]
  46. Tenne, R.; Margulis, L.; Genut, M.; Hodes, G. Polyhedral and cylindrical structures of tungsten disulphide. Nature 1992, 360, 444–446. [Google Scholar] [CrossRef]
  47. Imai, H.; Matsuta, M.; Shimizu, K.; Hirashima, H.; Negishi, N. Preparation of TiO2 fibers with well-organized structures. J. Mater. Chem. 2000, 10, 2005–2006. [Google Scholar] [CrossRef]
  48. Krivovichev, S.V.; Burns, P.C. Geometrical isomerism in uranyl chromates II. Crystal structures of Mg2[(UO2)3(CrO4)5] 17(H2O) and Ca2[(UO2)3(CrO4)5] 19(H2O). Z. Crystallogr. 2003, 218, 683–690. [Google Scholar] [CrossRef]
  49. Krivovichev, S.V.; Kahlenberg, V. Structural diversity of sheets in rubidium uranyl oxoselenates: Synthesis and crystal structures of Rb2((UO2)(SeO4)2(H2O))(H2O), Rb2((UO2)2(SeO4)3(H2O)2)(H2O)4 and Rb4((UO2)3(SeO4)5(H2O)). Z. Anorg. Allg. Chem. 2005, 631, 739–744. [Google Scholar] [CrossRef]
  50. Nazarchuk, E.V.; Krivovichev, S.V.; Filatov, S.K. Phase transitions and high-temperature crystal chemistry of polymorphous modifications of Cs2(UO2)2(MoO4). Radiochemistry 2004, 46, 438–440. [Google Scholar] [CrossRef]
  51. Siidra, O.I.; Nazarchuk, E.V.; Kayukov, R.A.; Bubnova, R.S.; Krivovichev, S.V. CrVI–CrV transition in Uranyl Chromium compounds: Synthesis and high-temperature X-ray diffraction study of Cs2[(UO2)2(CrO4)3]. Z. Anorg. Allg. Chem. 2013, 639, 2302–2306. [Google Scholar] [CrossRef]
Figure 1. Heteropolyhedral complexes with corner sharing only between UOn bipyramids and TO4 tetrahedra in the mineral crystal structures: (a) belakovskiite, (b) bluelizardite, (c) bobcookite, (d) meisserite, (e) shumwayite, (f) walpurgite, (g) autunite, (h) straßmannite, (i) wetherillite and (j) beshtauite.
Figure 1. Heteropolyhedral complexes with corner sharing only between UOn bipyramids and TO4 tetrahedra in the mineral crystal structures: (a) belakovskiite, (b) bluelizardite, (c) bobcookite, (d) meisserite, (e) shumwayite, (f) walpurgite, (g) autunite, (h) straßmannite, (i) wetherillite and (j) beshtauite.
Minerals 10 00659 g001
Figure 2. (a) General projection of the crystal structure of 1 along the c axis and (b) hydrogen bonding with isopropylamine molecules in the interlayer (UO7 bipyramids = orange, MoO4 = blue, H = black, C = grey).
Figure 2. (a) General projection of the crystal structure of 1 along the c axis and (b) hydrogen bonding with isopropylamine molecules in the interlayer (UO7 bipyramids = orange, MoO4 = blue, H = black, C = grey).
Minerals 10 00659 g002
Figure 3. (a) The [(UO2)3(MoO4)5]4− layer in the structure of 1; (b) orientation matrix of the terminal oxygen atoms (non-shared vertices of MoO4 tetrahedra), and (с) the corresponding graph.
Figure 3. (a) The [(UO2)3(MoO4)5]4− layer in the structure of 1; (b) orientation matrix of the terminal oxygen atoms (non-shared vertices of MoO4 tetrahedra), and (с) the corresponding graph.
Minerals 10 00659 g003
Figure 4. The topology of [(UO2)3(MoO4)5]4 layer in 1.
Figure 4. The topology of [(UO2)3(MoO4)5]4 layer in 1.
Minerals 10 00659 g004
Figure 5. (a) Structure of the [(UO2)3(TO4)5]4− layers in Zn2((UO2)3(SeO4)5)∙17(H2O) and (c) Rb2((UO2)2(SeO4)3(H2O)2)∙4(H2O) and (b, d) their split into fundamental chains.
Figure 5. (a) Structure of the [(UO2)3(TO4)5]4− layers in Zn2((UO2)3(SeO4)5)∙17(H2O) and (c) Rb2((UO2)2(SeO4)3(H2O)2)∙4(H2O) and (b, d) their split into fundamental chains.
Minerals 10 00659 g005
Figure 6. (a) Known uranyl-based nanotubules in K5(UO2)3(SeO4)5(NO3)·3.5(H2O), (b) Na(C8H10NO2)7[(UO2)6(SO4)10]·3.5H2O, (c) (C4H12N)14[(UO2)10(SeO4)17(H2O)]. (df) Nanotubules are cut into equal parts. The color indicates the chain type used in the construction of the NT. (gi) Schematic representation of the tube. The inner diameters are defined as the distances between the opposite closest oxygen atoms.
Figure 6. (a) Known uranyl-based nanotubules in K5(UO2)3(SeO4)5(NO3)·3.5(H2O), (b) Na(C8H10NO2)7[(UO2)6(SO4)10]·3.5H2O, (c) (C4H12N)14[(UO2)10(SeO4)17(H2O)]. (df) Nanotubules are cut into equal parts. The color indicates the chain type used in the construction of the NT. (gi) Schematic representation of the tube. The inner diameters are defined as the distances between the opposite closest oxygen atoms.
Minerals 10 00659 g006
Figure 7. (a) The distribution of the U-O-T angles in known UO2:TO4 = 3:5 layers and (b) in NT.
Figure 7. (a) The distribution of the U-O-T angles in known UO2:TO4 = 3:5 layers and (b) in NT.
Minerals 10 00659 g007
Figure 8. Fundamental chains with edge-sharing polyhedral in the structures of known layered topologies with UO2:TO4 = 3:5 and known complex NT derivatives.
Figure 8. Fundamental chains with edge-sharing polyhedral in the structures of known layered topologies with UO2:TO4 = 3:5 and known complex NT derivatives.
Minerals 10 00659 g008
Table 1. Crystallographic data and refinement parameters for [C3H9NH+]4[(UO2)3(MoO4)5]. Experiments were carried out at 100 K with Mo Kα radiation on a Bruker Smart DUO, CCD.
Table 1. Crystallographic data and refinement parameters for [C3H9NH+]4[(UO2)3(MoO4)5]. Experiments were carried out at 100 K with Mo Kα radiation on a Bruker Smart DUO, CCD.
Space GroupCc
a(Å)16.768 (6)
b(Å)20.553 (8)
c(Å)11.897 (4)
β, °108.195 (7)
V3)3895 (2)
Dx3.155
θ max (°).1.62–28.00
No. of measured, independent reflections9176/7160
Rint0.0694
Rsigma0.1213
wR10.1182
R10.0520
S0.944
CCDC2014144
Table 2. Selected bond lengths Å in the structure of 1.
Table 2. Selected bond lengths Å in the structure of 1.
U1-O31.765 (13)U2-O11.724 (15)U3-O41.729 (14)
U1-O51.796 (12)U2-O21.767 (15)U3-O81.791 (13)
<Ur1-O>1.780<Ur2-O>1.745<Ur3-O>1.760
U1-O122.299 (12)U2-O132.304 (11)U3-O262.287 (14)
U1-O182.36 7(11)U2-O92.320 (10)U3-O242.332 (12)
U1-O222.371 (11)U2-O232.372 (11)U3-O202.369 (10)
U1-O162.408 (10)U2-O172.375 (11)U3-O112.391 (11)
U1-O72.437 (11)U2-O62.583 (12)U3-O192.497 (12)
<U1-Oeq>2.376<U2-Oeq>2.391<U3-Oeq>2.375
Mo1-O211.688 (12)Mo2-O101.691 (15)Mo3-O151.698 (13)
Mo1-O221.735 (11)Mo2-O61.742 (13)Mo3-O171.774 (10)
Mo1-O91.770 (11)Mo2-O161.766 (10)Mo3-O181.779 (11)
Mo1-O121.776 (12)Mo2-O111.798 (11)Mo3-O71.789 (11)
<Mo1-O>1.742<Mo2-O>1.749<Mo3-O>1.760
Mo4-O141.703 (13)Mo5-O251.666 (16)
Mo4-O191.761 (12)Mo5-O261.702 (14)
Mo4-O231.761 (11)Mo5-O241.765 (13)
Mo4-O201.795 (11)Mo5-O131.767 (12)
<Mo4-O>1.755<Mo5-O>1.725

Share and Cite

MDPI and ACS Style

Nazarchuk, E.; Charkin, D.; Siidra, O.; Kalmykov, S. Organically Templated Layered Uranyl Molybdate [C3H9NH+]4[(UO2)3(MoO4)5] Structurally Based on Mineral-Related Modular Units. Minerals 2020, 10, 659. https://doi.org/10.3390/min10080659

AMA Style

Nazarchuk E, Charkin D, Siidra O, Kalmykov S. Organically Templated Layered Uranyl Molybdate [C3H9NH+]4[(UO2)3(MoO4)5] Structurally Based on Mineral-Related Modular Units. Minerals. 2020; 10(8):659. https://doi.org/10.3390/min10080659

Chicago/Turabian Style

Nazarchuk, Evgeny, Dmitri Charkin, Oleg Siidra, and Stepan Kalmykov. 2020. "Organically Templated Layered Uranyl Molybdate [C3H9NH+]4[(UO2)3(MoO4)5] Structurally Based on Mineral-Related Modular Units" Minerals 10, no. 8: 659. https://doi.org/10.3390/min10080659

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