1D and 3D Polymeric Manganese(II) Thiolato Complexes: Synthesis

Dec 19, 2017 - ... direct methods program SHELXS of the SHELXTL PC suite programs,(40) and ...... Dance , I. G. Formation, crystal structure, and reac...
0 downloads 0 Views 1MB Size
Article Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

pubs.acs.org/IC

1D and 3D Polymeric Manganese(II) Thiolato Complexes: Synthesis, Structure, and Properties of ∞3[Mn4(SPh)8 ] and ∞1[Mn(SMes)2 ] Andreas Eichhöfer*,†,‡,§ and Sergei Lebedkin† †

Institut für Nanotechnologie, Karlsruher Institut für Technologie (KIT), Campus Nord, Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany ‡ Lehn Institute of Functional Materials, Sun Yat-Sen University, Guangzhou 510275, China § Karlsruhe Nano Micro Facility (KNMF), Hermann-von-Helmholtz-Platz 1, 76344 Eggenstein-Leopoldshafen, Germany S Supporting Information *

ABSTRACT: Reactions of [Mn{N(SiMe3)2}2]2 with 2.1 equiv of RSH, R = Ph or Mes = C6H2−2,4,6-(CH3)3, yield compounds of the formal composition “Mn(SR)2”. Singlecrystal X-ray diffraction reveals that ∞1[Mn(SMes)2 ] forms one-dimensional chains in the crystal via μ2-SMes bridges, whereas ∞3[Mn4(SPh)8 ] comprises a three-dimensional network in which adamantanoid cages composed of four Mn atoms and six μ2-bridging SPh ligands are connected in three dimensions by doubly bridging SPh ligands. Thermogravimetric analysis and powder diffractometry indicate an reversible uptake of solvent molecules (tetrahydrofuran) into the channels of ∞1[Mn(SMes)2 ]. Magnetic measurements reveal antiferromagnetic coupling for both compounds with J = −8.2 cm−1 (∞1[Mn(SMes)2 ]) and −10.0 cm−1 (∞3[Mn4(SPh)8 ]), respectively. Their optical absorption and photoluminescence (PL) excitation spectra display characteristic d−d bands of Mn2+ ions in the visible spectral region. Both compounds emit bright phosphorescence at ∼800 nm at low temperatures (24° 2θ are very low in intensity, suggesting insufficient long-range order, especially of the mesityl groups, which are expected to contribute the most to the scattering intensity in this region of the diffractogram. Magnetic Properties. The static magnetic behavior of complex 1 was studied between 2 K and 300 K in a field of 100 Oe (Figure 3). A maximum in the magnetic susceptibility

Figure 3. Temperature dependence of (■) χ and (○) χT for 3 ∞[Mn4(SPh)8 ] (1). Solid green lines represent the results of the simulation using a Spin Hamiltonian (eq S1 in the SI) and the PHI program.

Figure 2. Molecular structure of ∞1[Mn(SMes)2 ] (2) viewed along (a) the crystallographic a-axis and (b) the c-axis (H atoms omitted for clarity). Selected bond lengths: Mn(1)−S(1), 246.12(3) pm. Selected bond angles: ∠S(1″)−Mn(1)−S(1‴), 89.36(1)°; ∠S(1)−Mn(1)− S(1″), 120.37(1)°; and ∠Mn(1)−S(1)−Mn(1″), 90.65(1)°. Symmetry transformations used to generate equivalent atoms: single prime symbol (′) denotes −x + 2, −y + 1, z; double prime symbol (′′) denotes −y + 3/2, x − 1/2, z − 1/2; triple prime symbol (‴) denotes y + 1 /2, −x + 3/2, −z − 1/2; and superscript “IV” denotes −y + 3/2, x − 1/2, z + 1/2.

versus temperature (χ vs T) at temperatures below 100 K is an indication for the presence of antiferromagnetic interactions in 1. An attempt was made to model this behavior with the PHI program28 by means of an isotropic spin Hamiltonian (SH), accounting for the exchange coupling (Heisenberg−Dirac-van Vleck Hamiltonian) (eq 1 in the SI). For 1, there exist theoretically six equal exchange pathways between the paramagnetic Mn2+ ions (d5, S = 5/2) through −SPh− bridges within the adamantanoid cages and four exchange pathways through −SPh− bridges between the cages in the 3-D crystal structure. The simulated curves for an isolated “Mn4(SPh)6(SPh)4/2” unit with six equal coupling constants J were found to describe the temperature-dependent susceptibility reasonably well, if antiferromagnetic coupling (J = −8.2 cm−1) between the Mn ions in the adamantanoid cages is assumed (Figure 3). The obvious mismatch of the simulated and experimental curve should result from the disregard of the coupling between the cluster cages. In this respect, we wish to note that we could not find a better model for the magnetic data of 1 within the program limitations of five spin centers. Isothermal magnetization 1 at 5 K shows no saturation up to 4.5 T, which is consistent with antiferromagnetic behavior (Figure S7 in the SI). It would be interesting to compare 1 with the related cluster anion [Mn4(SPh)10]2−.27 Unfortunately, magnetic data for the latter, so far, have not been reported. The χ vs T curve of 2 (Figure 4) measured between 2 K and 300 K in a field of 100 Oe displays a broad maximum centered at ∼125 K, which is characteristic for strong antiferromagnetic exchange in the polymeric chain. A similar behavior has been recently observed for the isostructural chain compounds

similar to isostructural ∞1[Co(SMes)2 ],21 in a different space group with a cell volume that is 200 Å3 larger, compared to the selenolato complex mentioned above. Therefore, the crystal structure of 2 consists of channels that are filled with solvent molecules (0.5 equiv, based on elemental analysis; i.e., the potential solvent volume amounts to 400 Å3). A comparison of the measured and calculated (based on the single-crystal structure data) powder X-ray diffraction (XRD) patterns for 1 and 2 indicates the crystalline purity, with respect to the formation of other crystalline compounds (Figures S2 and S3 in the SI). Slightly increasing differences in the positions of the peaks with increasing diffraction angle arise from the temperature difference between the data collections (singlecrystal XRD at 180 K and powder XRD at room temperature). TGA analysis of 2 reveals that the lattice solvent THF can be removed under vacuum in a well-separated “drying step” at temperatures between 100 and 160 °C, before decomposition (cleavage of SMes2) at ∼200 °C starts to give MnS (Figure S5 in the SI). Powder XRD shows that the first step with the accompanying change in the crystal structure can be partially reversed, if the dried material (heated to ∼155 °C) is exposed to THF vapor for several hours (Figure S6 in the SI). A reversal of the crystal structure of dried 2 is achieved by suspending the C

DOI: 10.1021/acs.inorgchem.7b02411 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. Temperature dependence of (■) χ and (○) χT for 1 ∞[Mn(SMes)2 ] (2). Solid green lines represent the results of the fit to the Fisher equation (eq S2 in the SI). 1 ∞[M(SeMes)2 ]

(M = Mn, Fe)19,20 and ∞1[Co(SMes)2 ].21 In close analogy to these earlier results, the classical Fisher model fit29,30 of the χ vs T curve of 2 for the data between 50 K and 300 K (eq S2 in the SI) yields J = −10.0(1) cm−1, suggesting a slightly weaker coupling in 2 than in the related mesitylselenolato complexes ∞1[M(SeMes)2 ] (M = Mn (J = −11.2 cm−1), Fe (J = −16.2 cm−1)). The deviation from the model at low temperatures, which has also been observed for the complexes ∞1[M(SeMes)2 ] (M = Mn, Fe) and ∞1[Co(SMes)2 ], could originate from an increasing influence of zero field splitting where states with low Jeff may be preferentially populated as the temperature is decreasing. A paramagnetic tail that is observed below ca. 25 K might be attributed to a small amount of paramagnetic impurities and/or contributions of paramagnetic ends of the chains. Similar to 1, the isothermal magnetization curve of 2 at 5 K shows no saturation up to 4.5 T (see Figure S8 in the SI). Optical Properties. Electronic absorption spectra of polycrystalline 1 and 2 can be roughly divided into a region of strong absorption below ca. 350 nm, because of charge transfer and ligand bands, and a region between ca. 350 and 600 nm with the less-intense bands due to spin-forbidden but symmetry-allowed d−d transitions in the tetrahedrally coordinated Mn2+ ions (Figure S9 in the SI). These regions are mirrored in the PL excitation (PLE) spectra of 1 and 2, with particularly well-structured d−d bands observed at low temperatures (see Figure 5, as well as Figures S10 and S11 in the SI). These are observed for 1 at 580, ∼535 (shoulder), 488, 460, 427, and 389 nm and may be assigned to 4T1(G) ← 6A1, 4 T2(G) ← 6A1, 4E(G) ← 6A1, 4A1(G) ← 6A1, 4T2(D) ← 6A1, 4 E2(D) ← 6A1 transitions, respectively.31 The d−d features for 2 are found at 553, 514, 478, 442, and 384 nm. Similar PLE spectra were also recorded for the related polymeric compounds ∞1[Mn(SePh)2 ] (3)19 and ∞1[Mn(SeMes)2 ] (4)18 (see Figure 5, as well as Figures S12 and S13 in the SI). In particular, the thiolato compounds 1 and 2 and selenolato compounds 3 and 4 demonstrate rather similar PLE patterns of d−d transitions (Figure 5) − reflecting close S−Mn−S and Se−Mn−Se environments of the tetrahedrally coordinated Mn(II) ions, respectively. Note that, because of the nonvalidity of the Lambert−Beer law for this type of sample preparation (see the Experimental Section), strong absorption bands have a tendency to be saturated and a comparison of absorption (PLE) intensities must be considered carefully.

Figure 5. Photoluminescence emission (PL) and excitation (PLE) spectra of polycrystalline ∞3[Mn4(SPh)8 ] (1), ∞1[Mn(SMes)2 ] (2), 1 1 ∞[Mn(SeMes)2 ] (3), and ∞[Mn(SePh)2 ] (4) at a temperature of 17 K. PL and PLE spectra were excited at 330 nm for all compounds and recorded at 780, 780, 920, and 740 nm for 1, 2, 3, and 4, respectively.

All polymeric complexes 1−4 show bright red/NIR phosphorescence at temperatures below ∼100 K (∼20 K for 3). It peaks at 809, 793, 863, and 742 nm for 1−4, respectively (see Figure 5, as well as Figures S10−S13 in the SI) and decays on the time scale of a few hundred microseconds at T = 17 K (see Figure S14 in the SI). Interestingly, the PL intensity of the phenyl-substituted compounds 1 and 3 decreases drastically by increasing the temperature up to 293 K (by a factor of ∼6500 for 1), whereas that of the mesityl compounds 2 and 4 decreases only moderately. The PL quantum efficiency of 2 at ambient temperature was determined to be 1.2%, by using an integrating sphere. According to the temperature-dependent PL spectra, it can be estimated as 9.5% below 100 K (see Figure S11 in the SI). The red/NIR PL of 1−4, with a relatively large Stokes shift, is in contrast to a characteristic green-to-orange emission that has been observed for various Mn(II)-doped inorganic materials, e.g., ZnS:Mn and CdSe:Mn nanoparticles,32,33 and attributed to a d−d transition (6A1 ← 4T1(G)) of Mn ions. The emission of 1−4 appears to be of a different origin and may be tentatively assigned to “delocalized” triplet excitations contributed by the Mn(II) metal ion(s), S (Se) bridges, and phenyl (mesityl) ligands. A similarly red-shifted PL has also been observed for manganese-benzimidazole complexes and attributed to metal-to-ligand charge-transfer states.34 This distinct characteristic of the emission in 1−4 is also supported by highpressure PL measurements of 2 (a relatively strong emitter under ambient conditions) in a diamond anvil cell, as described in the SI. By applying a hydrostatic pressure up to ca. 10 kbar, the PLE bands of 2 within 350−600 nm due to d−d transitions of Mn(II) show a moderate red shift of ∼0.6 nm/kbar (Figure S15 in the SI). Such behavior is typical for the d−d absorption bands of Mn(II).35 The d−d emission of Mn(II), which is also D

DOI: 10.1021/acs.inorgchem.7b02411 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

started to sublime into a connected Schlenk tube. Sublimation continued until the oil bath reached 95 °C and the residue turned almost white in color, to give a final yield of 1.26 g (86.3%) of [Mn{N(SiMe3)2}2]2. C24H72Mn2N4Si8 (751.42): calcd. C 38,4, H 9,7, N 7,5; found C 38.1, H 9.4, N 7.9%. A powder XRD pattern is given in Figure S16 in the SI. 3 ∞[Mn4(SPh)8 ] (1). [Mn{N(SiMe3)2}2]2 (0.7 mL, 631 mg, 0.84 mmol) was reacted with HSPh (0.36 mL, 3.53 mmol) in 30 mL of THF at 0 °C to yield a beige solution. After 2 h, the reaction solution was layered with Et2O through slow diffusion by evaporation from a connected flask to yield, within three days, red-brown crystals of 1. These were filtered after two additional days and washed two times with 10 mL of THF/Et2O (1:1) to give 355 mg (77.3%). C48H40Mn4S8 (1093.1): calcd. C 52.7%, H 3.7%, S 23.5%; found C 52.5%, H 3.5%, S 23.6%. 1 ∞[Mn(SMes)2 ] · 0.5C4 H8O (2). [Mn{N(SiMe3)2}2]2 (0.5 mL, 451 mg, 0.6 mmol) was reacted with HSMes (0.38 mL, 2.52 mmol) in 25 mL of THF at 0 °C to yield a beige solution. Upon standing, pink crystals of 2 soon started to grow, which were filtered after one additional day, washed two times with 10 mL of cold (0 °C) THF to give a total yield of 443 mg (94%). C18H22MnS2·0.5C4H8O (393.5): calcd. C 61.1, H 6.7, S 16.3; found C 60.8, H 6.7, S 15.9%. 1 1 ∞[Mn(SePh)2 ] and ∞[Mn(SeMes)2 ] have been prepared according to ref 19. Crystallography. Crystals suitable for single-crystal XRD were taken directly from the reaction solution and then selected in perfluoroalkylether oil. Single-crystal XRD data of 1 and 2 were collected using graphite-monochromatised Mo Kα radiation (λ = 0.71073 Å) on a STOE IPDS II (Imaging Plate Diffraction System). Raw intensity data were collected and treated with the STOE X-Area software, Version 1.39. Data for all compounds were corrected for Lorentz and polarization effects. Based on a crystal description, numerical absorption corrections were applied for 1.39 The structures were solved with the direct methods program SHELXS of the SHELXTL PC suite programs,40 and were refined with the use of the full-matrix least-squares program SHELXL. Molecular diagrams were prepared using Diamond.41 All Mn, S, and C atoms were refined in 1 and 2 with anisotropic displacement parameters, while H atoms were computed and refined, using a riding model, with an isotropic temperature factor equal to 1.2 times the equivalent temperature factor of the atom to which they are linked. Crystals of 1 display a low scattering intensity at higher 2θ and were therefore only measured up to 49° 2θ. In 1, phenyl rings of the −SPh ligands are partially disordered and C atoms were therefore refined with a split model of site disorder. Lattice solvent molecules were identified within the structure of 2. These are located at a special position but could not be adequately refined, because of disorder. Therefore, the data were corrected for these, using the SQUEEZE option within the PLATON program package,42 finding a total of 41 electrons in a potentially solvent accessible volume of ∼400 Å3. CCDC-1553253 (1) and 1553252 (2) contain supplementary crystallographic data for this paper. These data can be obtained free of charge at www.ccdc.cam.ac.uk/conts/retrieving.html (or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (international) +44 1223/336 033; Email:[email protected]). Powder XRD patterns were measured on a STOE STADI P diffractometer (Cu Kα1 radiation, germanium monochromator, Debye−Scherrer geometry, PSD) in sealed glass capillaries. Theoretical powder diffraction patterns were calculated on the basis of the atom coordinates obtained from single-crystal X-ray analysis by using the program package STOE WinXPOW.43 Physical Measurements. C, H, S elemental analysis was performed on an Elementar vario Micro cube instrument. TGA experiments were run in Al2O3 crucibles on a thermobalance STA 409 from Netzsch, under a vacuum of 1.5 × 10−6 mbar at a heating rate of 2 °C/min. The crucibles were filled with microcrystals of 1 and 2 (ca. 20−35 mg) inside an argon glovebox (which is due to a high degree of

red-shifted under pressure, is persistent up to at least 100 kbar.35 In contrast, the emission spectrum of 2 changes dramatically above ca. 3 kbar, with a PL maximum shifted from 750 nm to 800 nm at ca. 10 kbar. Furthermore, by increasing pressure up to 24 kbar, the emission disappears. This suggests a considerable structural transformation of 2 under (relatively moderate) pressure, which may be facilitated by the presence of channels in its crystal structure (see above). On the other hand, this transformation is reversible: the PL of 2 is practically completely recovered after the pressure has been released (see Figure S15).



CONCLUSION Two polymeric Mn(II) thiolato complexesthree-dimensional ( 3 D ) ∞3[Mn4(SPh)8 ] a n d o n e - d i m e n s i o n a l ( 1 D ) 1 ∞[Mn(SMes)2 ]have been successfully synthesized and structurally characterized via single-crystal XRD. 3 ∞[Mn4(SPh)8 ] is the first example of a 3D network chalcogenolato compound for other than group 12 transition metals. The fact that ∞1[Mn(SePh)2 ] instead forms a 1D chain structure indicates only slight energetic differences between both types of structures in the case of manganese as a transition-metal ion. Thus, these structures might probably be controlled by variation of the reaction/crystallization conditions. The structure of ∞1[Mn(SMes)2 ] is porous and consists of channels that can be reversibly filled with solvent (tetrahydrofuran, THF) molecules. Both complexes display strong antiferromagnetic coupling of the Mn2+ ions. The polymeric Mn(II) thiolato complexes, as well as the related selenolato complexes ∞1[Mn(SePh)2 ] and ∞1[Mn(SeMes)2 ] reported earlier were found to exhibit similar bright phosphorescence at ∼800 nm at temperatures below ∼100 K. The most intense emitter at ambient temperature is 1 ∞[Mn(SMes)2 ] with the PL quantum yield of 1.2%. PL measurements of this compound in a diamond anvil cell revealed very different behavior of the Mn(II) d−d absorption bands (in the PLE spectra) and the emission band under high pressure. These results support an assignment of the broad red/ NIR emission to triplet excitations contributed by the both Mn(II) metal ion(s), S (Se) bridges, and phenyl (mesityl) ligands. However, the exact character of the emissive excitation(s) in polymeric Mn(II) chalcogenolato compounds and the degree of their possible delocalization over the networks (i.e., beyond a single “monomeric unit”) are not clear at the moment and open to future investigations.



EXPERIMENTAL SECTION

Synthesis. Standard Schlenk techniques were employed throughout the syntheses using a double manifold vacuum line with highpurity dry nitrogen (99.9994%) and a MBraun Glovebox with highpurity dry argon (99.999%). The solvents THF (tetrahydofuran) and diethyl ether were dried over sodium benzophenone, and distilled under nitrogen. HSPh (distilled prior to use), LiN(SiMe3)2 (sublimed prior to use), and anhydrous MnCl2 were purchased from Aldrich. HSMes was synthesized according to the literature.36 [Mn{N(SiMe3)2}2]2 was prepared by a modified procedure according to refs37 and 38; anhydrous MnCl2 (0.5 g, 3.97 mmol) and Li(N(SiMe3)2 (1.33 g, 7.95 mmol) were suspended in 10 mL of toluene. Upon the addition of 5 mL of Et2O, the reaction mixture gently grow warm and was further heated for 12 h in an oil bath at 70 °C. The solvents were then removed by reduced pressure and the residue heated in vacuum (10−3 Torr). At 60 °C, a pale pink powder E

DOI: 10.1021/acs.inorgchem.7b02411 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry moisture and oxygen sensitivity of the compounds), transferred in Schlenk tubes and mounted under a stream of argon to the balance. Caution should be taken, with respect to the bad odor from the decomposition products. Zero-field-cooled (ZFC) temperature-dependent magnetic susceptibilities were recorded for 1 and 2 in RSO mode, using a MPMS-5S (Quantum Design) SQUID magnetometer over a temperature range from 2 K to 300 K in a homogeneous external magnetic field of 0.01 T. Magnetization curves were measured on the same instrument in a dc field up to 4.5 T. The samples were contained in gelatin capsules filled in a glovebox. They were transferred in sealed Schlenk tubes from the glovebox to the magnetometer and then rapidly transferred to the helium-purged sample space of the magnetometer. The data were corrected for the sample holder, including the gelatin capsule, and for diamagnetism using Pascal’s constants.30,44,45 UV-vis absorption spectra of polycrystalline samples dispersed (in a glovebox) in a mineral oil (Sigma−Aldrich) layer between two quartz plates were measured on a PerkinElmer Lambda 900 spectrophotometer, using an integrating sphere (Labsphere). Photoluminescence measurements of the same sample preparations were performed on a Horiba JobinYvon Fluorolog-3 spectrometer that was equipped with a Hamamatsu R5509 UV-vis-NIR photomultiplier (∼300−1400 nm) and an optical close-cycle cryostat (Leybold) for cooling samples down to 16 K. The emission spectra were corrected for the wavelength-dependent response of the spectrometer and detector (in relative photon flux units). PL decay traces were recorded by connecting the photomultiplier to an oscilloscope (typically with a 500 Ω load) and using a N2 laser for pulsed excitation at 330 nm (∼2 ns, ∼5 μJ per pulse). PL quantum yield of 2 at ambient temperature was determined using an integrating sphere, according to ref 46, with the uncertainty estimated as ±10%. PL measurements of 2 under high pressure in a diamond anvil cell are described in the SI.



K. Powell for a generous support, E. Tröster and N. Metz for assistance in the synthesis, and S. Stahl for performing the elemental analysis.



(1) He, Y.; Zhou, W.; Qian, G.; Chen, B. Methane storage in metal− organic frameworks. Chem. Soc. Rev. 2014, 43, 5657−5678. (2) van de Voorde, B.; Bueken, B.; Denayer, J.; De Vos, D. Adsorptive separation on metal−organic frameworks in the liquid phase. Chem. Soc. Rev. 2014, 43, 5766−5688. (3) Hu, Z.; Deibert, B. J.; Li, J. Luminescent metal−organic frameworks for chemical sensing and explosive detection. Chem. Soc. Rev. 2014, 43, 5815−5840. (4) Zhang, T.; Lin, W. Metal−organic frameworks for artificial photosynthesis and photocatalysis. Chem. Soc. Rev. 2014, 43, 5982− 5993. (5) Zhou, H. C.; Kitagawa, S. Metal−Organic Frameworks (MOFs). Chem. Soc. Rev. 2014, 43, 5415−5418. (6) Cave, D.; Gascon, J.-M.; Bond, A. D.; Teat, S. J.; Wood, P. T. Layered metal organosulfides: hydrothermal synthesis, structure and magnetic behaviour of the spin-canted magnet Co(1,2-(O2C)(S)C6H4. Chem. Commun. 2002, 1050−1051. (7) Lobana, T. S.; Sharma, R.; Bermejo, E.; Castineiras, A. Construction of a novel copper(I) chain polymer of hexanuclear Cu6(2-SC5H4NH)6I6 cores with a new (μ3-S) mode of bonding of pyridine-2-thione and of an unusual triangular Cu3I3(dppe)3(2SC5H4NH) cluster. Inorg. Chem. 2003, 42, 7728−7730. (8) Zhao, Y.; Hong, M.; Liang, Y.; Cao, R.; Li, W.; Weng, J.; Lu, S. A paramagnetic lamellar polymer with a high semiconductivity. Chem. Commun. 2001, 1020−1021. (9) He, J.; Yee, K. K.; Xu, Z.; Zeller, M.; Hunter, A. D.; Chui, S. S. Y.; Che, C. M. Thioether side chains improve the stability, fluorescence, and metal uptake of a metal organic framework. Chem. Mater. 2011, 23, 2940−2947. (10) Yee, K. K.; Reimer, N.; Liu, J.; Cheng, S. Y.; Yiu, S. M.; Weber, V.; Stock, N.; Xu, Z. Effective mercury sorption by thiol-laced metal− organic frameworks: in strong acid and the vapor phase. J. Am. Chem. Soc. 2013, 135, 7795−7798. (11) Huang, R.-W.; Wei, Y. S.; Dong, X.-Y.; Wu, X.-H.; Du, C.-X.; Zang, S.-Q.; Mak, T. C. W. Hypersensitive dual-function luminescence switching of a silver-chalcogenolate cluster-based metal−organic framework. Nat. Chem. 2017, 9, 689−697. (12) Yan, H.; Hohman, J. N.; Li, F. H.; Jia, C.; Solis-Ibarra, D.; Wu, B.; Dahl, J. E. P.; Carlson, R. M. K.; Tkachenko, B. A.; Fokin, A. A.; Schreiner, P. R.; Vailionis, A.; Kim, T. R.; Devereaux, T. P.; Shen, Z.X.; Melosh, N. A. Hybrid metal−organic chalcogenide nanowires with electrically conductive inorganic core through diamondoid-directed assembly. Nat. Mater. 2016, 16, 349−355. (13) Ananikov, V. P.; Beletskaya, I. P. Preparation of metal “nanosalts” and their application in catalysis: heterogeneous and homogeneous pathways. Dalton Trans. 2011, 40, 4011−4023. (14) Dance, I. G.; Garbutt, R. G.; Craig, D. C.; Scudder, M. L. The different nonmolecular polyadamantanoid crystal structures of cadmium benzenethiolate and 4-methylbenzenethiolate. Analogies with microporous aluminosilicate frameworks. Inorg. Chem. 1987, 26, 4057−4064. (15) Dance, I. G. Formation, crystal structure, and reactions of catena-(μ-SPh)[(μ-SPh)6Zn4(CH3OH)(SPh)], a model for zinc thiolate metalloenzymes. J. Am. Chem. Soc. 1980, 102, 3445−3451. (16) Anjali, K. S.; Vittal, J. J. Polylinked adamantanoid structure of Cd(SePh)2. Inorg. Chem. Commun. 2000, 3, 708−710. (17) Cuthbert, H. L.; Wallbank, A. I.; Taylor, N. J.; Corrigan, J. F. Synthesis and structural characterization of [Cu 20 Se 4 (μ 3 SePh)12(PPh3)6] and [Ag(SePh)]. Z. Anorg. Allg. Chem. 2002, 628, 2483−2488. (18) Bochmann, M.; Powell, A. K.; Song, X. Synthesis and characterisation of manganese(II) chalcogenolato complexes. Crystal

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b02411. Crystal structures, measured and simulated X-ray powder patterns, TGA data, magnetic measurements, solid-state UV-vis-NIR absorption and PL spectra (PDF) Accession Codes

CCDC 1553252−1553253 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



REFERENCES

AUTHOR INFORMATION

Corresponding Author

*Tel.: 49-(0)721-608-26371. Fax: 49-(0)721-608-26368. Email: [email protected]. ORCID

Andreas Eichhöfer: 0000-0002-3412-6280 Author Contributions

A.E. performed synthesis, XRD, and magnetic measurements, S.L. performed PL measurements. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the Karlsruhe Institut für Technologie (KIT, Campus Nord). A.E. wishes to thank A. F

DOI: 10.1021/acs.inorgchem.7b02411 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry and molecular structure of [{Mn(μ-SeC6H2Me3-2,4,6)2}∞]. J. Chem. Soc., Dalton Trans. 1995, 1645−1648. (19) Eichhöfer, A.; Wood, P. T.; Viswanath, R. N.; Mole, R. A. Synthesis, structure and magnetic behaviour of manganese(II) selenolate complexes [Mn(SePh)2], [Mn(SePh)2(bipy)2] and [Mn(SePh)2(phen)2] (bipy = bipyridyl, phen = phenanthroline). Eur. J. Inorg. Chem. 2007, 2007, 4794−4799. (20) Eichhöfer, A.; Buth, G.; Dolci, F.; Fink, K.; Mole, R. A.; Wood, P. T. Homoleptic 1-D iron selenolate complexesSynthesis, structure, magnetic and thermal behaviour of [Fe(SeR)2] (R = Ph, Mes). Dalton Trans. 2011, 40, 7022−7032. (21) Eichhö fer, A.; Buth, G. Polymeric cobalt(II) thiolato complexesSyntheses, structures and properties of [Co(SMes)2] and [Co(SPh)2NH3]. Dalton Trans. 2016, 45, 17382−17391. (22) Bochmann, M.; Coleman, A. P.; Webb, K. J.; Hursthouse, M. B.; Mazid, M. Synthesis of sterically hindered tellurophenols and the structure of [Cd(μ-TeC6H2M3)2]. Angew. Chem., Int. Ed. Engl. 1991, 30, 973−975. (23) Fenske, D.; Fischer, A. New Selenolato-Bridged Clusters of Iron and Nickel; the Structures of [Fe12(SePh)24] and [Na2(POPh3)6][Ni20Se12(SeMe)10]. Angew. Chem., Int. Ed. Engl. 1995, 34, 307−309. (24) Wang, J.; Jian, F.; Huang, B.; Bai, Z. Two molecular wheels 12MC-6 complexes: synthesis, structure and magnetic properties of [Co(μ2-SEt)2]6 and [Fe(μ2-SEt)2]6. J. Solid State Chem. 2013, 204, 272−277. (25) Ivanov, A. A.; Kozee, M. A.; Merrill, W. A.; Agarwal, S.; Dahl, L. Cyclo-[Ni(μ2-SPh)2]9 and cyclo-[Ni(μ2-SPh)2]11: new oligomeric types of toroidal nickel(II) thiolates containing geometrically unprecedented 9- and 11-membered ring systems. J. Chem. Soc., Dalton Trans. 2002, 4105−4115. (26) Zhang, C.; Takada, S.; Kölzer, M.; Matsumoto, T.; Tatsumi, K. Nickel(II) thiolate complexes with a flexible cyclo-{Ni10S20} framework. Angew. Chem., Int. Ed. 2006, 45, 3768−3772. (27) Costa, T.; Dorfman, J.; Hagen, K. S.; Holm, R. H. Synthesis and stereochemistry of mono-, bi-, and tetranuclear manganese thiolates. Inorg. Chem. 1983, 22, 4091−4099. (28) Chilton, N. F.; Anderson, R. P.; Turner, L. D.; Soncini, A.; Murray, K. S. PHI: A powerful new program for the analysis of anisotropic monomeric and exchange-coupled polynuclear d- and fblock complexes. J. Comput. Chem. 2013, 34, 1164−1175. (29) Fisher, M. E. Magnetism in one-dimensional systemsThe Heisenberg model for infinite spin. Am. J. Phys. 1964, 32, 343−346. (30) Kahn, O. Molecular Magnetism; Wiley−VCH: Weinheim, Germany, 1993. (31) Lever, A. B. P. Inorganic Electronic Spectroscopy, 2nd Edition; Elsevier: Amsterdam, 1984; p 448. (32) Cao, L.; Zhang, J.; Ren, S.; Huang, S. Luminescence enhancement of core-shell ZnS:Mn/ZnS nanoparticles. Appl. Phys. Lett. 2002, 80, 4300−4302. (33) Beaulac, R.; Archer, P. I.; Ochsenbein, S. T.; Gamelin, D. R. Mn2+-doped CdSe quantum dots: new inorganic materials for spinelectronics and spin-photonics. Adv. Funct. Mater. 2008, 18, 3873− 3891. (34) Wei, Y.; Yu, Y.; Wu, K. Highly stable five-coordinated Mn(II) polymer [Mn(Hbidc)]n (Hbidc = 1H-Benzimidazole-5,6-dicarboxylate): Crystal structure, antiferromegnetic property, and strong longlived luminescence. Cryst. Growth Des. 2008, 8, 2087−2089. (35) Nataf, L.; Barreda-Argüeso, J. A.; Valiente, R.; González, J.; Rodríguez, F. Effects of Cu2+ doping and pressure on the exchangemediated exciton dynamics in one-dimensional N(CH3)4MnCl3. Phys. Rev. B 2014, 89, 115120. (36) Chisholm, M.; Corning, J.; Folting, K.; Huffman, J. Triple bonds between molybdenum and tungsten atoms supported by thiolate ligands: (RS)3MM(SR)3. Polyhedron 1985, 4, 383−390. (37) Bürger, H.; Wannagat, U. Silylamido-Verbindungen von Chrom, Mangan, Nickel und Kupfer. Monatsh. Chem. 1964, 95, 1099−1102. (38) Bradley, D. C.; Hursthouse, M. B.; Abdul Malik, K. M.; Möseler, R. The crystal molecular structure of “bis(hexamethyldisilylamido) manganese”. Transition Met. Chem. 1978, 3, 253−254.

(39) X-RED32 1.01, Data Reduction Program; Stoe & Cie GmbH: Darmstadt, Germany, 2001. (40) Sheldrick, G. M. SHELXTL, PC version 5.1; An Integrated System for Solving, Refining, and Displaying Crystal Structures from Diffraction Data; Bruker Analytical X-ray Systems: Karlsruhe, Germany, 2000. (41) Brandenburg, K. Diamond Version 2.1d; Crystal Impact GbR, 1996−2000. (42) Spek, A. L. PLATON. A Multipurpose Crystallographic Tool; Utrecht University: The Netherlands, 2002. (http://www.cryst.chem. uu.nl/platon). (43) STOE WinXPOW; STOE & Cie GmbH: Darmstadt, Germany, 2000. (44) Lueken, H. Magnetochemie; B. G. Teubner Stuttgart: Leipzig, Germany, 1999; p 426. (45) Haberditzl, W. Advances in molecular diamagnetism. Angew. Chem., Int. Ed. Engl. 1966, 5, 288−323. (46) de Mello, J. C.; Wittmann, H. F.; Friend, R. H. An improved experimental determination of external photoluminescence quantum efficiency. Adv. Mater. 1997, 9, 230−232.

G

DOI: 10.1021/acs.inorgchem.7b02411 Inorg. Chem. XXXX, XXX, XXX−XXX