An MOF-like Interpenetrated 2D Plus 2D to 3D ... - ACS Publications

Apr 27, 2018 - •S Supporting Information. ABSTRACT: Remarkable progresses regarding pure inorganic frameworks and metal−organic frameworks (MOF) ...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. XXXX, XXX, XXX−XXX

An MOF-like Interpenetrated 2D Plus 2D to 3D Inorganic Grid Assembled by Linear Inorganic Pillars, Structures, and Properties in Supercapacitance Nan-Nan Xu, Li-Wen Qian, Zhao-Qi Li, Guo-Qing Bian, Qin-Yu Zhu,* and Jie Dai* College of Chemistry, Chemical Engineering and Materials Science, Soochow University, Suzhou 215123, People’s Republic of China

Downloaded via UNIV OF MASSACHUSETTS AMHERST on July 10, 2018 at 13:38:21 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: Remarkable progresses regarding pure inorganic frameworks and metal−organic frameworks (MOF) have been made. However, pure inorganic frameworks with MOFlike grid structures are rarely reported due to the weakness of inorganic moiety as a long linear linker. We report herein a fascinating inorganic framework assembled by a [Ge4S10]4− cluster node and a linear [−Cu−MS4−Cu−] (M = Mo (1) and W (2)) inorganic pillar. Their network shows MOF-like orthogonal structure with two interpenetrated two-dimensional (2D) plus 2D to 3D framework and a 1D nano tunnel. Electrodes with crystalline sample of 1 and 2, inorganic sulfide framework, were prepared, and their quasi-capacitance behaviors were investigated. Electrochemical performances were evaluated by cyclic voltammetry and galvanostatic charge−discharge techniques in CsOH, KOH, NaOH, and LiOH electrolytes. The results revealed that the crystal materials exhibit moderate specific capacitance values that are comparable to those of porous sulfide materials.



INTRODUCTION During the past decades, inorganic−organic hybrid metal− organic frameworks (MOFs), typically made of a metal ion or a metal coordination cluster and an organic linker (rigid or flexible), have attracted immense attention due to their fascinating topologies and potential applications in gas storage, separation, ion exchange, and catalysis.1,2 A huge number of coordination networks are tuned and accessed attributed to the designable nature of organic ligands and the rich coordination geometries of metal ions. Recently, remarkable progresses regarding MOF materials have been made for additional applications in electronic devices, energy storage, and conversion.3−6 The recent development in the field of MOF-derived porous carbon, metal oxide, or chalcogenide nanostructures for supercapacitor applications has been greatly interesting.7−14 It appears that there is limited work on pristine MOF materials especially for high-performance supercapacitor. This is mainly due to the drawback of MOF materials: the low conductivity. Actually, using inorganic building blocks to construct mesostructures emerged prior to the creation of MOFs. Inorganic porous structures using relatively large building blocks were synthesized as early as decades ago.15,16 Recently, excellent progress has also been made in the inorganic mesostructure materials; for example, a large family of polyoxometalate (POM)-based porous materials17−22 and a family of metal chalcogenides-based open frameworks23−27 were developed. Unlike the MOFs, these inorganic porous structures, such as © XXXX American Chemical Society

POMs, have no distinct long connecting-pillar and grid structure characters and, consequently, are mainly modulated in void sizes from the nano- to micrometer scales by carefully selecting templating materials (Chart 1). Therefore, in contrast Chart 1. Schematic View of the Difference between POM and MOF

to the large number of inorganic−organic hybrid MOFs, inorganic frameworks featured with pure inorganic nodes and long connecting pillars are relatively uncommon. Received: April 27, 2018

A

DOI: 10.1021/acs.inorgchem.8b01168 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 1. (a) XRD patterns of the microcrystal samples of 1 and 2 with those simulated from the data of single-crystal analysis. (b) Solid-state absorption spectra of 1 and 2 in representation of electronvolts. (c) Excitation and emission spectra of 1 in solid state. (d) IR spectra of 1 and 2.

Both of the compounds showed desirable features for supercapacitor, and the relationship of the specific capacitances with the ion exchanging properties of the electrode materials was evaluated.

Interpenetrating network structures have rapidly attracted current attention, not only for their intriguing variety of topologies but also for their functions in modulating the space size. Currently, most of the reported interpenetrating structures were based on MOF materials and have been extensively discussed in comprehensive reviews.28−30 Although there are a few interpenetrating pure inorganic polymers,31−34 they are still very limited, not only in numerical number but also in structural topologies. Moreover, very few inorganic interpenetrating frameworks are assembled by linear connectors. This is not surprising, because the prerequisite for constructing interpenetrating frameworks is that the connecting rods must have sufficient length, which usually could not be met for inorganic connectors. An example of inorganic linear linker is the binary edge-shared bi-tetrahedra P2S6.35 If the metal cyanide polymers were considered as the inorganic moieties, some interpenetrating frameworks were reported with linear [−NC−M″−CN−] linkers (M″ = Ag(I), Cu(I), and Au(I)).36,37 Nevertheless, the cyanide ion has different nature from inorganic metalate components of oxide and chalcogenides. In this work we report a fascinating inorganic interpenetrating framework assembled by [Ge4S10]4− clusters and linear [−Cu−MS4−Cu−] (M = Mo (1) and W (2)) inorganic pillars. The network shows an MOF-like orthogonal structure with twofold interpenetrated two-dimensional (2D) inorganic grid and 1D nano tunnel character. As mentioned above, MOFs as the precursors of porous materials can be integrated into supercapacitor devices. However, electron-conductive MOFs as materials in supercapacitor research seem to be just the beginning of the field of MOFs for energy-storage applications.38,39 The frameworks of 1 and 2 are semiconductive sulfides. In this work, electrodes with semiconductive crystalline samples of 1 and 2 were fabricated to uncover their capacitance behaviors.



RESULTS AND DISCUSSION To prepare compounds 1 and 2, analytically pure [(CH3)4N]4Ge4S10 and (NH4)2MS4 (M = Mo and W) at 1:1 mol ratio reacted with excessive Cu(OAc)2 in dimethylformamide (DMF). A small quantity of water and ethanol was added to modulate the solubility of the starting materials. The reaction mixture was sealed and heated at 100 °C for 3 d yielding red and yellow crystals, [(CH3)4N]4[(Ge4S10)Cu4Mo2S8] (1) and [(CH3)4N]4[(Ge4S10)-Cu4W2S8] (2). The purity of the bulky microcrystal samples of 1 and 2 was confirmed by elemental analysis and by comparing the experimental X-ray diffraction (XRD) patterns with the calculated patterns from the crystal data (Figure 1a). Solid-state UV−vis−NIR (NIR = niearinfrared) absorption spectra of crystals 1 and 2 were calculated from their diffuse-reflection spectra (Figure 1b). The onset energies of the absorption bands are 1.9 eV for 1 and 2.4 eV for 2; therefore, 1 and 2 belong to the narrow-band materials. Excitation and emission spectra of 1 gave curves in a wellmirror symmetry (Figure 1c), which shows a typical exciting and emitting energy-transfer mechanism. The IR peaks at 3016 and 1478 cm−1 in IR spectra of the compounds are contributed by stretches and bending vibrations of [(CH3)4N]+ ion, respectively (Figure 1d). The bands at 795 cm−1 are attributed to the vibration of the Ge4S10 cluster. The results of thermogravimetry are in agreement with the formulas of the compounds, which are shown in Figure S1, in Supporting Information. The compounds are stable before 240 °C and decompose approximately in two major steps in the range from 240 to 850 °C. B

DOI: 10.1021/acs.inorgchem.8b01168 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry An XRD study performed on a single crystal reveals that compound 1 crystallizes in a tetragonal space group I41/acd (Table S1). The anion structure of 1 features a 2D plus 2D to 3D interpenetrating framework constructed from [Ge4S10]4−cluster nodes and inorganic linear [−Cu−MoS4−Cu−] pillars (Figure 2). The MoS42−, with twofold disorder, coordinates linearly to two

Cu(I) ions that are in planar triangle coordination conformation to be a rigid pillar. This pillar moiety has been reported to assemble a square molecule.40 The super-tetrahedral T223 [Ge4S10]4− cluster takes a D2d symmetry and coshares four corner sulfur atoms with the linear pillars to form a 2D (4,4) grid topology with node-to-node distances of 15.03 and 14.47 Å. One of the 2D frameworks is parallel to ac plane, and the other is parallel to bc plane. The [S−Cu−MoS4−Cu−S] distance is ∼1 nm (9.590 Å) along the a or b direction. Along the c direction the [S−Cu−MoS4−Cu−S] distance is 9.577 Å. The two interpenetrated 2D nets construct a MOF-like 3D structure (Figure 2b) that is uncommon for inorganic materials, since most of the inorganic porous solids lack the grid character as the MOFs with long pillars. Figure 2c shows the space-filling model of 1 viewed in c direction. The worth noting feature of this structure is the tunnels (∼6 Å in diameter estimated from the space-filling model) formed along the c axis. The tetramethylammonium cations locate in the tunnel (Figure S2). The structure of compound 2 is the same as that of 1, except that Mo atom is replaced by W atom (Figure S3). The two frameworks are isostructural, and their XRD spectra are identical (Figure 1a). On the basis of our knowledge, this is a first pure inorganic MOF-like interpenetrating inorganic framework with clear node and connecting-pillar characters. The supercapacitance performances of compounds 1 and 2 were evaluated by cyclic voltammetry (CV) and galvanostatic charge−discharge (GCD) techniques in KOH electrolyte with a concentration of 6.0 mol L−1,41 respectively (Figure 3, Figure S4). Figure 3a shows the typical CV curves of 1 at different scan rates with a potential window ranging from 0.15 to 0.6 V. The electroactive electrode oxidized at 0.40 V and reduced at 0.32 V versus saturated calomel electrode (SCE), showing a reversible redox reaction that is expected to exhibit quasi-capacitance. When the scan rates are increased from 5 to 150 mV s−1, the

Figure 2. (a) Assembly of the 2D framework from the Ge4S10 cluster node and [−Cu−MS4−Cu−] (M = Mo, W) pillar. (b) The interpenetrated 2D plus 2D to 3D frameworks. (c) The space-filling model of the MOF-like framework, showing the tunnel structure.

Figure 3. (a) CV curves of 1 at different scan rates. (b) The integrated area of the CV curves and the specific capacitances of 1 at different scan rates. (c) GCD curves of 1 at different current intensities. (d) Variation of specific capacitance of 1 with applied currents. C

DOI: 10.1021/acs.inorgchem.8b01168 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. Specific capacitances of 1 (a) and 2 (b) in different MOH media (M = Li, Na, K, and Cs, 3.0 mol L−1). (c) The SEM-EDS mapping images of a single crystal of 1 ion-exchanged with CsCl showing the uniform distribution of Cs(I) element and other framework elements.

To understand the effect of the electrolyte on the Csp, the supercapacitance was measured in a series of M′OH (M′ = Li, Na, K, and Cs) aqueous solutions with a concentration of 3.0 mol L−1 (nearly saturated concentration for CsOH). The Csp data of the electrodes of 1 and 2 were calculated based on CV at different scan rates, and the results are shown in Figure 4a,b. By comparing the result measured in 3 mol L−1 KOH with that in 6 mol L−1 KOH, the Csp data are increased with the increase of concentration. When the concentrations of the electrolyte solutions are constant, the Csp data are related to the cations of the M′OH, and the CsOH gives the largest one for the both electrode materials. It has been reported that some metal chalcogenides containing organic amine cations have excellent ion-exchange properties.25,45 The tetramethylammonium cations are charge-compensating ions for the anion frameworks of 1 and 2, which might be exchanged by other cations under a proper condition. To explore the relationship between the Csp and ion-exchange properties, ion-exchange experiments were conducted. After the crystals are soaked in 0.1 mol L−1 alkali chlorides for 3 d, the average percentages of the exchange ions are ∼80−90%, 20−30%, and 5−10% for Cs+, K+, and Na+, respectively, based on average atom ratios of energy-dispersive X-ray spectroscopy (EDS) data of several measurements (Figure S6). The Li+ can not exchange the ammonium cation that was confirmed by inductively coupled plasma (ICP) measurement. Compound 1 exhibited the specificity for Cs+ ion against the other cations, which might be due to the large size of the Cs+ ion that matches with that of Me4N+ cation. The order of ion exchange percentages is in accordance with that of Csp data, which might be due to the increase of the ionic conductivity. The EDS scan maps of a single crystal clearly confirmed that the Cs+ ions entered the framework with a uniform distribution (Figure 4c), and the crystal was not destroyed after the exchange. The XRD spectra of the ion-exchanged products are

shapes of the curves are maintained, and the peak currents increase as a function of the square root of scan rate, which indicates that the redox reaction kinetics may be controlled by diffusion processes (Figure S5). To highlight the electrochemical properties, Figure 3b shows the integrated area of the CV curves and the calculated capacitances at different scan rates. The specific capacitances (Csp) of the electrode material were calculated using the eq (1)42 in Supporting Information (SI) materials. The Csp values of 1 are 290 F g−1 at a scan rate of 5 mV s−1 and 174 F g−1 at a scan rate of 20 mV s−1, which are in the moderate range of the sulfide materials.38,39,43,44 The Csp values of 2 are 362 F g−1 at a scan rate of 5 mV s−1 and 208 F g−1 at a scan rate of 20 mV s−1, which are somewhat larger than those of 1 but still belong to the same order of magnitude. Figure 3c shows the GCD profiles of 1(Mo) at current densities from 1.25 to 10.0 A g−1. The galvanostatic capacitances (C′sp) at various current densities calculated from the GCD curves using the eq (2)42 in SI are presented in Figure 3d. As seen, the discharge time decreases with the increase of discharge current. The specific capacitances of 1 measured by GCD are 240 F g−1 at 1.25 A g−1 and 197 F g−1 at 2.50 A g−1. The charge time and discharge time are 102 and 190 s at 1.25 A g−1. Electrochemical performances of 2(W) were evaluated as 1 (Figure S4a−d), and the results show that 2 gave comparable data as those of 1(Mo). The C′sp data of 2 measured by GCD are 231 F g−1 at 1.25 A g−1 and 188 F g−1 at 2.50 A g−1. The charge time and discharge time are 110 and 195 s at 1.25 A g−1. Judged from the Csp data from CV and C′sp data from GCD, the capacitance properties of 1 and 2 are comparable with each other, which can be explained by their similar crystal structures. The two frameworks are isostructural, and the radius of W(VI) is only slightly larger than that of Mo(IV) (0.65 Å for W and 0.60 Å for Mo). D

DOI: 10.1021/acs.inorgchem.8b01168 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

SHELXS-2016 program, and the refinement was performed against F2 using SHELXL-2016.47 All the non-hydrogen atoms were refined anisotropically. The tetramethylammonium cations were not completely solved for 2, and the hydrogen atoms were not treated. Those disordered atoms were removed using the PLATON/SQUEEZE procedure.48 Relevant crystal data, collection parameters, and refinement results can be found in Table S1. Fabrication of the Working Electrodes and Electrochemical Measurements. The working electrodes were obtained with the following procedure. The compounds as active materials, acetylene black, and a poly(vinylene fluoride) (PVDF) emulsion, were mixed in ethanol at a weight ratio of 80:10:10, and dispersed by ultrasonic treatment for 10 min. The slurry was coated on a nickel foam (1.0 × 1.0 cm2) current collector and then pressed and dried under vacuum at 60 °C for 12 h. The mass of the active loading on the nickel foam was ∼3 mg. The electrochemical properties of the active material were tested using a CHI660E electrochemical workstation by a conventional three-electrode system with a calomel reference electrode and a platinum counter electrode. The average specific capacitances of the electrodes were calculated based on the CV curves and the discharge curves, respectively. All of the electrochemical measurements were performed in a 6.0 or 3.0 M KOH solution at room temperature. Ion exchanges of the crystals were performed by soaking the crystals in a solution of alkali chlorides in the concentration of 0.10 mol L−1 for 3 d.

provided in Figure S7. Compared with those freshly isolated, the crystals are basically unchanged after the ion changes.



CONCLUSIONS In summary, a fascinating inorganic MOF-like framework assembled by a [Ge4S10]4− cluster node and a linear [−Cu− MS4−Cu−] (M = Mo and W) inorganic pillar is reported. The network shows a twofold interpenetrated 2D plus 2D to 3D framework with 1D nano tunnels. Although inorganic open structures are well-documented, the orthogonal topology with interpenetrated framework is unusual for inorganic materials. In addition, the pure inorganic MOF-like frameworks were used as supercapacitor materials, and their capacitance behaviors were investigated. Electrochemical measurements revealed that 1 and 2 showed quasi-capacitance performance with moderate specific capacitances. The effect of alkali ions on the capacitances was evaluated by changing the electrolytes of M′OH (M′ = Li, Na, K, and Cs). The conclusion is that the Csp data are related to the size of the cations of M′OH, and the CsOH gives the best performance for the both Mo and W electrode materials. This work described a novel pure inorganic MOF-like structure and revealed the potential application of the crystal frameworks as supercapacitor materials.





ASSOCIATED CONTENT

S Supporting Information *

EXPERIMENTAL SECTION

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.8b01168. Equations, TGA data, illustrated molecular packing of 1, illustrated structures of 2, CV and GCD curves, EDS diagrams, XRD spectra, tabulated crystal data and structural refinement parameters (PDF)

Materials and Instrumentation. Starting material [(CH3)4N]4Ge4S10 was prepared according to the literature method.46 [(NH4]2MS4 (M = Mo and W) were purchased from Aldrich company. All other reagents were purchased commercially and used without further purification. The IR spectra were recorded using KBr pellets on a Nicolet Magna 550 FT-IR spectrometer. Elemental analyses of C, H, and N were performed using a VARIDEL III elemental analyzer. Solid-state room-temperature optical diffuse reflectance spectra of the microcrystal samples were obtained with a Shimadzu UV-2600 spectrometer using BaSO4 as a standard reference. Thermal analysis was conducted on a TGA-DCS 6300 microanalyzer. The samples were heated under a nitrogen stream of 100 mL min−1 with a heating rate of 20 °C min−1. Room-temperature powder X-ray diffraction data were collected on a D/MAX-3C diffractometer using a Cu tube source (Cu Kα, λ = 1.5406 Å). Fluorescence spectra of the solid samples were obtained using a Hitachi F-2500 spectrometer. Synthesis of [(CH3)4N]4[(Ge4S10)Cu4Mo2S8] (1). A mixture of [(CH3)4N]4Ge4S10 (9.1 mg, 0.01 mmol), (NH4)2MoS4 (2.6 mg, 0.01 mmol), Cu(OAc)2 (2.0 mg, 0.011 mmol), H2O (three drops), butanol (three drops), and DMF (0.55 mL) was sealed in a thick Pyrex tube. The sealed tube was heated at 100 °C for 3 d to yield red crystals that were washed with ethanol and dried in air (22% yield based on (NH4)2MoS4). The compound was sensitive against oxygen and slowly changed into black in air. Anal. Calcd for C16H48Cu4Ge4Mo2N4S18 (FW 1610.09): C, 11.94; H, 3.00; N, 3.48%. Found: C, 11.73; H, 3.31; N, 3.52%. Important IR data (cm−1): 3426 (m), 3016 (w), 1645 (m), 1478(vs), 1284 (w), 1119(m), 947 (vs),795 (s), 465 (vs), 434 (m). Synthesis of [(CH3)4N]4[(Ge4S10)Cu4W2S8] (2). Compound 2 was prepared by a similar method used in the synthesis of 1, except that the (NH4)2MoS4 was replaced by (NH4)2WS4 (3.3 mg, 0.01 mmol) (35% yield based on (NH4)2WS4). Anal. Calcd for C16H48Cu4Ge4N4S18W2 (FW 1785.91): C, 10.76; H, 2.71; N, 3.14%. Found: C, 10.79; H, 3.10; N, 3.15%. Important IR data (cm−1): 3443 (m), 3016 (w), 1627 (m), 1478(vs), 1281 (w), 1119(w), 947 (vs),795 (s), 456 (vs), 410 (m). Crystallographic Study. The diffraction data were collected at 273 K for 1 and 120 for 2 using Bruker APEX-II single-crystal X-ray diffractometer. The diffractometer is equipped with a brilliant Turbo X-ray Source with Mo Kα (λ = 0.710 75 Å) radiation and a CMOS detector. The intensities were corrected for Lorentz and polarization effects. The structure was solved with direct methods using the

Accession Codes

CCDC 1817763−1817764 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.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. (J.D.) *E-mail: [email protected]. (Q.-Y.Z.) ORCID

Qin-Yu Zhu: 0000-0003-1864-1175 Jie Dai: 0000-0002-3549-726X Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We gratefully acknowledge financial support by the NSF of China (21571136 & 21771130), the Program of the Priority Academic Program Development of Jiangsu Higher Education Institutions, and by the project of scientific and technologic infrastructure of Suzhou (SZS201708).



REFERENCES

(1) Stock, N.; Biswas, S. Synthesis of metal-organic frameworks (MOFs): routes to various MOF topologies, morphologies, and composites. Chem. Rev. 2012, 112, 933−969. (2) Dhakshinamoorthy, A.; Asiri, A. M.; García, H. Metal−organic framework (MOF) compounds: photocatalysts for redox reactions

E

DOI: 10.1021/acs.inorgchem.8b01168 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry and solar fuel production. Angew. Chem., Int. Ed. 2016, 55, 5414− 5445. (3) Wu, H. B.; Lou, X. W. Metal-organic frameworks and their derived materials for electrochemical energy storage and conversion: Promises and challenges. Sci. adv. 2017, 3, 1−16. (4) Sun, L.; Campbell, M. G.; Dinca, M. Electrically conductive porous metal-organic frameworks. Angew. Chem., Int. Ed. 2016, 55, 3566−3579. (5) Li, W.-J.; Tu, M.; Cao, R.; Fischer, R. A. Metal−organic framework thin films: electrochemical fabrication techniques and corresponding applications & perspectives. J. Mater. Chem. A 2016, 4, 12356−12369. (6) Stassen, I.; Burtch, N.; Talin, A.; Falcaro, P.; Allendorf, M.; Ameloot, R. An updated roadmap for the integration of metal− organic frameworks with electronic devices and chemical sensors. Chem. Soc. Rev. 2017, 46, 3185−3241. (7) Cao, X.; Tan, C.; Sindoro, M.; Zhang, H. Hybrid micro-/nanostructures derived from metal−organic frameworks: preparation and applications in energy storage and conversion. Chem. Soc. Rev. 2017, 46, 2660−2677. (8) Zhou, J.; Wang, B. Emerging crystalline porous materials as a multifunctional platform for electrochemical energy storage. Chem. Soc. Rev. 2017, 46, 6927−6945. (9) Salunkhe, R. R.; Kaneti, Y. V.; Yamauchi, Y. Metal-organic framework-derived nanoporous metal oxides toward supercapacitor applications: progress and prospects. ACS Nano 2017, 11, 5293− 5308. (10) Hao, L.; Ning, J.; Luo, B.; Wang, B.; Zhang, Y.; Tang, Z.; Yang, J.; Thomas, A.; Zhi, L. Structural evolution of 2D microporous covalent triazine-based framework toward the study of highperformance supercapacitor. J. Am. Chem. Soc. 2015, 137, 219−225. (11) Wu, R.; Wang, D. P.; Kumar, V.; Zhou, K.; Law, A. W. K.; Lee, P. S.; Lou, J.; Chen, Z. MOFs-derived copper sulfides embedded within porous carbon octahedra for electrochemical capacitor applications. Chem. Commun. 2015, 51, 3109−3112. (12) Zhang, P.; Guan, B. Y.; Yu, L.; Lou, X. W. Formation of doubleshelled zinc−cobalt sulfide dodecahedral cages from bimetallic zeolitic imidazolate frameworks for hybrid supercapacitors. Angew. Chem., Int. Ed. 2017, 56, 7141−7145. (13) Cao, F.; Zhao, M.; Yu, Y.; Chen, B.; Huang, Y.; Yang, J.; Cao, X.; Lu, Q.; Zhang, X.; Zhang, Z.; Tan, C.; Zhang, H. Synthesis of twodimensional CoS1.097/nitrogen-doped carbon nanocomposites using metal−organic framework nanosheets as precursors for supercapacitor application. J. Am. Chem. Soc. 2016, 138, 6924−6927. (14) Yu, X.-Y.; Yu, L.; Wu, H. B.; Lou, X. W. Formation of nickel sulfide nanoframes from metal−organic frameworks with enhanced pseudocapacitive and electrocatalytic properties. Angew. Chem., Int. Ed. 2015, 54, 5331−5335. (15) Yaghi, O. M.; Sun, Z.; Richardson, D. A.; Groy, T. L. Directed transformation of molecules to solids: synthesis of a microporous sulfide from molecular germanium sulfide cages. J. Am. Chem. Soc. 1994, 116, 807−808. (16) Cheetham, A. K.; Fèrey, G.; Loiseau, T. Open-framework inorganic materials. Angew. Chem., Int. Ed. 1999, 38, 3268−3292. and references therein. (17) Seo, Y.; Lee, S.; Jo, C.; Ryoo, R. Microporous aluminophosphate nanosheets and their nanomorphic zeolite analogues tailored by hierarchical structuredirecting amines. J. Am. Chem. Soc. 2013, 135, 8806−8809. (18) Pang, H.; Zhang, C.; Shi, D.; Chen, Y. Synthesis of a purely inorganic three-dimensional porous framework based on polyoxometalates and 4d-4f heterometals. Cryst. Growth Des. 2008, 8, 4476− 7780. (19) Zhou, Y.-Y.; Yao, S.; Yan, J.-H.; Chen, L.; Wang, T.-T.; Wang, C.-J.; Zhang, Z.-M. Design and synthesis of purely inorganic 3D frameworks composed of reduced vanadium clusters and manganese linkers. Dalton Trans. 2015, 44, 20435−20440. (20) Zhao, C.; Glass, E. N.; Chica, B.; Musaev, D. G.; Sumliner, J. M.; Dyer, R. B.; Lian, T.; Hill, C. L. All-inorganic networks and

tetramer based on tin(II)-containing polyoxometalates: tuning structural and spectral properties with lone-pairs. J. Am. Chem. Soc. 2014, 136, 12085−12091. (21) Lin, Z.-E.; Yang, G.-Y. Oxo boron clusters and their open frameworks. Eur. J. Inorg. Chem. 2011, 2011, 3857−3867. (22) Lin, H.-Y.; Chin, C.-Y.; Huang, H.-L.; Huang, W.-Y.; Sie, M.-J.; Huang, L.-H.; Lee, Y.-H.; Lin, Ch.-H.; Lii, K.-H.; Bu, X.; Wang, S.-L. Crystalline inorganic frameworks with 56-ring, 64-ring, and 72-ring channels. Science 2013, 339, 811−813. (23) Feng, P.; Bu, X.; Zheng, N. The interface chemistry between chalcogenide clusters and open framework chalcogenides. Acc. Chem. Res. 2005, 38, 293−303. (24) Kanatzidis, M. G. Beyond silica: Nonoxidic mesostructured materials. Adv. Mater. 2007, 19, 1165−1181. (25) Feng, M.-L.; Kong, D.-N.; Xie, Z.-L.; Huang, X. Y. Threedimensional chiral microporous germanium antimony sulfide with ion-exchange properties. Angew. Chem., Int. Ed. 2008, 47, 8623−8626. (26) Zhang, Z.; Zhang, J.; Wu, T.; Bu, X.; Feng, P. Threedimensional open framework built from Cu-S icosahedral clusters and its photocatalytic property. J. Am. Chem. Soc. 2008, 130, 15238− 15239. (27) Zheng, N.; Bu, X.; Feng, P. Synthetic design of crystalline inorganic chalcogenides exhibiting fast-ion conductivity. Nature 2003, 426, 428−432. (28) Wei, L.; Wei, Q.; Lin, Z.-E.; Meng, Q.; He, H.; Yang, B.-F.; Yang, G.-Y. A 3D aluminoborate open framework interpenetrated by 2D zinc−amine coordination-polymer networks in its 11-ring channels. Angew. Chem., Int. Ed. 2014, 53, 7188−7191. (29) Gong, Y.-N.; Zhong, D.-C.; Lu, T.-B. Interpenetrating metal− organic frameworks. CrystEngComm 2016, 18, 2596−2606. (30) Li, N.; Xu, J.; Feng, R.; Hu, T.-L.; Bu, X.-H. Governing metal− organic frameworks towards high stability. Chem. Commun. 2016, 52, 8501−8513. (31) Wang, C.-L.; Liu, S.-X.; Xie, L.-H.; Ren, Y.-H.; Liang, D.-D.; Sun, C.-Y.; Cheng, H.-Y. New 3D two-fold interpenetrating polyoxometallate compounds built up of dititanium-substituted Keggin polyoxotungstates and transition metals. Polyhedron 2007, 26, 3017−3022. (32) Cahill, C. L.; Parise, J. B. J. Chem. Soc. On the formation of framework indium sulfides. Dalton Trans. 2000, 1475−1482. (33) Zhang, H.; Lin, P.; Chen, E.; Tan, Y.; Wen, T.; Aldalbahi, A.; Alshehri, S. M.; Yamauchi, Y.; Du, S.; Zhang, J. Encapsulation of an interpenetrated diamondoid inorganic building block in a metalorganic framework. Chem. - Eur. J. 2015, 21, 4931−4934. (34) Sun, Y.; Lu, J.; Li, D.; Dou, J. A novel POM-based inorganic porous framework: Synthesis, structure and properties. Inorg. Chem. Commun. 2013, 36, 166−169. (35) Gieck, C.; Rocker, F.; Ksenofontov, V.; Gütlich, P.; Tremel, W. Supramolecular solid-state chemistry: interpenetrating diamond-type frameworks of U4+ ions linked by S,S’-bidentate P2S62− molecular rods in UP4S12. Angew. Chem., Int. Ed. 2001, 40, 908−910. (36) Hill, J. A.; Thompson, A. L.; Goodwin, A. L. Dicyanometallates as model extended frameworks. J. Am. Chem. Soc. 2016, 138, 5886− 5896. (37) Chippindale, A. M.; Cheyne, S. M.; Hibble, S. J. Interpenetrating copper−silver cyanometallate networks: polymorphs and topological isomers. Angew. Chem., Int. Ed. 2005, 44, 7942−7946. (38) Ramasamy, K.; Gupta, R. K.; Palchoudhury, S.; Ivanov, S.; Gupta, A. Layer-structured copper antimony chalcogenides (CuSbSexS2−x): stable electrode materials for supercapacitors. Chem. Mater. 2015, 27, 379−386. (39) Choi, K. M.; Jeong, H. M.; Park, J. H.; Zhang, Y.-B.; Kang, J. K.; Yaghi, O. M. Supercapacitors of nanocrystalline metal-organic frameworks. ACS Nano 2014, 8, 7451−7457. (40) Li, Z.-H.; Du, S.-W.; Wu, X.-T. Assembly of molecular squares and helical chain polymers of Mo(W)/Cu/S clusters using thiolato ligands as linkers. Polyhedron 2005, 24, 2988−2993. (41) Guan, B. Y.; Kushima, A.; Yu, L.; Li, S.; Li, J.; Lou, X. W. Coordination polymers derived general synthesis of multishelled F

DOI: 10.1021/acs.inorgchem.8b01168 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry mixed metal-oxide particles for hybrid supercapacitors. Adv. Mater. 2017, 29, 1605902. (42) Lu, M.; Yuan, X.-P.; Guan, X.-H.; Wang, G.-S. Synthesis of nickel chalcogenide hollow spheres using an L-cysteine-assisted hydrothermal process for efficient supercapacitor electrodes. J. Mater. Chem. A 2017, 5, 3621−3627. (43) Tang, Y.; Chen, T.; Yu, S. Morphology controlled synthesis of monodispersed manganese sulfide nanocrystals and their primary application in supercapacitors with high performances. Chem. Commun. 2015, 51, 9018−9021. (44) Tang, J.; Ge, Y.; Shen, J.; Ye, M. Facile synthesis of CuCo2S4 as a novel electrode material for ultrahigh supercapacitor performance. Chem. Commun. 2016, 52, 1509−1512. (45) Zhang, B.; Feng, M.-L.; Cui, H.-H.; Du, C.-F.; Qi, X.-H.; Shen, N.-N.; Huang, X.-Y. Syntheses, crystal structures, ion-exchange, and photocatalytic properties of two amine-directed Ge−Sb−S compounds. Inorg. Chem. 2015, 54, 8474−8481. (46) Bowes, C. L.; Lough, A. J.; Malek, A.; Ozin, G. A.; Petrov, S.; Twardowski, M.; Young, D.; et al. Dimetal linked open frameworks: [(CH3)4N]2(Ag2,Cu2)Ge4S10. Chem. Mater. 1996, 8, 2147−2152. (47) Sheldrick, G. M. SHELXS-2014, Program for structure solution; Universität of Göttingen: Germany, 2014. (48) Spek, A. L. Structure validation in chemical crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2009, D65, 148−155.

G

DOI: 10.1021/acs.inorgchem.8b01168 Inorg. Chem. XXXX, XXX, XXX−XXX