Structural Evolution from 0D Units to 3D Frameworks in Pb Oxyhalides

Nov 19, 2015 - The properties and structure are compared to other phases crystallizing in the anhydrous PbO–PbX2 (X = F, Cl, Br, I) systems, illustr...
0 downloads 0 Views 6MB Size
Article pubs.acs.org/IC

Structural Evolution from 0D Units to 3D Frameworks in Pb Oxyhalides: Unexpected Strongly Corrugated Layers in Pb7O6Br2 Oleg I. Siidra,*,† Mathias Gogolin,†,‡ Evgeniya A. Lukina,† Houria Kabbour,§ Rimma S. Bubnova,† Olivier Mentré,§ Atali A. Agakhanov,† Sergey V. Krivovichev,† Marie Colmont,§ and Thorsten Gesing‡ †

Department of Crystallography, Saint-Petersburg State University, University emb. 7/9, 199034 St. Petersburg, Russia Chemische Kristallographie fester Stoffe, Universität Bremen, Leobener Straße/NW2, 28359 Bremen, Germany § UCCS, UMR 8181, Université Lille Nord de France, USTL, 59655 Villeneuve d’Ascq, France ‡

S Supporting Information *

ABSTRACT: Novel Pb7O6Br2 (1) lead oxybromide was prepared from Pb oxybromide melt by the “rapid quenching” route. Bonding scheme, thermal expansion, and structural properties were studied. The structural features of this unexpectedly complex phase are described on the basis of lone electron pair stereochemical activity and Pb−Br versus Pb−O bonding scheme. The structure of 1 contains a number of cavities, which can be assigned to the self-containments of the lone electron pairs on Pb2+ cations. “Empty” □Pb4 chains are observed in between of the folding sides of the adjacent strongly corrugated oxocentered [Pb7O6]2+ layers. Highly isotropic thermal expansion of 1 appeared to be unexpected. The possible explanations of such a behavior in 1 are given. The structure of 1 is an interesting example of tetrahedral framework with mixed chemical bonding and is the densest known among Pb oxyhalides with the density of 18.4 tetrahedra/1000 Å3. Current study shows that oxocentered layers derivatives from α-PbO can be very flexible and form rather dense three-dimensional structural topologies. The properties and structure are compared to other phases crystallizing in the anhydrous PbO−PbX2 (X = F, Cl, Br, I) systems, illustrate the complexity of lead oxyhalides, and reveal new and general pathways for the targeted synthesis of new phases with the Pb−O units of desired dimensionality. The indirect gap value of ∼2.04 eV obtained from generalized gradient approximation calculations demonstrates potentially good photocatalytic properties of 1.



INTRODUCTION Pb(II) oxyhalides is a class of inorganic compounds and materials with variety of possible and existing applications.1 They are also important from the viewpoint of environmental issues since tetraethyl lead was widely used as an octane booster to the road vehicle and aviation fuels. As the result, many road side soils are heavily contaminated by these compounds.2 Lead oxyhalide particles are also known as considerable pollutants in the atmosphere.3 The number of “pure” lead oxyhalides4 (i.e., containing no additional cations and anions) is rather limited, but layered Pb oxychloride minerals5 with different anionic and cationic substituents are known as common constituents of oxidation zones of different mineral deposits. Crystal chemistry of similar mineral-like synthetic phases has recently been reviewed.6 It is remarkable that almost no data are known for the Pb oxyhalide with the Pb/X ratio of 7:2 (X = halide ion). For instance, Pb7O6Cl2 was previously reported as natural minerals chubutite and lorettoite, but later both were discredited as mineral species.7 The “Pb7O6Cl2·2H2O” phase was identified as a corrosion product of various lead cable sheaths.2 © XXXX American Chemical Society

Herein we report on the synthesis, structure, bonding scheme, and thermal expansion properties of novel Pb7O6Br2 (1) compound, which was prepared from Pb oxybromide melt. The structural features of this unexpectedly complex phase are described here on the basis of lone electron pair stereochemical activity and Pb−Br versus Pb−O bonding scheme. The properties and structure are compared to other phases crystallizing in the anhydrous PbO−PbX2 (X = F, Cl, Br, I) systems and illustrate the complexity of lead oxyhalides and reveal new and general pathways for the targeted synthesis of new phases with the Pb−O units of desired dimensionality.



EXPERIMENTAL SECTION

Synthesis. Single crystals of 1 were obtained by the solid-state high-temperature “rapid cooling” route6 from the mixture of 0.5000(10) g of PbBr2 (Aldrich, 99.999%) and 0.1000 g of yellow tetragonal α-PbO (Merck, ≥ 99.0%). The reagents were placed in an agate-mortar, wetted with acetone, and ground to a smooth paste until dry. The mixture was then placed into a platinum crucible and kept at Received: September 28, 2015

A

DOI: 10.1021/acs.inorgchem.5b02227 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry 750 °C for 1 h in air, followed by cooling to 25 °C for 20 h. The product consisted of yellow platy crystals (Figure 1S) of 1 with minor impurities of poorly crystallized unidentified phase. The products were manually separated from each other under optical microscope and washed with hexane. The reproducibility of synthesis is 100%. We did not succeed in obtaining Cl-analogue of 1 either in air or in closed inert atmosphere using stoichiometric mixtures. Crystallographic Studies. Single crystal of 1 was mounted on a thin glass fiber for the X-ray diffraction analysis using Bruker APEX II DUO X-ray diffractometer with a microfocus X-ray tube operated with Mo Kα radiation at 50 kV and 40 mA. The data were integrated and corrected for absorption using a multiscan-type model using the Bruker programs APEX and SADABS. More than a hemisphere of Xray diffraction data were collected. Crystallographic information is summarized below.8 Atomic coordinates and additional structural information are provided in the Supporting Information (CIF). High-Temperature X-ray Powder Diffraction Study. Thermal behavior of 1 was studied in air by means of a Rigaku Ultima X-ray diffractometer (Cu Kα radiation) with a high-temperature camera Rigaku HTA 1600. The sample was prepared from heptane’s suspension on a Pt−Rh plate. The temperature step was 15 °C in the range of 25−670 °C. Unit-cell parameters at different temperatures were refined by least-squares methods. Main coefficients of the thermal expansion tensor were determined using linear approximation of temperature dependencies by the ThetaToTensor program.9 Computational Methods. Density functional theory (DFT) simulations were performed using the Vienna ab initio simulation package (VASP).10 The calculations were performed within the generalized gradient approximation (GGA) for the electron exchange and correlation corrections using the Perdew−Wang functional and the frozen core projected wave vector method.11 The geometry optimization (both unit cell parameters and atomic positions) was performed using a plane wave energy cutoff of 550 eV and 42 k points in the irreducible Brillouin zone. The convergence was reached with residual Hellman−Feynman forces on the atoms smaller than 0.03 eV/ Å. Electronic structure calculations were then performed using the optimized structure. In that purpose, a plane wave energy cutoff of 400 eV was used as well as an energy convergence criterion of 1 × 10−6 eV and 250 k points in the irreducible Brillouin zone.

Results of the calculations reveal the indirect bandgap deduced as the maximum of the valence band (VB) and the minimum of the conduction band (CB) located at different k points of the Brillouin zone, between Γ and B and at Γ, respectively. The indirect gap value of ∼2.04 eV obtained from GGA calculations demonstrates potentially good photocatalytic properties of 1. Related PbBi2OX (X = Cl, Br) lead oxyhalides have already been proved as very active for the degradation of methyl orange and stereoselective alkylation under visible light.1c−e The density of states (DOS) and total projected density of states (PDOS) focused in the energy range from −10 to 7 eV (Figure 2). The PDOS are shown for two representative entities: Pb(2) (PbO4 coordination polyhedron) and Pb(3) (PbO3Br2 coordination polyhedron). In the latter, the contribution to the upper VB is extended in the energy range from −4 eV to the Fermi level and composed of the mixing of O p states with Pb s and p states. Very similar features are observed in the same region for Pb(3) also, except for additional contribution from Br 3p states hybridized with the Pb states. Both observations demonstrate the comparable degree of covalent character between Pb−O and Pb−Br. However, the Br 3p states appear slightly more localized than the O 2p states. Such features were also observed, for instance, recently in bismuth oxyhalides BiOX (X = F, Cl, Br).14 The bottom of the CB is predominantly composed of Pb 6p states from 2.4 to ∼5 eV. The predominant presence of O and Br p states at the top of the VB and of Pb p states at the bottom of the CB might refer to a p−p charge transfer upon photonic excitation observed for BiOBr. Concerning the Pb2+ lone electron pair (LP) stereoactivity, the mixing between the anion 2p level and Bi 6s and Bi 6p states at the top of the VB is crucial in producing the asymmetric electronic density15 and in good agreement with our PDOS topology. The lone pair localization for each Pb site was performed using the Verbaere method implemented in the program Hybride.16 The electrostatic E field on the lead atoms is calculated as the induced polarization P = αE roughly equal to the Pb−LP dipolar momentum (−2d), which enables refining of the LP coordinates. We used the Pb2+ Shannon polarizability17 α = 6.58 A3 and the partial charges discussed above for the electric field calculation. Results are provided in Table 1S. Pb−LP distances are rather short (0.16− 0.22 Å) and similar for all of the seven symmetrically independent Pb sites in agreement with the very homogeneous distribution of residual within a highly covalent context. Note, Pb−LP distances should not be taken literally but reflect rather well the lone pair orientation and relative stereoactivity; that is, Pb(1) and Pb(3) show the most pronounced Pb−LP polarization. The structure of 1 contains six oxygen sites. All of the O atoms are tetrahedrally coordinated by Pb atoms, which results in the formation of oxocentered18 OPb4 tetrahedra (Figure 1). The average ⟨O−Pb⟩ distances within the OPb4 tetrahedra are in the range of 2.21−2.57 Å, which is in good agreement with the average value of 2.33 Å derived earlier. There are two symmetrically independent Br sites. Br atoms are coordinated in a typical symmetrical manner. General projection of the structure of 1 is shown in Figure 3a. The OPb4 oxocentered tetrahedra share common edges and corners to produce two types of [Pb5O3]2+ chains depicted in Figure 4. The [Pb5O3]2+ chains shown in Figure 4a were previously observed in a number of compounds with the general formula of [Pb5O3]O(TO4) (T = Cr, S, Mo).19 The second type of the [Pb5O3]2+ chain (Figure 4b) observed in 1 has not been previously



RESULTS AND DISCUSSION The structure of 1 (Figure 1) contains seven symmetrically independent Pb sites. The coordination environments of the Pb atoms are strongly asymmetric in agreement with the presence of stereochemically active “lone pairs” on divalent lead cations. The Pb2+ cations of the Pb(1), Pb(3), Pb(5), Pb(6), and Pb(7) sites have mixed ligand coordination with mO + nBr anions, where m values range from 3 to 4 and n values range from 1 to 2. Pb(2) and Pb(4) are coordinated exclusively by four oxygen atoms at distances shorter than 2.6 Å to form distorted PbO4 tetragonal pyramid with Pb at its apex. These observations raise the question about the relative ionicity of Pb−O bonds versus Pb−Br bonds in 1. We used the Henry’s model12 for determination of partial charges using nonempirical scales of atomic electronegativity and hardness implemented in the program PACHA.12 Results are given in Table 1S and show very similar partial charges for each Pb (ca. 0.44 to 0.48), O (ca. −0.41), and Br (ca. −0.37) site. It does not allow differentiating of the Pb(3), Pb(5), Pb(6), and Pb(7) from Pb(2) and Pb(4) bonding scheme, which shows similar covalent bonds by O and Br anions despite very different electronegativities, that is, χO = 3.44, χBr = 2.96, χPb = 2.33 according to the Pauling scale.13 Pb−X electronegativity differences play in favor of more covalent Pb−Br bonds, but the effective overlap reinforces the Pb−O overlap. The latter was verified by the calculation of the electronic structure. B

DOI: 10.1021/acs.inorgchem.5b02227 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

layer, which correspond to one missing oxygen atom each (Figure 4c). Most of the Pb2+ lone pairs point inside the “empty” Pb4 tetrahedra. The most probable role of the vacancies is to reduce structural strain due to the layer corrugation. The adjacent [Pb7O6]2+ layers in the structure of 1 create one-dimensional channels occupied by Br atoms. The arrangement of the Br atoms in channels corresponds to the fragments excised from an ideal tetragonal halogen layer typical for layered Pb oxyhalides6 (Figure 3e). It is remarkable that corrugated PbO-type layers are also present in the structure of orthorhombic PbO (β-PbO)21 and Ag2Pb8O7Cl4,22 though the degree of distortion is much smaller than that observed in 1. Another interesting aspect of the structure of 1 is the presence of empty □Pb4 chains (Figure 3c) between the folding sides of the adjacent [Pb7O6]2+ layers (Figure 3a). Similar chains formed by Pb4 tetrahedra centered by oxygen atoms were previously described in Pb8O4(Si4O12).23 Figure 1 shows that the size of the Pb4 and OPb4 tetrahedra in the structure of 1 is very similar. In general, the title compound is another novel representative of a Pb oxyhalide structural derivative from the tetragonal modification of PbO (α-PbO).24 Most of the known layered Pb oxyhalides can be considered as Aurivillius-type phases based upon flat PbO cationic layers. In contrast, the structure of 1 can be obtained from the ideal α-PbO structure according the following sequential transformations (Figure 5): (1) incorporation of Br atoms into the interlayer space; (2) corrugation of the PbO layers; (3) removal of some of the oxygen atoms on the folding edges to reduce the stress induced by corrugation and to compensate for the negative charge of the inserted Br− anions. This imaginary scheme generally differs from those that describe the layered Pb oxyhalides with flat layers4,5 (Figure 5).

Figure 1. OPb4 oxocentered (red) and □Pb4 tetrahedra in the structure of 1. Localized Pb2+ lone-pair positions are marked by blue.

described in Pb oxysalts but was found previously in La oxysalts.20 Both types of the [Pb5O3]2+ chains are perpendicular to each other and are interconnected via common Pb atoms to form unusual strongly corrugated [Pb7O6]2+ layers (Figure 3a). There are vacancies near the folding edge of the [Pb7O6]2+

Figure 2. Band structure of 1 in the energy range from −2 to 3 eV. The indirect band gap is highlighted in red (a). Total DOS of 1 and PDOS for Pb(2) and coordinating O(4), O(5), and O(6) anions (PbO4 coordination polyhedron) in the energy range from −10 to 7 eV (b). PDOS for Pb(3) and coordinating O(2), O(3), and Br(2) anions (PbO2Br2 coordination polyhedron) in the energy range from −10 to 7 eV (c). C

DOI: 10.1021/acs.inorgchem.5b02227 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 3. General projection of the crystal structure of 1 along the c axis (a) (gray spheres = Pb; red spheres = O; khaki spheres = Br; localized Pb2+ lone pair positions = blue). [Pb7O6]2+ layers can be symbolized (b) as built from two types of chains depicted in Figure 4. Empty [□Pb4] tetrahedral chains in between corrugated layers (c). Enlarged fragment (bulb) of the corrugated [Pb7O6]2+ layer in 1 (d). Pseudotetragonal arrangement of Br atoms within the bulbs (e). See text for details.

10−6 K−1, μ = (α33ĉ) = 22° at 520 °C. Highly isotropic thermal expansion (Figure 6; Figure 3S) of 1 appeared to be unexpected. Anisotropic thermal expansion is generally favored for layered crystal structures especially in the case of oblique crystals due to additional degree of freedom.25 The possible explanation of such a behavior in 1 is the presence of Pb···Pb interactions inside the empty □Pb2 chains (Figure 3c). It is important to mention that isotropic thermal expansion is also a typical feature for α-PbO.26 α-PbO, known as mineral litharge, also does not cleave along (001), despite the presence of the PbO layers parallel to this plane in the structure, which was explained by the presence of weak Pb···Pb attractive interactions confirmed by NMR27 studies. The NMR appeared to be a useful tool for the evaluation of lone electron pair stereochemical activity on Pb2+.28 The Pb···Pb distance between the layers in α-PbO is ∼3.65 Å and is close to the values of ∼3.70 Å observed in the empty Pb4 tetrahedra in 1. The role of similar cation−cation closed-shell interactions seems to be very important in stabilizing some structure types.29 All the above-mentioned peculiarities of the structure of 1 allow us to consider it as an interesting example of tetrahedral framework with mixed chemical bonding.

Evolution of powder diffraction patterns of 1 with the increasing temperature is shown in Figure 2S. The crystal structure of 1 is basically unchanged up to ca. 590 °C, when diffraction maxima start to disappear. The compound 1 decomposes into PbBr2, β-PbO, Pb2PtO4 (which forms as a result of reaction with Pt plate), and an unidentified phase. Temperature dependencies of the unit-cell parameters (Figure 3S) can be described by the following polynomial functions: a t = 13.4762 + (0.3087 × 10−3)t bt = 12.4863 + 0.2675 × 10−3)t

ct = 8.099 + (0.1896 × 10−3)t βt = 106.8897 + (0.0533 × 10−3)t

Vt = 1303.916 + (89.081 × 10−3)t

where t is a temperature in degrees Celsius. When heated, the unit cell expands almost isotropic along all crystallographic axes: α11 = 23, α22 = αb = 20, α33 = 19 × 10−6 K−1, μ = (α33ĉ) = 3° at room temperature and α11 = 25, α22 = αb = 22, α33 = 23 × D

DOI: 10.1021/acs.inorgchem.5b02227 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. [Pb7O6]2+ strongly corrugated layer in 1 can be split into two mutually perpendicular [Pb5O3]2+ oxocentered chains of different architecture (a, b). Polyhedral representation of [Pb7O6]2+ layer fragment showing interconnection of [Pb5O3]2+ chains along the b axis (c).

Figure 5. Two principal pathways for structural transformations of α-PbO in Pb oxyhalide melt into Aurivillius-type phases by spreading and shift of PbO layers with subsequent incorporation of halogen atoms into interlayer (above) and strong corrugation of PbO layers with crimped halogen atoms in the channels observed in 1 (below).

The following pure (i.e., without additional cations and anions) Pb oxyhalides are known to date: Pb[Pb8O4]X10,4d Pb 0.16 [Pb 2 O]X 2.32 ,4b,1j [Pb 3 O 2 ]X 2 , 4a,30 Pb[Pb 30 O 22 ]X 18 , 4g [Pb13O10]X6,4c,e and [Pb7O6]X2 (X = Cl, Br, I) reported herein (Figure 7). The [PbmOn]z+ units formed by edge and corner sharing of OPb4 oxocentered tetrahedra demonstrate all possible dimensionalities starting from zero-dimensional (0D) to three-dimensional (3D). The structural architecture of the oxocentered units in Pb oxyhalides correlates with the O/X ratio. The phase with equal 1:1 ratio corresponds to [Pb3O2] X24a,30 with the crystal structure based upon double [Pb3O2]2+

Figure 6. Pole figures of the thermal expansion coefficients of 1 in ca (α11−α33 section) and b−α11 (α11−α22 section) planes at 25 °C.

E

DOI: 10.1021/acs.inorgchem.5b02227 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

structural architecture and physical properties of 1. The probable explanation for this effect is the electrophilicity of lone electron pairs that can be understood in terms of the hard−soft acid−base (HSAB) concept.32 Current study also shows that oxocentered layers derivatives from α-PbO can be very flexible and form rather dense 3D structural topologies. “Rapid quenching” route is very fruitful for the synthesis of well-crystallized Pb oxyhalides. This observation makes the latter a viable exploration venue for the preparation of novel and unusual structural architectures with possible different applications.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.5b02227. Three-dimensional perspective plot showing all diffractograms over 8−59° 2Θ with increasing temperature, ellipsoids of thermal expansion at different temperatures, comparison of the calculated and experimental powder patterns, the temperature dependences of the unit cell parameters and volume, table with atomic coordinates, calculated lone pair coordinates, bond-valence sums and displacement parameters, and photo of crystals of 1 under optical microscope. (PDF) File of X-ray crystallographic data in CIF format. (CIF)



Figure 7. Structural diversity of known Pb oxyhalides (X = Cl, Br, I) without additional anions and cations. The [PbmOn]z+ units formed by edge- and corner-sharing OPb4 tetrahedra (shown in red) vary from 0D to 3D structural architectures. Dimensionality of structural motif correlates with the O:X (X = Cl, Br, I) ratio. Density of lead oxyhalide structures (calculated as a number of OPb4 tetrahedra per 1000 Å3) increase (green arrow) with the increase of the oxygen content. Two projections of the structure of [Pb13O10]X6 are shown. Crystal structure of 1 (below) appears to be the densest Pb oxyhalide known to date and can be considered as a framework with Pb···Pb interactions in the regions marked by blue.

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by St. Petersburg State Univ. through the internal Grant Nos. 3.38.238.2015 and 3.50.2099.2013 (A.A.A.), RFBR 15-35-20632, and Russian President Grant No. MK-3756.2014.5 (O.I.S.). Support by the X-ray Diffraction Resource Centre of Saint-Petersburg State Univ. is gratefully acknowledged.

chains. This phase is also the most stable Pb oxyhalide in PbO− PbX2 (X = Cl, Br, I) anhydrous system and often forms as a byproduct in more complex systems with additional cations and anions. Further reduction of the oxygen content leads to the formation of [Pb2O]2+ single chains and [Pb8O4]8+ finite stella quadrangula clusters. The increase of the oxygen content relative to the X content finally results in the formation of 3D frameworks. There is also an obvious trend in the increase of the topological density of [PbmOn]z+ moieties defined in the manner similar to the framework density (FD) in zeolites31 and calculated as the number of OPb4 tetrahedra per 1000 Å3. The structure is 1 is the densest known among Pb oxyhalides with the density of 18.4 tetrahedra/1000 Å3. It can be compared with the density of α-PbO equal to 24.6. Another framework Pb oxyhalide structure of [Pb13O10]X6 is porous and less dense than 1.



REFERENCES

(1) (a) Matsumoto, H.; Miyake, T.; Iwahara, H. Mater. Res. Bull. 2001, 36, 1177−1184. (b) Sigman, M. B., Jr.; Korgel, B. A. J. Am. Chem. Soc. 2005, 127, 10089−10095. (c) Shan, Z.; Lin, X.; Liu, M.; Ding, H.; Huang, F. Solid State Sci. 2009, 11, 1163−1169. (d) Füldner, S.; Pohla, P.; Bartling, H.; Dankesreiter, S.; Stadler, R.; Gruber, M.; Pfitzner, A.; König, B. Green Chem. 2011, 13, 640−643. (e) Cherevatskaya, M.; Neumann, M.; Füldner, S.; Harlander, C.; Kümmel, S.; Dankesreiter, S.; Pfitzner, A.; Zeitler, K.; König, B. Angew. Chem., Int. Ed. 2012, 51, 4062−4066. (f) Siidra, O. I.; Krivovichev, S. V.; Armbruster, T.; Depmeier, W. Inorg. Chem. 2007, 46, 1523−1525. (g) Shu, J.; Ma, R.; Shao, L.; Shui, M.; Wang, D.; Wu, K.; Long, N.; Ren, Y. Electrochim. Acta 2013, 102, 381−387. (h) Aliev, A.; Olchowka, J.; Colmont, M.; Capoen, E.; Wickleder, C.; Mentre, O. Inorg. Chem. 2013, 52, 8427−8435. (i) Noren, L.; Tan, E. S. Q.; Withers, R. L.; Sterns, M.; Rundlof, H. Mater. Res. Bull. 2002, 37, 1431−1442. (j) Zhang, H.; Zhang, M.; Pan, S.; Dong, X.; Yang, Z.; Hou, X.; Wang, Z.; Chang, K. B.; Poeppelmeier, K. R. J. Am. Chem. Soc. 2015, 137, 8360−8363. (2) Edwards, R.; Gillard, R. D.; Williams, P. A.; Pollard, A. M. Mineral. Mag. 1992, 56, 53−65.



CONCLUSION The structure of 1 contains a number of cavities, which can be assigned to the self-containments of the lone electron pairs on Pb2+ cations. Formation of such cavities is typical for the “lone pair cations” of such metals as Tl, Pb, and Bi and controls F

DOI: 10.1021/acs.inorgchem.5b02227 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (3) (a) Post, J.; Buseck, P. R. Environ. Sci. Technol. 1985, 19, 682− 685. (b) Cziczo, D. J.; Stetzer, O.; Worringen, A.; et al. Nat. Geosci. 2009, 2, 333−336. (4) (a) Siidra, O. I.; Krivovichev, S. V.; Armbruster, T.; Depmeier, W. Z. Kristallogr. 2008, 223, 204−211. (b) Siidra, O. I.; Krivovichev, S. V.; Depmeier, W. Dokl. Phys. Chem. 2007, 414, 128−131. (c) Siidra, O. I.; Krivovichev, S. V.; Depmeier, W. Geol. Ore Deposits 2007, 49, 827− 834. (d) Keller, H. L. Angew. Chem. 1983, 95, 318−319. (e) Riebe, H. J.; Keller, H. L. Z. Anorg. Allg. Chem. 1989, 571, 139−147. (f) Kramer, V.; Post, E. Mater. Res. Bull. 1985, 20, 407−412. (g) Krivovichev, S. V.; Siidra, O. I.; Nazarchuk, E. V.; Burns, P. C.; Depmeier, W. Inorg. Chem. 2006, 45, 3846−3848. (h) Aurivillius, B. Chem. Scripta 1976, 10, 156− 158. (5) (a) Humphreys, D. A.; Thomas, J. H.; Williams, P. A.; Symes, R. F. Mineral. Mag. 1980, 43, 901−904. (c) Cooper, M. A.; Hawthorne, F. C. Am. Mineral. 1994, 79, 550−554. (d) Welch, M. D.; Criddle, A. J.; Symes, R. F. Mineral. Mag. 1998, 62, 387−393. (e) Welch, M. D.; Cooper, M. A.; Hawthorne, F. C.; Criddle, A. J. Am. Mineral. 2000, 85, 1526−1533. (f) Bonaccorsi, E.; Pasero, M. Mineral. Mag. 2003, 67, 15−21. (g) Krivovichev, S. V.; Turner, R.; Rumsey, M.; Siidra, O. I.; Kirk, C. A. Mineral. Mag. 2009, 73, 75−89. (h) Siidra, O. I.; Krivovichev, S. V.; Turner, R. W.; Rumsey, M. S.; Spratt, J. Am. Mineral. 2013, 98, 248−255. (i) Siidra, O. I.; Krivovichev, S. V.; Turner, R. W.; Rumsey, M. S.; Spratt, J. Am. Mineral. 2013, 98, 256− 261. (j) Turner, R. W.; Siidra, O. I.; Krivovichev, S. V.; Stanley, C. J.; Spratt, J. Mineral. Mag. 2012, 76, 1247−1255. (6) Siidra, O. I.; Zinyakhina, D. O.; Zadoya, A. I.; Krivovichev, S. V.; Turner, R. W. Inorg. Chem. 2013, 52, 12799−12805. (7) White, J. S. Am. Mineral. 1979, 64, 1303−1305. (8) Crystal data of 1: A total of 9736 reflections of a yellow plate crystal with dimensions of 0.09 × 0.09 × 0.01 mm3, monoclinic, P21/c (No. 14), a = 13.494(2) Å, b = 12.483(2) Å, c = 8.1028(15) Å, β = 107.173(4), V = 1304.0(4) Å3, Z = 4, ρcalc = 8.690 g·cm−3, and μ(Mo Kα) = 96.162 mm−1. The following twinning matrix was applied during the refinement: [−10−1 0−10 001] The final cycles of refinement converged at R1 = 0.0425, wR2 = 0.109, and GOF = 1.030 with 149 parameters, ρmax,min = +4.63/−4.46 e·Å−3. Qualitative electron microprobe analysis (Hitachi TM-3000) of 1 revealed no other elements, except Pb and Br, with the atomic number greater than 11 (Na). (9) Firsova, V. A.; Bubnova, R. S.; Filatov, S. K. Program for the Thermal Expansion Tensor Determination for Crystalline Materials; Institute of Silicate Chemistry of Russ. Acad. Sci.: St. Petersburg, Russia, 2011. (10) Kresse, G.; Furthmüller, J. Vienna Ab-initio Simulation Package (VASP); Institut für Materialphysik: Vienna, 2012. (11) (a) Perdew, J. P.; Wang, Y. Phys. Rev. B: Condens. Matter Mater. Phys. 1992, 45, 13244. (b) Kresse, G.; Joubert, D. Phys. Rev. B: Condens. Matter Mater. Phys. 1999, 59, 1758. (12) (a) Henry, M. ChemPhysChem 2002, 3, 561−569. (b) Henry, M. In Advances in Quantum Chemical Bonding Structures; Putz, M. V., Ed.; Transworld Research Network: India, 2008; pp 153−211. (13) Pauling, L. The Nature of the Chemical Bond, 3rd ed; Cornell Univ: Ithaca, NY, 1960. (14) Zhao, L.; Zhang, X.; Fan, C.; Liang, Z.; Han, P. Phys. B 2012, 407, 3364−3370. (15) Walsh, A.; Watson, G. W.; Payne, D. J.; Edgell, R. G.; Guo, J. H.; Glans, P. A.; Learmonth, T.; Smith, K. E. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 73, 235104. (16) (a) Verbaere, A.; Marchand, R.; Tournoux, M. J. Solid State Chem. 1978, 23, 383−390. (b) Morin, G. E.; Wallez, G.; Jaulmes, S.; Couturier, J. C.; Quarton, M. J. Solid State Chem. 1998, 137, 283−288. (17) Shannon, R. D. J. Appl. Phys. 1993, 73, 348−366. (18) (a) Krivovichev, S. V.; Mentré, O.; Siidra, O. I.; Colmont, M.; Filatov, S. K. Chem. Rev. 2013, 113, 6459−6535. (b) Siidra, O. I.; Krivovichev, S. V.; Filatov, S. K. Z. Kristallogr. - Cryst. Mater. 2008, 223, 114−126. (c) Thirumurugan, A.; Sanguramath, R. A.; Rao, C. N. R. Inorg. Chem. 2008, 47, 823−831.

(19) (a) Steele, I. M.; Pluth, J. J. J. Electrochem. Soc. 1998, 145, 528− 533. (b) Krivovichev, S. V.; Armbruster, T.; Depmeier, W. J. Solid State Chem. 2004, 177, 1321−1332. (c) Vassilev, P.; Nihtianova, D. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 1998, C54, 1062−1068. (20) Carré, D.; Guittard, M.; Jaulmes, S.; Mazurier, A.; Palazzi, M.; Pardo, M. P.; Laruelle, P.; Flahaut, J. J. Solid State Chem. 1984, 55, 287−292. (21) Kay, M. I. Acta Crystallogr. 1961, 14, 80−81. (22) Langecker, C.; Keller, H. L. Z. Anorg. Allg. Chem. 1994, 620, 1229−1233. (23) Dent Glasser, L. S.; Howie, R. A.; Smart, R. M. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1981, B37, 303−306. (b) Kato, K. Acta Crystallogr., Sect. B: Struct. Crystallogr. Cryst. Chem. 1980, B36, 2539−2545. (24) Boher, P.; Garnier, P.; Gavarri, J. R.; Hewat, A. W. J. Solid State Chem. 1985, 57, 343−350. (25) Filatov, S. K. Phys. Status Solidi B 2008, 245, 2490−2496. (26) Sorrell, C. A. J. Am. Ceram. Soc. 1970, 53, 641−644. (27) Gabuda, S. P.; Kozlova, S. G.; Terskikh, V. V.; Dybowski, C.; Neue, G.; Perry, D. L. Chem. Phys. Lett. 1999, 305, 353−358. (28) Greer, B. J.; Michaelis, V. K.; Katz, M. J.; Leznoff, D. B.; Schreckenbach, G.; Kroeker, S. Chem. - Eur. J. 2011, 17, 3609−3618. (29) Pyykkö, P. Chem. Rev. 1997, 97, 597−636. (30) Kramer, V.; Post, E. Mater. Res. Bull. 1985, 20, 407−412. (31) Bnmner, G. O.; Meier, W. M. Nature 1989, 337, 146−147. (32) Siidra, O. I.; Zenko, D. S.; Krivovichev, S. V. Am. Mineral. 2014, 99, 817−823.

G

DOI: 10.1021/acs.inorgchem.5b02227 Inorg. Chem. XXXX, XXX, XXX−XXX