Article pubs.acs.org/IC
Two-Dimensional Inorganic Cationic Network of Thorium Iodate Chloride with Unique Halogen−Halogen Bonds Huangjie Lu,†,‡,⊥ Yaxing Wang,†,‡,⊥ Congzhi Wang,§,⊥ Lanhua Chen,†,‡ Weiqun Shi,*,§ Juan Diwu,†,‡ Zhifang Chai,†,‡ Thomas E. Albrecht-Schmitt,∥ and Shuao Wang*,†,‡ †
School for Radiological and Interdisciplinary Sciences and ‡Collaborative Innovation Center of Radiation Medicine of Jiangsu Higher Education Institutions, Soochow University, 199 Ren’ai Road, Suzhou 215123, China § Laboratory of Nuclear Energy Chemistry and Key Laboratory for Biomedical Effects of Nanomaterials and Nanosafety, Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, China ∥ Department of Chemistry and Biochemistry, Florida State University, Tallahassee, Florida 32306, United States S Supporting Information *
ABSTRACT: A unique two-dimensional inorganic cationic network with the formula [Th3O2(IO3)5(OH)2]Cl was synthesized hydrothermally. Its crystal structure can best be described as positively charged slabs built with hexanuclear thorium clusters connected by iodate trigonal pyramids. Additional chloride anions are present in the interlayer spaces but surprisingly are not exchangeable, as demonstrated by a series of CrO42− uptake experiments. This is because all chloride anions are trapped by multiple strong halogen− halogen interactions with short Cl−I bond lengths ranging from 3.134 to 3.333 Å, forming a special Cl-centered trigonalpyramidal polyhedron as a newly observed coordination mode for halogen bonds. Density functional theory calculations clarified that electrons transformed from central Cl atoms to I atoms, generating a halogen−halogen interaction energy with a value of about −8.3 kcal mol−1 per Cl···I pair as well as providing a total value of −57.9 kcal mol−1 among delocalized halogen− halogen bonds, which is a new record value reported for a single halogen atom. Additional hydrogen-bonding interaction is also present between Cl and OH, and the interaction energy is predicted to be −8.1 kcal mol−1, confirming the strong total interaction to lock the interlayer Cl anions.
1. INTRODUCTION Most extended structures of inorganic minerals in nature are typically anionic in which the charge is balanced by exchangeable cations (e.g., alkali-metal cations) that fill in the void space within the anionic framework, leading to an overall neutral charge. Only a handful of inorganic cationic framework materials exist where the charge is balanced by unbound/ weakly bound anions.1,2 These compounds have been intensively studied as a new type of anion-exchange material with an applications of removing key anionic pollutants in the environment.3 The best known cationic materials that exist in nature are hydrotalcite clays, also known as the layered double hydroxides, which are based on the trivalent-ion-substituted brucite structure rendering cationic layers.4 In recent years, a series of cationic extended materials were also prepared by several groups, mostly utilizing weakly coordinated anions as structure-directing templates.5 In addition, a family of layered lanthanide hydroxyhalide intercalated hosts and transitionmetal-based cationic networks reported recently also exhibit anion-exchange properties or selective catalytic activities.6 More examples of inorganic cationic network materials containing © XXXX American Chemical Society
unbound/weakly bound halogen anions are francisite compounds Cu3Bi(SeO3)2O2X (X = Cl, Br, and I) and Te4M3O15· Cl (M = Nb5+ or Ta5+).7 Six years ago we prepared an inorganic supertetrahedral cationic framework material, [ThB5O6(OH)6][BO(OH)2]· 2.5H2O (NDTB-1), which can selectively remove TcO4− anions from nuclear waste streams.2a,b One of the intrinsic reasons for formation of the cationic framework is that the structure is based on a tetravalent Th cation with high positive formal charge coordinated by condensed hydroborate ligands with reduced negative charge. During the continued course of searching new inorganic cationic materials based on high-valent metal centers and large coordinating ligands with low negative charge, a highly unusual thorium iodate chloride with a cationic network was discovered with the formula [Th3O2(IO3)5(OH)2] Cl (abbreviated as ThIOCl). This crystalline compound contains strong halogen−halogen interactions, forming a special Cl-centered trigonal-pyramidal polyhedron as a newly Received: May 6, 2016
A
DOI: 10.1021/acs.inorgchem.6b01110 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry
range of 200−1000 cm−1 were collected with crystals placed on quartz slides. Fourier Transform Infrared (FTIR) Spectrum. The IR spectrum of ThIOCl was recorded using a FTIR spectrometer instrument in the range of 400−4000 cm−1 (Thermo Nicolet 6700 spectrometer). Energy-Dispersive Spectroscopy (EDS) Analysis. Scanning electron microscopy (SEM) images and EDS data were recorded on a FEI Quanta 200FEG scanning electron microscope, with the energy of the electron beam being 30 keV. Samples were mounted directly on the carbon conductive tape with gold coating. Theoretical Methods. Density functional theory (DFT) calculations were carried out with the Gaussian 098 program package using the M06-2X9 functional. The LANL2DZ basis sets were used for the I atom, and the 6-311+G(d,p) basis sets were adopted for the H, O, and Cl atoms. On the basis of the model fragment from the crystal structure of Th2O3(IO3)5Cl(OH)2, single-point calculations were carried out at the M06-2X/RECP/6-311+G(d,p) level of theory. For the halogen−halogen and hydrogen-bonding interaction energy analyses, the basis set superposition error (BSSE)10 was taken into account by the counterpoise method. The quantum theory of atoms in molecules (QTAIM)11 analysis was performed with the Multiwf n 3.3.8 package12 at the same level of theory.
observed coordination mode for halogen bonds, further leading to its completely unexchangeable nature.
2. EXPERIMENTAL SECTION Caution! Th-232 used in this study is an α emitter with the daughter of radioactive Ra-228. All of the thorium compounds used and investigated were operated in an authorized laboratory designed for actinide element studies. Standard protections for radioactive materials should be followed. Synthesis of [Th3O2(IO3)5(OH)2]Cl (ThIOCl). Th(NO3)4·6H2O (55.8 mg, 0.1 mmol), HIO3 (0.0176 g, 0.1 mmol), KCl (0.0149 g, 0.2 mmol), and deionized water (2 mL) were loaded into a 10 mL autoclave. The autoclave was sealed, heated to 230 °C in a box furnace for 3 days, and cooled to room temperature for 2 days. The product was washed with a deionized water and ethanol solution, and the colorless hexagon-shaped crystals suitable for X-ray structural analysis were collected. All reagents were used as received from commercial suppliers without further purification. X-ray Crystallography. Data collection was performed on a Bruker D8-Venture diffractometer with a Turbo X-ray source (Mo Kα radiation, λ = 0.71073 Å) adopting the direct-drive rotating-anode technique and a CMOS detector at room temperature. The data frames were collected using the program APEX2 and processed using the program SAINT routine in APEX2. The structures were solved by direct methods and refined by the full-matrix least squares on F2 using the SHELXTL-2014 program. All non-H atoms were refined with anisotropic displacement parameters. Crystallographic and refinement details are summarized in Table 1.
3. RESULTS AND DISCUSSION Structure Description. ThIOCl crystallizes in centrosymmetric trigonal space group R3̅. A view of the structure is shown in Figure 1. The overall structure can be divided into
Table 1. Crystallographic Data for ThIOCl formula Mr (g mol−1) cryst syst space group a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3) Z Dc (g cm−3) μ (mm−1) F(000) T (K) GOF on F2 R1,a wR2b [I > 2σ(I)] R1,a wR2b (all data) a
[Th3O2(IO3)5(OH)2]Cl 1670.07 trigonal R3̅ 9.8117(3) 9.8117(3) 33.049(2) 90 90 120 2755.3(2) 6 6.039 32.869 4224.0 298 0.853 0.0236, 0.0521 0.0240, 0.0522
R1 = ∑||Fo| − |Fc||/∑|Fo|. bwR2 = [∑w(Fo2 − Fc2)2/∑w(Fo2)2]
1/2
Figure 1. Coordination mode of thorium (a), hexanuclear clusters achieved by μ3-hydroxo (b), and a two-dimensional cationic network structure achieved by iodate ligands (c). Views of the cationic network structures of [Th3O2(IO3)5(OH)2]+ along the c axis (d) (color scheme: Th, green; O, red; I, purple) and [Th3O2(IO3)5 (OH)2]Cl along the a axis (e).
.
Powder X-ray Diffraction (PXRD). PXRD patterns were collected from 5° to 50°, with a step of 0.02°, using a Bruker D8 Advance X-ray diffractometer with Cu Kα radiation (λ = 1.54056 Å) equipped with a Lynxeye one-dimensional detector. Thermal Stability Analysis. Thermogravimetric analysis (TGA) was carried out on a NETZSCH STA 449F3 instrument in the range of 30−900 °C under a nitrogen flow at a heating rate of 10 °C min−1. Variable-temperature PXRD analysis was carried out on a Bruker D8 Advance X-ray diffractometer under a nitrogen flow in the range of 25−700 °C. Solid-State UV−Vis−Near-IR (NIR) Absorption and Raman Spectral Measurements. All spectra were recorded on a Craic Technologies microspectrophotometer. For UV−vis−NIR spectra, crystals were placed on quartz slides, and data were collected after optimization of the microspectrophotometer. Raman spectra in the
[Th3O2(IO3)5(OH)2]+ cationic layers and interlayer Cl ions. The layer is perpendicular to the c axis and built with Th-based clusters further bridged by IO3−. Each ThIV center is eightcoordinated, adopting a coordination geometry of a standard D4d square antiprism (Figure 1a), which is bound by four μ3oxo groups (O2, O3, O4, and O5) from four different IO3− ligands and four μ3-hydroxo groups (O1, O7A, O7B, and O7C).13 Six ThIV ions are bridged by four μ3-oxo and four μ3hydroxo groups, forming a hexanuclear thorium core (Figure 1b). This type of cluster was observed in several hydrolysis products of tetravalent actinides.14 I(2)O3− and I(3)O3− provided all three O atoms serving as bridges to connect B
DOI: 10.1021/acs.inorgchem.6b01110 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry adjacent thorium clusters, while I(1)O3− supplies two O atoms to connect two Th metal centers within the thorium cluster (Figure 1d,e). One of the key features for the structure of ThIOCl is that Cl ions are orderly distributed in the interlayer spaces between the neighboring [Th3O2(IO3)5(OH)2]+ layers. The interlayer distance of the cationic layers is about 1.7 Å, and Cl ions do not covalently bond to any network atoms at first glance. In addition, elemental analysis determined by SEM and EDS, as shown in Figure S1 and Table S1, gave an atomic ratio of 15.31:26.87:6.47 for Th/I/Cl, which is consistent with the crystallographic results. Stability Measurements. Pure inorganic materials, such as zeolite or zeolite analogue materials, often possess high thermal stability and hydrolytic resistance.15 Variable-temperature PXRD (VT-PXRD) patterns and TGA were performed to investigate the thermal stability of ThIOCl. As shown in Figures 2 and 3, the bulk of ThIOCl can maintain structural
Figure 4. Simulated PXRD patterns for ThIOCl and experimental PXRD patterns for ThIOCl after it was soaked in aqueous solutions at pH 0, 3, 6, 9, 12, and 14 for 24 h, respectively.
after they were soaked in aqueous solutions with different pH values (0, 3, 6, 9, 12, and 14) for 24 h, respectively. The pattern is consistent with that calculated from the single-crystal structure, and the results show that the crystal began to crumble only in a strong alkaline environment. Anion-Exchange Test. Considering the decent stability as well as cationic nature of the extended structure, ThIOCl should be considered as a potential anion-exchange material. A preliminary anion-exchange experiment was implemented by immersing crystals into a potassium chromate solution (0.1 mol L−1) at room temperature. However, as shown in Figure S3, the crystal remains robust and colorless after 12 h of the exchange experiment. The UV−vis absorption spectra of immersed crystals exhibit no characteristic peaks of CrO42− at 425 nm. The EDS analysis shown in Figure S4 also confirms that no anion exchange occurs. Additionally, the interlayer Cl anions are completely unexchangable regardless of the concentration of potassium chromate at room temperature. Furthermore, even in a concentrated CrO42− solution under elevated temperature (e.g., 60 °C), the crystals of ThIOCl can still preserve its crystallinity and unexchangeable nature, further demonstrating the robustness of interlayer Cl anions trapped in this cationic network compound (PXRD shown in Figure S5 and EDS in Figure S6). Another anion-exchange experiment was implemented by immersing crystals into different NaF solutions (0.1 and 1 mol L−1) at room temperature for 12 h. As shown in Figure S8, the structure of ThIOCl crumbled with the 1 mol L−1 F− concentration, probably converting to thorium fluoride because of the strong affinity between F− and Th4+. However, as shown in Figures S9 and S10, no obvious ion exchange can be observed in the low concentration solution (0.1 mol L−1) because no characteristic peak of the F element was shown in the EDS spectra. The EDS spectra collected on the soaked crystals show the molar ratio of Th/I/Cl to be ca. 3:5:1, which is almost identical with that of the original sample. In addition, as shown in Figure 5, elemental mapping analysis16 showed that Th, I, O, and Cl atoms are present in the hexagon crystal, while no F signal can be probed. Therefore, the structure of ThIOCl is not even able to exchange with smaller F− anions, surprisingly. Halogen-Bonding Description. The anion-unexchangeable phenomenon is attributed to the deep structural features in ThIOCl. It is found that the coordination spheres of interlayer
Figure 2. VT-PXRD patterns for ThIOCl.
Figure 3. TGA (black) and differential scanning calorimetry (blue) curves for ThIOCl.
integrity up to 450 °C. Figure 2 shows that the main peak (2θ ≈ 24.35°) correponding to the (2, −1, 6) plane disappeared until 450 °C. TGA was carried out by the as-synthesized dried samples, which were heated at a constant rate of 10 K min−1 in nitrogen from 30 to 900 °C. ThIOCl shows high thermal stability and no obvious weight loss occurred until 450 °C, further demonstrating the robustness of the cationic materials. In addition, the structure shows excellent hydrolytic stability because it is stable in aqueous solutions over a wide pH range from 0 to 12 (Figure 4). PXRD was performed on the samples C
DOI: 10.1021/acs.inorgchem.6b01110 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry
Figure 6. Views of the fragment models of ThIOCl, containing all types of interactions between I···I, I···Cl, and Cl···O1(μ3-hydroxo) (a), Cl− coordination spheres containing all I···I and I···Cl interactions (b), all I···Cl interactions (c), and only strong I···Cl interactions (d), respectively.
is 3.194 Å, suggesting that an intermediate hydrogen bond is also involved in the coordination sphere of Cl−. Therefore, the Cl ions are effectively locked within the interlayer spaces, giving rise to the completely unexchangeable nature of ThIOCl. The halogen bonding effect can also be probed by the change of I−O vibrations. As shown in Figure 7, the IR spectrum of
Figure 5. SEM image and EDX elemental maps of a ThIOCl crystal: (a) Crystalline SEM image. Elemental maps of (b) Th (yellow), (c) I (purple), (d) O (red), (e) Cl (dark blue), and (f) F (green) are shown in the figure.
Cl ions are rather complicated because they are bound to multiple neighboring iodate ions as well as a hydroxo group, indicating that both halogen−halogen interactions and hydrogen bonds exist in the structure. As a versatile linkage, the halogen−halogen bond has rapidly emerged in various supermolecular assemblies, which could find potential applications in anion recognition, magnetic, optical, and catalysis materials.17,18 However, it is seldom reported and often not recognized in purely inorganic compounds. Considering the sum of the van der Waals radii of Cl, I, and O atoms,19 there are seven halogen−halogen bonds with three different Cl···I distances and one hydrogen bond around each Cl− anion (Figure 6). Shown as blue dotted lines in Figure 6c, the three Cl···I1A distances are 3.601 × 3 Å, indicating relatively weak interactions. The O2−I1A···Cl angle of 69.600°, O3−I1A···Cl angle of 163.266°, and O6−I1A···Cl angle of 73.621° suggest that these halogen bonds are at a intermediate state between type I and II contacts.17a In comparison, the Cl··· I1B distances of 3.134 × 3 Å and the Cl···I2 distance of 3.333 Å are significantly shorter (note that I1A and I1B are symmetryrelated). These four bonds result in a unique trigonal-pyramidal coordination geometry for the central Cl ion (Figure 6d). The O6−I1B···Cl angle of 166.859°, O3−I1B···Cl angle of 76.492°, and O2−I1B···Cl angle of 91.898° indicate that the contacts of the halogen and anion are roughly perpendicular to the O2− I1B axis.17a The O3−I2···Cl angles of 118.282°, 118.264°, and 118.295° are approximately equal, indicating that these halogen−halogen interactions are of the pseudo-type II contact. Note that, in this case, the halogen interaction is between the anions Cl− and IVO3−, which deviates from the common cases between halides. These strong halogen bonds can therefore be attributed to the participatation of the lone electron pair of IV in bonding, forming a special type of halogen bonding and anion interaction.18b In addition, the distance of Cl···O1 (μ3-hydroxo)
Figure 7. IR and Raman spectra of compound ThIOCl.
ThIOCl exhibits a series of absorption bands from 400 to 2000 cm−1, where the absorption bands at 498 and 542 cm−1 should be attributed to the symmetric I−O stretching (ν2), the features at 673 and 719 cm−1 can be assigned to the symmetric I−O stretching (ν1), and the features at 795 and 823 cm−1 are the asymmetric I−O stretching (ν3).20 In addition, as complementary information, the Raman spectrum was also recorded from 200 to 1000 cm−1, showing four vibrational modes for IO3− in the range from 300 to 1000 cm−1, as depicted in the inset of Figure 7. The corresponding bands are 317 cm−1 (ν4), 333 cm−1 (ν4), 432 cm−1 (ν2), 729 cm−1 (ν1), 772 cm−1 (ν1), 797 cm−1 (ν3), and 837 cm−1 (ν3).20,21 In comparison to those for a free IO3− anion with C3ν symmetry, the four fundamental vibration modes all shift to lower energy regions with shift values of about 26 and 32 cm−1 (ν1), 34 cm−1 (ν2), 37 and 3 cm−1 (ν3), and 5 and 23 cm−1 (ν4), which are also demonstrated by the DFT calculations discussed in the next part.20 This is because formation of the Cl−···IO3− interaction D
DOI: 10.1021/acs.inorgchem.6b01110 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry reduces the electron density of I−O, similar to the cases of vibration peak shifting induced by the hydrogen-bonding effect.22 Quantum-Chemical Calculations. Quantum-chemical calculations were carried out to investigate the halogen− halogen (I···Cl and I···I) and hydrogen-bonding [Cl···O1(μ3hydroxo)] interactions in ThIOCl with the model fragment (Figure 8) at the M06-2X/LANL2DZ/6-311+G(d,p) level of theory.
Figure 9. RDG isosurfaces for the model fragment referring to the Cl···I and I···I bonding interactions (a). RDG isosurfaces for the model fragment referring to the Cl···I and O−H···Cl bonding interactions (b).
studied fragments (Figure S14), the noncovalent interactions of the Cl···I, I···I, and Cl···O1 bonding seem to be weak hydrogenbonding interactions in nature, which is consistent with QTAIM analysis. On the basis of binding energy analysis, the Cl···I average interaction energy in the model fragment is −8.3 kcal mol−1 with BSSE correction. Additionally, the sum of the Cl···I halogen−halogen energies is −57.9 kcal mol−1 among a delocalized halogen−halogen bond, while the Cl···O1 interaction energy is predicted to be −8.1 kcal mol−1. All of these results confirm the strong interaction to localize the interlayer Cl atoms. To the best of our knowledge, this is a new record value reported for a single halogen atom in a complex, further supporting the unexchangeable nature of Cl− anions in ThIOCl.
Figure 8. Structure for the model fragment of ThIOCl. Red, green, and purple spheres represent the O, Cl, and I atoms, respectively (all H atoms are omitted for clarity).
Natural population analysis shows that the natural charge on the Cl atom is −0.694, indicating nonnegligible charge transfer in the Cl···I and Cl···O1 interactions. As expected, based on the Raman spectral simulation (Figure S11), the calculated spectral bands are shifted to lower energy regions compared with the free IO3− ion20 with shift values of about 54 and 44 cm−1 (ν1), 67 cm−1 (ν2), 28 and 9 cm−1 (ν3), and 62 and 31 cm−1 (ν4). This indicates that the I−O bonds are weakened by the Cl···I halogen bonding. According to QTAIM11 analysis, the properties at bond critical points (BCPs) are important indicators for characterization of the interatomic bonding interactions. For the studied model fragment, the BCPs were clearly found for each of the Cl···I and Cl···O1 bonding interactions (Figure S12). Besides, the BCPs associated with I··· I bonding interactions were also found in the model fragment. As shown in Table S8, at the Cl···I and I···I BCPs, the electron densities (ρ) are between 0.0120 and 0.0233 au, and Laplacian of electron density (∇2ρ) values range from 0.0266 to 0.0506 au. All of these values are in the range of hydrogen bonds (0.002 < ρ < 0.035 au; 0.024 < ∇2ρ < 0.139 au).23 The ρ and ∇2ρ values for the Cl···O1 hydrogen-bonding interaction are predicted to be 0.0227 and 0.0812 au, respectively. It is worth noting that the Cl···I bonds with shorter interatomic lengths show higher ρ and ∇2ρ values and relatively stronger interactions. Additionally, all of the energy density (H) values are close to zero, also indicating closed-shell interaction of the Cl···I, I···I, and Cl···O1 bonding. To visualize the Cl···I, I···I, and Cl···O1 bonding interactions in ThIOCl, reduced density gradient (RDG)24 analysis was carried out at the same level of theory. Figure S13 shows the plot of reduced gradient versus sign(λ2) (ρ multiplied by the sign of the second Hessian eigenvalue). For the model fragment, the low-density and low-gradient spikes denote noncovalent interactions. From the RDG isosurfaces (Figure 9), it is clear that weak interactions exist in the Cl···I, I···I, and Cl···O1 bonding. According to the values of sign(λ2)ρ, different noncovalent interactions can be identified by color.24 For the
4. CONCLUSIONS In conclusion, a robust cationic network material featured with trigonal-pyramidal halogen bond geometries was obtained. The interlayer anions cannot be exchanged with other anions under ambient conditions or elevated temperature predominately stemming from the inherent affinity of strong halogen bonds. ThIOCl is a rare example of a pure inorganic cationic material constructed from halogen−halogen bonds. The present work demonstrated a new strategy in the combining of weak coordinated anions and weak interaction to synthesize inorganic cationic materials. Although strong halogen−halogen interaction is an obstacle to real anion-exchange applications, future work will be focused on how to deactivate these interactions in cationic materials.
■
ASSOCIATED CONTENT
S Supporting Information *
The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.6b01110. X-ray crystallographic data in CIF format (CIF) Detailed synthesis methods, PXRD, materials and characterization, X-ray crystallography, and Nyquist plots (PDF)
■
AUTHOR INFORMATION
Corresponding Authors
*E-mail:
[email protected]. *E-mail:
[email protected]. Author Contributions ⊥
E
These authors contributed equally. DOI: 10.1021/acs.inorgchem.6b01110 Inorg. Chem. XXXX, XXX, XXX−XXX
Article
Inorganic Chemistry Notes
(10) Boys, S. F.; Bernardi, F. Mol. Phys. 1970, 19, 553−566. (11) (a) Bader, R. F. W. Atoms in Moleculesa Quantum Theory; Oxford University Press: Oxford, U.K., 1990. (b) Bader, R. F. W. Chem. Rev. 1991, 91, 893−928. (c) Bader, R. F. W. J. Phys. Chem. A 1998, 102, 7314−7323. (d) Bader, R. F. W. Phys. Rev. B: Condens. Matter Mater. Phys. 1994, 49, 13348−13356. (12) Lu, T.; Chen, F. W. J. Comput. Chem. 2012, 33, 580−592. (13) (a) Shamov, G. A. Inorg. Chem. 2012, 51, 6507−6516. (b) Ma, P.; Wan, R.; Wang, Y.; Hu, F.; Zhang, D.; Niu, J.; Wang, J. Inorg. Chem. 2016, 55, 918−924. (14) (a) Knope, K. E.; Soderholm, L. Chem. Rev. 2013, 113, 944− 994. (b) Qian, X. Y.; Zhou, T. H.; Mao, J. G. Dalton Trans. 2015, 44, 13573−13580. (c) Knope, K. E.; Wilson, R. E.; Vasiliu, M.; Dixon, D. A.; Soderholm, L. Inorg. Chem. 2011, 50, 9696−9704. (15) Li, Y.; Yu, J. Chem. Rev. 2014, 114, 7268−7316. (16) (a) Conny, J. M.; Norris, G. A. Environ. Sci. Technol. 2011, 45, 7380−7386. (b) Ault, A. P.; Peters, T. M.; Sawvel, E. J.; Casuccio, G. S.; Willis, R. D.; Norris, G. A.; Grassian, V. H. Environ. Sci. Technol. 2012, 46, 4331−4339. (17) (a) Gilday, L. C.; Robinson, S. W.; Barendt, T. A.; Langton, M. J.; Mullaney, B. R.; Beer, P. D. Chem. Rev. 2015, 115, 7118−7195. (b) Mei, L.; Wu, Q. Y.; An, S. W.; Gao, Z. Q.; Chai, Z. F.; Shi, W. Q. Inorg. Chem. 2015, 54, 10934−10945. (c) Mei, L.; Wang, C.-z.; Wang, L.; Zhao, Y.-l.; Chai, Z.-f.; Shi, W.-q. Cryst. Growth Des. 2015, 15, 1395−1406. (d) Aiswaryalakshmi, P.; Mani, D.; Arunan, E. Inorg. Chem. 2013, 52, 9153−9161. (e) Ivanov, D. M.; Novikov, A. S.; Ananyev, I. V.; Kirina, Y. V.; Kukushkin, V. Y. Chem. Commun. 2016, 52, 5565−5568. (f) Pyziak, M.; Pyziak, J.; Hoffmann, M.; Kubicki, M. Cryst. Growth Des. 2015, 15, 5223−5232. (g) Pfrunder, M. C.; Micallef, A. S.; Rintoul, L.; Arnold, D. P.; McMurtrie, J. Cryst. Growth Des. 2016, 16, 681−695. (h) Mínguez Espallargas, G.; Zordan, F.; Arroyo Marín, L.; Adams, H.; Shankland, K.; van de Streek, J.; Brammer, L. Chem. Eur. J. 2009, 15, 7554−7568. (18) (a) Wang, H.; Wang, W.; Jin, W. J. Chem. Rev. 2016, 116, 5072− 5104. (b) Brown, A.; Beer, P. D. Chem. Commun. 2016, 52, 8645. (c) Sivchik, V. V.; Solomatina, A. I.; Chen, Y. T.; Karttunen, A. J.; Tunik, S. P.; Chou, P. T.; Koshevoy, I. O. Angew. Chem., Int. Ed. 2015, 54, 14057−14060. (d) Shi, H.; An, Z.; Li, P.-Z.; Yin, J.; Xing, G.; He, T.; Chen, H.; Wang, J.; Sun, H.; Huang, W.; Zhao, Y. Cryst. Growth Des. 2016, 16, 808−813. (e) Wiscons, R. A.; Zeller, M.; Rowsell, J. L. C. Cryst. Growth Des. 2016, 16, 2201−2210. (f) Cavallo, G.; Metrangolo, P.; Milani, R.; Pilati, T.; Priimägi, A.; Resnati, G.; Terraneo, G. Chem. Rev. 2016, 116, 2478−2601. (g) Khavasi, H. R.; Azhdari Tehrani, A. Inorg. Chem. 2013, 52, 2891−2905. (h) Bayse, C. A.; Rafferty, E. R. Inorg. Chem. 2010, 49, 5365−5367. (i) Derossi, S.; Brammer, L.; Hunter, C. A.; Ward, M. D. Inorg. Chem. 2009, 48, 1666−1677. (19) Pedireddi, V. R.; Reddy, D. S.; Goud, B. S.; Craig, D. C.; Rae, A. D.; Desiraju, G. R. J. Chem. Soc., Perkin Trans. 2 1994, 2, 2353−2360. (20) Bushiri, M. J.; Kochuthresia, T. C.; Vaidyan, V. K.; GautierLuneau, I. J. Nonlinear Opt. Phys. Mater. 2014, 23, 1450039. (21) Kochuthresia, T. C.; Gautier-Luneau, I.; Vaidyan, V. K.; Bushiri, M. J. J. Appl. Spectrosc. 2016, 82, 941−946. (22) (a) Geboes, Y.; Nagels, N.; Pinter, B.; De Proft, F.; Herrebout, W. A. J. Phys. Chem. A 2015, 119, 2502−16. (b) Nagels, N.; Hauchecorne, D.; Herrebout, W. A. Molecules 2013, 18, 6829−51. (c) Hauchecorne, D.; Szostak, R.; Herrebout, W. A.; van der Veken, B. J. ChemPhysChem 2009, 10, 2105−15. (23) Koch, U.; Popelier, P. J. Phys. Chem. 1995, 99, 9747−9754. (24) Johnson, E. R.; Keinan, S.; Mori-Sanchez, P.; Contreras-Garcia, J.; Cohen, A. J.; Yang, W. T. J. Am. Chem. Soc. 2010, 132, 6498−6506.
The authors declare no competing financial interest.
■
ACKNOWLEDGMENTS This work was supported by the National Natural Foundation of China (Grants 91326112 and 21422704), the Science Foundation of Jiangsu Province (Grant BK20140007), a Project Funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions (PAPD), and the “Young Thousand Talented Program” in China. Support for T.E.A.-S. was provided by the Chemical Sciences, Geosciences, and Biosciences Division, Office of Basic Energy Sciences, Office of Science, Heavy Elements Chemistry Program, U.S. Department of Energy, under Grant DE-FG0213ER16414.
■
REFERENCES
(1) Rives, V. Layered Double Hydroxides: Present and Future; Nova Science Publishers: New York, 2001. (2) (a) Wang, S.; Alekseev, E. V.; Diwu, J.; Casey, W. H.; Phillips, B. L.; Depmeier, W.; Albrecht-Schmitt, T. E. Angew. Chem., Int. Ed. 2010, 49, 1057−1060. (b) Lin, J.; Diwu, J.; Cross, J. N.; Villa, E. M.; Albrecht-Schmitt, T. E. Inorg. Chem. 2012, 51, 10083−10085. (3) (a) Wang, S.; Yu, P.; Purse, B. A.; Orta, M. J.; Diwu, J.; Casey, W. H.; Phillips, B. L.; Alekseev, E. V.; Depmeier, W.; Hobbs, D. T.; Albrecht-Schmitt, T. E. Adv. Funct. Mater. 2012, 22, 2241−2250. (b) Fei, H.; Pham, C. H.; Oliver, S. R. J. Am. Chem. Soc. 2012, 134, 10729−10732. (c) Zhao, X.; Mao, C.; Luong, K. T.; Lin, Q.; Zhai, Q. G.; Feng, P.; Bu, X. Angew. Chem., Int. Ed. 2016, 55, 2768−2772. (d) Desai, A. V.; Manna, B.; Karmakar, A.; Sahu, A.; Ghosh, S. K. Angew. Chem., Int. Ed. 2016, 55, 7811−7815. (e) Mao, C.; Kudla, R. A.; Zuo, F.; Zhao, X.; Mueller, L. J.; Bu, X.; Feng, P. J. Am. Chem. Soc. 2014, 136, 7579−7582. (f) Colinas, I. R.; Silva, R. C.; Oliver, S. R. Environ. Sci. Technol. 2016, 50, 1949−1954. (4) Chubar, N. J. Mater. Chem. A 2014, 2, 15995−16007. (5) Oliver, S. R. J. Chem. Soc. Rev. 2009, 38, 1868−1881. (6) (a) Goulding, H. V.; Hulse, S. E.; Clegg, W.; Harrington, R. W.; Playford, H. Y.; Walton, R. I.; Fogg, A. M. J. Am. Chem. Soc. 2010, 132, 13618−13620. (b) Poudret, T. J. P.; Prior, T. J.; McIntyre, L. J.; Fogg, A. M. Chem. Mater. 2008, 20, 7447−7453. (c) Fei, H.; Han, C. S.; Robins, J. C.; Oliver, S. R. J. Chem. Mater. 2013, 25, 647−652. (d) Manna, B.; Chaudhari, A. K.; Joarder, B.; Karmakar, A.; Ghosh, S. K. Angew. Chem., Int. Ed. 2013, 52, 998−1002. (e) Fu, H.-R.; Xu, Z.-X.; Zhang, J. Chem. Mater. 2015, 27, 205−210. (7) (a) Millet, P.; Bastide, B.; Pashchenko, V.; Gnatchenko, S.; Gapon, V.; Ksari, Y.; Stepanov, A. J. Mater. Chem. 2001, 11, 1152− 1157. (b) Ok, K. M.; Halasyamani, P. S. Inorg. Chem. 2002, 41, 3805− 3807. (8) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Montgomery, J. A., Jr.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Mennucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji, H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.; Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich, S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A. D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A. G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.; Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham, M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.; Johnson, B.; Chen, W.; Wong, M. W.; Gonzalez, C.; Pople, J. A. Gaussian 09; Gaussian, Inc.: Wallingford, CT, 2009. (9) Zhao, Y.; Truhlar, D. G. Acc. Chem. Res. 2008, 41, 157−167. F
DOI: 10.1021/acs.inorgchem.6b01110 Inorg. Chem. XXXX, XXX, XXX−XXX