Zinc(II) and Copper(II) Hybrid Frameworks via Metal-Ion Metathesis

Nov 1, 2017 - (2) Considering MOFs as hybrid crystalline materials featuring metal-based nodes connected by organic linkers, the secondary building un...
0 downloads 6 Views 2MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: Inorg. Chem. XXXX, XXX, XXX-XXX

pubs.acs.org/IC

Zinc(II) and Copper(II) Hybrid Frameworks via Metal-Ion Metathesis with Enhanced Gas Uptake and Photoluminescence Properties Pei-Pei Cui,†,‡ Xiu-Du Zhang,† Peng Wang,† Yue Zhao,† Mohammad Azam,§ Saud I Al-Resayes,§ and Wei-Yin Sun*,† †

Coordination Chemistry Institute, State Key Laboratory of Coordination Chemistry, School of Chemistry and Chemical Engineering, Nanjing National Laboratory of Microstructures, Collaborative Innovation Center of Advanced Microstructures, Nanjing University, Nanjing 210023, China ‡ College of Life Science, Dezhou University, Dezhou 253023, China § Department of Chemistry, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia S Supporting Information *

ABSTRACT: The fabrication of metal−organic frameworks with controlled structure and desired properties is important but still a challenge. In this work, a zinc(II) framework, {[Zn3(L)2(DABCO)(H2O)]·9DMF} (named as Zn-1), has been synthesized based on [1,1′:3′,1″-terphenyl]-4,4″,5′-tricarboxylic acid (H3L) and 1,4-diazabicyclo[2.2.2]octane (DABCO), which is isostructural to the previously reported copper(II) analogue, {[Cu3(L)2(DABCO)(H2O)]· 15H2O·9DMF} (named as Cu-1). Interestingly, hybrid zinc(II) and copper(II) bimetallic frameworks have been obtained via metal-ion metathesis and found to show enhanced adsorption and photoluminescence properties. Such a postmetal-ion metathesis method can be used to synthesize new and desired frameworks that could not be obtained by direct synthesis.



INTRODUCTION Porous metal−organic frameworks (MOFs) as an emerging class of new materials have attracted remarkable attention because of their intriguing structural diversity and potential application in gas adsorption, separation and storage, photoluminescence, catalysis, and so on.1 Generally speaking, high stability is essential for the potential application of MOFs. However, the design and synthesis of porous MOFs with high stability is still a challenge.2 Considering MOFs as hybrid crystalline materials featuring metal-based nodes connected by organic linkers, the secondary building units (SBUs) and coordination geometry of metal ions that constitute the SBUs contribute remarkably to the stability of MOFs. As one particularly promising and extensively examined case, MOFs with zinc(II) paddlewheel [Zn2(OCO)4] SBUs frequently exhibit lower surface area than expected or collapse when the axial ligands or coordinated solvent molecules of the paddlewheel SBU are removed.3 However, the ZnII ion in the unstable SBU may be exchanged by other metal ions via metalion metathesis to generate more stable MOFs.4 In this way, some new isostructural MOFs that are difficult to fabricate by conventional methods can be successfully achieved.5 Hence, metal-ion metathesis has been considered to be a promising tool for the synthesis of new MOF materials. In this work, we prepared a zinc(II) MOF, {[Zn 3 (L) 2 (DABCO)(H 2 O)]·9DMF} (Zn-1; H 3 L = [1,1′:3′,1″-terphenyl]-4,4″,5′-tricarboxylic acid, DABCO = © XXXX American Chemical Society

1,4-diazabicyclo[2.2.2]octane, DMF = N,N-dimethylformamide; Scheme S1), with paddlewheel SBUs that is isostructural to the previously reported copper(II) analogue, {[Cu3(L)2(DABCO)(H2O)]·15H2O·9DMF} (Cu-1).6 The general strategy toward the construction of bimetallic MOFs is that two different metal salts were employed simultaneously as reactants during the conventional hydro/solvothermal reaction process.7 However, in this work, we achieved bimetallic zinc(II) and copper(II) MOFs via metal-ion metathesis in a single-crystal-to-single-crystal (SCSC) manner at room temperature. Furthermore, the bimetallic MOFs show improved gas adsorption and photoluminescence properties compared with the sole metal-ion MOFs.



EXPERIMENTAL SECTION

Materials. Ligand H3L and {[Cu3(L)2(DABCO)(H2O)]·15H2O· 9DMF} (Cu-1) were obtained according to the procedures reported in the literature.6,8 Synthesis of {[Zn3(L)2(DABCO)(H2O)]·9DMF} (Zn-1). Concentrated hydrochloric acid was added to a mixture of Zn(NO3)2·6H2O (44.6 mg, 0.15 mmol), H3L (36.2 mg, 0.1 mmol), and DABCO (11.2 mg, 0.1 mmol) in DMF/ethanol [16 mL; 1/1 (v/v)] until the mixture turned clear. Then, the resulting solution was sealed in a 20 mL bottle and heated at 90 °C for 3 days. After cooling to room temperature, colorless block-shaped crystals of Zn-1 were obtained in 35% yield. Received: September 5, 2017

A

DOI: 10.1021/acs.inorgchem.7b02235 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

SQUEEZE result, elemental analysis, and thermogravimetric analysis (TGA). Anal. Calcd for C81H127N13O31Zn2Cu: C, 49.30; H, 6.49; N, 9.23. Found: C, 49.35; H, 6.52; N, 9.19. IR (KBr): 3440 (s), 1659 (m), 1603 (s), 1534 (m), 1393 (s), 1181 (w), 1097 (w), 1011 (w), 863 (w), 786 (s), 749 (m), 661 (w). Synthesis of Cu-1′. Considering the similarity of Zn-1 and Cu-1, metal-ion metathesis was also tested for Cu-1. The experiments were carried out in the same way as described above, except that the crystals of Cu-1 were soaked in a 0.5 M Zn(NO3)2/DMF solution (2 mL). The color of the block-shaped crystals did not show a remarkable change upon the soaking (Figure S3). The ratio of CuII/ZnII was also determined by ICP (Table S2). When the crystals were soaked for 5 days at room temperaure, about one-third of CuII ions were exchanged by ZnII ions, upon which the sample was named as Cu-1′ (Figure S4).

Anal. Calcd for C75H99N11O22Zn3: C, 52.90; H, 5.86; N, 9.05. Found: C, 53.02; H, 5.78; N, 9.11. IR (KBr): 3379 (s), 1644 (m), 1597 (s), 1549 (m), 1388 (s), 1183 (m), 1110 (w), 861 (w), 815 (s), 783 (s), 750 (m), 714 (m), 665 (w). Synthesis of {[Zn2Cu(L)2(DABCO)(H2O)]·11DMF·7H2O} (Zn-1′). In a typical metal-ion-exchange experiment, the crystals of Zn-1 (0.02 mmol) were soaked in a 0.5 M Cu(NO3)2/DMF solution (2 mL) at room temperature. During that period, the solution was replaced with a fresh 0.5 M Cu(NO3)2 solution once a day. The color of the blockshaped crystals changed from colorless to green upon soaking (Figure 1). The metal-ion-exchanged crystals were washed with DMF/ethanol.



RESULTS AND DISCUSSION Structure Description of Zn-1. Crystallographic analysis reveals that Zn-1 has the same framework structure as that of the previously reported copper(II) analogue Cu-1.6 As shown in Figure 2a, there are two independent ZnII atoms with different coordination environments in the asymmetric unit of Zn-1. Zn1 is five-coordinated by four O atoms from carboxylate ligands and one N atom from DABCO. Two Zn1 atoms joined together by four carboxylate groups to form a [Zn2(CO2)4] paddlewheel SBU. Such adjacent SBUs linked together by axial DABCO ligands to form an infinite one-dimensional (1D) chain (Figure 2b). At the same time, each Zn2 atom is coordinated by five O atoms and two Zn2 atoms are also linked together by four carboxylate groups to give another [Zn2(CO2)4] paddlewheel SBU, in which the axial positions are occupied by H2O molecules. Two types of [Zn2(CO2)4] paddlewheel SBUs are connected together by L3− and DABCO ligands to generate a three-dimensional (3D) framework with two different kinds of channels (Figure 2b), as observed in Cu1. It is worth noting that, for each L3− ligand, two carboxylate groups at the para position link two [Zn2(CO2)4] subunits based on Zn1 and the one carboxylate group at the meta position links one [Zn2(CO2)4] subunit based on Zn2. Hence, the Zn1/Zn2 radio is 2:1. From a topological viewpoint, the framework belongs to a rare 3,4,6-connected net with the point symbol of (510·85)2(53)4(54·82) calculated by TOPOS software.6b,9 PLATON calculation suggests that the effective free volume, after removal of the coordinated water and noncoordinated solvent molecules, in Zn-1 is 71.7%, which is close to the one in Cu-1 (71.4%).6b,10 Structure Description of Zn-1′. Crystallographic analysis reveals that Zn-1′ has the same framework structure with Zn-1 as well as Cu-1 (Figure 3). The Zn1−O bond lengths are 2.035(2) Å (Zn1−O1) and 2.0213(19) Å (Zn1−O2), and the Zn1−N one is 2.067(3) Å (Table S1). The bond lengths around Zn1 are similar to those in Zn-1. Similarly, the bond distances of Cu1−O3 and Cu1−O4 are 1.9524(19) and 2.152(4) Å, respectively, which are close to those around the same position (Cu2) in Cu-1. The solvent-accessible volume of Zn-1′ is 72.3% as calculated by PLATON, which is close to those in Zn-1 and Cu-1. Metal-Ion Metathesis. From single-crystal X-ray structural analysis and the results of metal-ion metathesis, we can infer that Zn2 in Zn-1 was exchanged by CuII ions to give Zn-1′. In order to further ensure this speculation, Raman spectral measurements were carried out. In the Raman spectra of Zn1 and Zn-1′, a broad peak around 290 cm−1 corresponding to the v(Zn−N) band was observed; however, no peak at 215 cm−1 due to the v(Cu−N) band was detected (Figure S5).11,12

Figure 1. Photographs of the crystals of Zn-1 before (a) and after (b) metal-ion metathesis. The ratio of ZnII/CuII in the crystals was determined by inductively coupled plasma optical emission spectrometry (ICP-OES) measurements (Table S2). The results show that the percentage of CuII ions at first increases with prolonged time. When the crystals were soaked for 4 days, the results of ICP analysis indicated that about one-third of ZnII ions were exchanged by CuII ions. After that, at room temperature, prolonging the soaking time, the ratio no longer changed. The sample was named as Zn-1′. The exchange process was also monitored by energy-dispersive X-ray and elemental mapping, which also show the substitution of ZnII ions by CuII ions in Zn-1 (Figures S1 and S2). Fortunately, the diffraction data of Zn-1′ were successfully collected (Table 1), and the final formula {[Zn2Cu(L)2(DABCO)(H2O)]·11DMF·7H2O} was calculated from the

Table 1. Crystal Data and Structure Refinements for Zn-1 and Zn-1′ compound formula Mr cryst syst space group a (Å) b (Å) c (Å) T (K) V (Å3) Z Dc (g cm−3) M (mm−1) F(000) no. of unique reflns no. of obsd reflns [I > 2σ(I)] no. of param GOF final R indices [I > 2σ(I)]a R indices (all data)a largest difference peak and hole (e Å−3)

Zn-1 C75H99N11O22Zn3 1702.76 tetragonal I41/amd 19.660(2) 19.660(2) 59.960(7) 153(2) 23175(5) 8 0.599 0.643 4256 5503 4207 175 1.069 R1 = 0.0387, wR2 = 0.1054 R1 = 0.0632, wR2 = 0.1243 0.482 and −0.340

Zn-1′ C81H127N13O31Zn2Cu 1973.24 tetragonal I41/amd 19.3406(5) 19.3406(5) 61.432(4) 298(2) 22979.3(18) 8 0.603 0.624 4248 5398 3957 175 1.045 R1 = 0.0399, wR2 = 0.1003 R1 = 0.0672, wR2 = 0.1196 0.510 and −0.502

a R1 = ∑||Fo| − |Fc||/∑|Fo|. wR2 = |∑w(|Fo|2 − |Fc|2)|/∑|w(Fo)2|1/2, where w = 1/[σ2(Fo2) + (aP)2 + bP]. P = (Fo2 + 2Fc2)/3.

B

DOI: 10.1021/acs.inorgchem.7b02235 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 2. (a) Coordination environment of ZnII ions in Zn-1 with ellipsoids drawn at the 50% probability level. (b) 3D structure of Zn-1 with 1D chains. H atoms are omitted for clarity.

Figure 3. (a) Coordination environment of ZnII and CuII ions in Zn-1′ with ellipsoids drawn at the 50% probability level. (b) 3D structure of Zn-1′. H atoms are omitted for clarity.

PXRD studies show that the framework has no change after metal-ion metathesis (Figure S10). In addition, when the crystals of Cu-1′ were soaked in a DMF solution of 0.5 M Cu(NO3)2 for 6 days, the sample still remained crystalline, and the ratio of CuII/ZnII was 93:7, as determined by ICP, and the framework was maintained, as evidenced by PXRD (Figure S6). The results show that metal-ion metathesis of Cu-1 is also reversible, the same as Zn-1. Such selective and reversible metatheses of the Zn- and Cu-MOFs may be attributed to the similarity of the [Zn2(CO2)4] and [Cu2(CO2)4] SBUs as well as the similar coordination strength of the Zn−O and Cu−O bonds. The above results demonstrate partial metal-ion metathesis of Zn-1 and Cu-1 to give zinc(II) and copper(II) hybrid frameworks, inheriting their original framework structures.14 As described above, in Zn-1 the axial sites of the paddlewheel [Zn2(OCO)4] SBUs containing Zn1 are coordinated by DABCO ligands to form 1D chains, which support the structure of Zn-1-like pillars. However, in the paddlewheel [Zn2(OCO)4] SBUs with Zn2, the axial positions are occupied by coordinated H2O molecules. It has been reported that metal centers with coordinated solvent molecules are flexible and can be exchanged by other metal ions, and most of them can realize transmetalation at room temperature.13,14 In contrast, if the central metal ions are coordinated without solvent molecules or embedded in a stable structures, they are difficult to exchange at room temperature.15 Therefore, it is reasonable that Zn1 atoms in Zn-1 are not exchanged by CuII ions. In addition, upon

This implies that there are no Cu−N bonds in Zn-1′. Namely, Zn1 coordinated with a N atom was not exchanged; instead, Zn2 without coordination of a N-donor atom was exchanged by CuII ions. More interestingly, no metal-ion metathesis of Zn-1 occurred when other metal ions, for example, a 0.5 M DMF solution of Mg(NO3)2, Mn(NO3)2, Co(NO3)2, Ni(NO3)2, and Cd(NO3)2, were employed. Furthermore, when Zn-1 crystals were immersed in the solution with mixed salts of Cu(NO3)2 with Mg(NO3)2, Mn(NO3)2, Co(NO3)2, Ni(NO3)2, or Cd(NO3)2, only CuII ions could be incorporated into the crystals, as determined by ICP. The results indicate that metal-ion metathesis of Zn-1 occurs selectively.13 Encouraged by the above results, we further investigated the exchange behavior of Zn-1′. Fortunately, when the crystals of Zn-1′ were soaked in a DMF solution of 0.5 M Zn(NO3)2, the CuII ions were exchanged out by ZnII ions. After 6 days, the ratio of CuII/ZnII can reach 4:96, as determined by ICP measurements. That is to say, metal-ion metathesis of Zn-1 is reversible at room temperature. After two exchange processes, the framework is still stable, as confirmed the powder X-ray diffraction (PXRD) data (Figure S6). On the basis of the above experimental results and considering the similarity between Zn-1 and Cu-1, metal-ion metathesis of Cu-1 was also carried out following the procedures described above. For Cu-1, although no obvious color change was observed, the ICP data show that about onethird of the CuII ions were repleaced by ZnII ions (Table S2). C

DOI: 10.1021/acs.inorgchem.7b02235 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

framework Cu-1′ has intense emission. This result shows that the postsynthesis method can be used to make complexes with improved fluorescence. Sorption Property. For Zn-1 and Zn-1′, the solventaccessible volumes are 71.7% and 72.3%, respectively, as calculated by PLATON. As described above, Cu-1′ also still remains the same as Cu-1. The high porosity of these frameworks encourages us to examine their gas adsorption property. The samples of Zn-1, Zn-1′, and Cu-1′ were activated by the same procedure as that used for the activation of Cu-1. The TGA and PXRD data ensure the removal of solvent molecules and maintain of the framework structures for Zn-1′ and Cu-1′ (Figures S8−S13). However, in the case of Zn-1, the solvent-exchanged and desolvated samples give PXRD patterns that deviate from that of the as-prepared sample (Figure S8a), suggesting structural variation upon solvent exchange and desolvation. The 1H NMR spectrum of the digested sample (Zn-1 after activation and gas adsorption) exhibited a L3−/DABCO ligand ratio of about 2:1, which is the same as that for Zn-1 without activation (Figure S14), implying that no composition change occurred upon activation and gas adsorption. The inherent instability of the zinc(II) paddlewheel SBU upon removal of solvent molecules has been reported,4 as reflected by the PXRD patterns of Zn-1 under different conditions (Figure S8a). In order to get more detail information from the PXRD data, the Pawley fitting of PXRD of the activated Zn-1 after gas adsorption was performed.18 The results indicate that the sample for activated Zn-1 after gas sorption crystallizes in a C2 space group with different cell parameters (Figure S8b). Such structural rearrangements have been detected in the replacement or removal of coordinated solvent molecules.19 As shown in Figure 5, the adsorption of N2 at 77 K displays typical type I adsorption isotherms for activated samples of Zn-

prolonging of the metal-ion metathesis time or an increase of the concentration of CuII ions at room temperature, no exchange of Zn1 by CuII ions was observed. When metal-ion metathesis was carried out in different concentrations of the outside metal ions, for example, 0.5, 1.0, and 2.0 M, the results show the higher the concentration, the faster the metal ion metathesis (Figure S7).16 The results of this work provide an example to achieve new heterometallic hybrid frameworks via metal-ion metathesis. More importantly, Zn-1′ and Cu-1′ could not be obtained by direct synthesis because reactions of H3L and DABCO with mixed zinc(II) and copper(II) salts under the same reaction conditions just gave unidentifiable precipitates. TGA and PXRD of the Complexes. TGA was performed under a N2 atmosphere to check the thermal stability of the frameworks, and the TGA curves of Zn-1 and Zn-1′ are shown in Figures S11 and S12. Zn-1 shows a weight loss of 38.89% in the temperature range of 30−350 °C, which corresponds the loss of free solvent molecules (calcd: 38.60%), and further the framework begins to collapse. For Zn-1′, the weight loss before 215 °C is 47.04%, which can be assigned to the losses of free DMF and H2O molecules (calcd: 47.13%). The framework remains stable up to approximately 350 °C. The pure phase of the complexes was confirmed by PXRD measurements. For Zn1 and Zn-1′, the PXRD results of the as-synthesized samples are in accordance with the simulated ones (Figures S8 and S9). Photoluminescence Properties. Because of their possible emission and potential applications in chemical sensors, photochemistry, and electroluminescent displays, MOFs based on d10 transition-metal centers and organic ligands are of great interest. Thus, the photoluminescence properties of Zn-1, Zn-1′, Cu-1, and Cu-1′ were examined in the solid state at room temperature (Figure 4). The photoluminescence

Figure 4. Emission spectra of Zn-1, Zn-1′, Cu-1, and Cu-1′ in the solid state at room temperature. Figure 5. N2 (77 K) adsorption isotherms of activated Zn-1, Zn-1′, Cu-1, and Cu-1′, where filled and open shapes represent adsorption and desorption, respectively.

properties of free ligand H3L were reported to show a maximum emission at 423 nm upon excitation at 346 nm.8c Compared with luminescence of the free H3L ligand, Zn-1, Zn1′, and Cu-1′ have obvious red shifts and intense emissions at ca. 470 nm for Zn-1, 460 nm for Zn-1′, and 437 nm for Cu-1′. The dissimilarity of the emission of Zn-1, Zn-1′, and Cu-1′ may be ascribed to the difference coordination of L3− to ZnII in the complexes.17 It should be pointed out that Cu-1 has almost no emission; however, the copper(II) and zinc(II) hybrid

1′, Cu-1, and Cu-1′. It is noteworthy that the maximum value of N2 adsorption of activated Zn-1′ is 590 cm3 g−1; in contrast, only surface adsorption happened for activated Zn-1 (Figure S15a). On the basis of the N2 adsorption isotherms, the BET surface area of activated Zn-1′ is estimated to be 2063 m2 g−1, while that of activated Zn-1 is only 59 m2 g−1 (Figures S16 and D

DOI: 10.1021/acs.inorgchem.7b02235 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry S17). The enhanced gas adsorption property of activated Zn-1′ is ascribed to the high stability of the hybrid framework (vide ante). The maximum value of N2 adsorption of activated Cu-1′ is 781 cm3 g−1, which is close to that of activated Cu-1 (756 cm3 g−1). On the basis of the N2 adsorption isotherms, the BET surface area of activated Cu-1′ is calculated to be 2658 m2 g−1 (Figure S18), also close to the previously reported value of activated Cu-1.6b CO2 adsorption measurements were also conducted for the activated samples at 195 K, and the results are illustrated in Figure 6. Under the conditions of 195 K and 1 bar, the

Experimental details, PXRD patterns, TGA, ICP, and additional figures (PDF) Accession Codes

CCDC 1571790−1571791 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 Author

*E-mail: [email protected]. Fax: +86 25 83314502. ORCID

Wei-Yin Sun: 0000-0001-8966-9728 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was financially supported by the National Natural Science Foundation of China (Grants 21331002 and 21573106), the Natural Science Foundation of Shandong Province (Grant ZR2016BL01), and the Talent Introduction Project of Dezhou University (Grant 320116). The authors extend their appreciation to the International Scientific Partnership Program ISPP at the King Saud University for funding this research work through ISPP#0090. This work was also supported by a Project Funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions.

Figure 6. CO2 (195 K) adsorption isotherms of activated Zn-1, Zn-1′, Cu-1, and Cu-1′, where filled and open shapes represent adsorption and desorption, respectively.



adsorbed amount of activated Zn-1′ is 490 cm3 g−1, whereas the corresponding value of activated Zn-1 is only 52 cm3 g−1, which is 89% lower than that of activated Zn-1′ (Figure S15b). For activated Cu-1′, the maximum value of CO2 adsorption is up to 770 cm3 g−1, while the one for activated Cu-1 is 703 cm3 g−1. It is interesting to note that the large hysteresis in the CO2 adsorption only appeared in the hybrid frameworks Zn-1′ and Cu-1′. The results imply that the hybrid bimetallic frameworks not only show improved adsorption capacity (Zn-1′ ≫ Zn-1 and Cu-1′ > Cu-1) but also enhance the interactions between the frameworks and CO2 molecules, leading to hindering the escape of the adsorbed CO2.20,21



CONCLUSIONS The construction of functional MOFs with high stability and porosity has been of considerable interest. In the present work, we have successfully used metal-ion exchange via direct SCSC transformation to fabricate hybrid bimetallic MOFs with improved stability and gas-uptake capability as well as photoluminescence properties. The postsynthesis method supports a way of constructing isostructural MOFs that is not available by direct synthesis.



REFERENCES

(1) (a) Cui, Y.; Li, B.; He, H.; Zhou, W.; Chen, B.; Qian, G. MetalOrganic Frameworks as Platforms for Functional Materials. Acc. Chem. Res. 2016, 49, 483−493. (b) Kitao, T.; Zhang, Y.; Kitagawa, S.; Wang, B.; Uemura, T. Hybridization of MOFs and Polymers. Chem. Soc. Rev. 2017, 46, 3108−3133. (c) Adil, K.; Belmabkhout, Y.; Pillai, R. S.; Cadiau, A.; Bhatt, P. M.; Assen, A. H.; Maurin, G.; Eddaoudi, M. Gas/ Vapour Separation Using Ultra-Microporous Metal-Organic Frameworks: Insights into the Structure/Separation Relationship. Chem. Soc. Rev. 2017, 46, 3402−3430. (d) Lin, R.-B.; Li, L.; Wu, H.; Arman, H.; Li, B.; Lin, R.-G.; Zhou, W.; Chen, B. Optimized Separation of Acetylene from Carbon Dioxide and Ethylene in a Microporous Material. J. Am. Chem. Soc. 2017, 139, 8022−8028. (e) Li, L.; Lin, R.B.; Krishna, R.; Wang, X.; Li, B.; Wu, H.; Li, J.; Zhou, W.; Chen, B. Flexible-Robust Metal-Organic Framework for Efficient Removal of Propyne from Propylene. J. Am. Chem. Soc. 2017, 139, 7733−7736. (2) (a) Wang, B.; Lv, X.-L.; Feng, D.; Xie, L.-H.; Zhang, J.; Li, M.; Xie, Y.; Li, J.-R.; Zhou, H.-C. Highly Stable Zr(IV)-Based MetalOrganic Frameworks for the Detection and Removal of Antibiotics and Organic Explosives in Water. J. Am. Chem. Soc. 2016, 138, 6204− 6216. (b) Doonan, C.; Riccò, R.; Liang, K.; Bradshaw, D.; Falcaro, P. Metal-Organic Frameworks at the Biointerface: Synthetic Strategies and Applications. Acc. Chem. Res. 2017, 50, 1423−1432. (c) Chen, C.; Wei, Z.; Jiang, J.; Zheng, S.; Wang, H.; Qiu, Q.; Cao, C.; Fenske, D.; Su, C.-Y. Dynamic Spacer Installation for Multirole Metal-Organic Frameworks: A New Direction toward Multifunctional MOFs Achieving Ultrahigh Methane Storage Working Capacity. J. Am. Chem. Soc. 2017, 139, 6034−6307. (3) (a) Feldblyum, J. I.; Liu, M.; Gidley, D. W.; Matzger, A. J. Reconciling the Discrepancies between Crystallographic Porosity and Guest Access As Exemplified by Zn-HKUST-1. J. Am. Chem. Soc. 2011, 133, 18257−18263. (b) Kole, G. K.; Vittal, J. J. Solid-State Reactivity and Structural Transformations Involving Coordination Polymers.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b02235. E

DOI: 10.1021/acs.inorgchem.7b02235 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Chem. Soc. Rev. 2013, 42, 1755−1775. (c) Zhou, H.; Liu, B.; Wang, H.; Hou, L.; Zhang, W.; Wang, Y. Construction of Highly Porous Pillared Metal-Organic Frameworks: Rational Synthesis, Structure, and Gas Sorption Properties. Inorg. Chem. 2017, 56, 9147−9155. (4) (a) Song, X.; Jeong, S.; Kim, D.; Lah, M. S. Transmetalations in Two Metal-Organic Frameworks with Different Framework Flexibilities: Kinetics and Core-Shell Heterostructure. CrystEngComm 2012, 14, 5753−5756. (b) Wang, X. J.; Li, P. Z.; Liu, L.; Zhang, Q.; Borah, P.; Wong, J. D.; Chan, X. X.; Rakesh, G.; Li, Y. X.; Zhao, Y. L. Significant Gas Uptake Enhancement by Post-Exchange of Zinc(II) with Copper(II) within a Metal-Organic Framework. Chem. Commun. 2012, 48, 10286−10288. (c) Bajpai, A.; Chandrasekhar, P.; Govardhan, S.; Banerjee, R.; Moorthy, J. N. Single Crystal-to-Single Crystal Site-Selective Postsynthetic Metal Exchange in a Zn-MOF Based on Semi-Rigid Tricarboxylic Acid and Access to Bimetallic MOFs. Chem. - Eur. J. 2015, 21, 2759−2765. (d) Brozek, C. K.; Dincă, M. Cation Exchange at the Secondary Building Units of Metal-Organic Frameworks. Chem. Soc. Rev. 2014, 43, 5456−5467. (e) Evans, J. D.; Sumby, C. J.; Doonan, C. J. Post-synthetic Metalation of MetalOrganic Frameworks. Chem. Soc. Rev. 2014, 43, 5933−5951. (f) Zou, R.; Li, P.-Z.; Zeng, Y.-F.; Liu, J.; Zhao, R.; Duan, H.; Luo, Z.; Wang, J.G.; Zou, R.; Zhao, Y. Bimetallic Metal-Organic Frameworks: Probing the Lewis Acid Site for CO2 Conversion. Small 2016, 12, 2334−2343. (5) (a) Yao, Q. X.; Sun, J. L.; Li, K.; Su, J.; Peskov, M. V.; Zou, X. D. A Series of Isostructural Mesoporous Metal-Organic Frameworks Obtained by Ion-Exchange Induced Single-Crystal to Single-Crystal Transformation. Dalton Trans. 2012, 41, 3953−3955. (b) Yang, J.; Wang, X.; Dai, F.; Zhang, L.; Wang, R.; Sun, D. Improving the Porosity and Catalytic Capacity of a Zinc Paddlewheel Metal-Organic Framework (MOF) through Metal-Ion Metathesis in a Single-Crystalto-Single-Crystal Fashion. Inorg. Chem. 2014, 53, 10649−10653. (c) Xiao, Z.; Wang, Y.; Zhang, S.; Fan, W.; Xin, X.; Pan, X.; Zhang, L.; Sun, D. Stepwise Synthesis of Diverse Isomer MOFs via Metal-Ion Metathesis in a Controlled Single-Crystal-to-Single-Crystal Transformation. Cryst. Growth Des. 2017, 17, 4084−4089. (6) (a) Schnobrich, J. K.; Lebel, O.; Cychosz, K. A.; Dailly, A.; WongFoy, A. G.; Matzger, A. J. Linker-Directed Vertex Desymmetrization for the Production of Coordination Polymers with High Porosity. J. Am. Chem. Soc. 2010, 132, 13941−13948. (b) Hou, C.; Liu, Q.; Fan, J.; Zhao, Y.; Wang, P.; Sun, W. Y. Novel (3,4,6)-Connected MetalOrganic Framework with High Stability and Gas-Uptake Capability. Inorg. Chem. 2012, 51, 8402−8408. (c) Fan, W.; Lin, H.; Yuan, X.; Dai, F.; Xiao, Z.; Zhang, L.; Luo, L.; Wang, R. Expanded Porous MetalOrganic Frameworks by SCSC: Organic Building Units Modifying and Enhanced Gas-Adsorption Properties. Inorg. Chem. 2016, 55, 6420− 6425. (d) Wang, P.; Chen, K.; Liu, Q.; Wang, H. W.; Azam, M.; AlResayes, S. I.; Lu, Y.; Sun, W. Y. Fine Tuning of Catalytic and Sorption Properties of Metal-Organic Frameworks via in Situ Ligand Exchange. Dalton Trans. 2017, 46, 11425−11430. (7) (a) Caskey, S. R.; Matzger, A. J. Selective Metal Substitution for the Preparation of Heterobimetallic Microporous Coordination Polymers. Inorg. Chem. 2008, 47, 7942−7944. (b) Cui, Y.; Xu, H.; Yue, Y.; Guo, X.; Yu, J.; Chen, Z.; Gao, J.; Yang, Y.; Qian, G.; Chen, B. A Luminescent Mixed-Lanthanide Metal-Organic Framework Thermometer. J. Am. Chem. Soc. 2012, 134, 3979−3982. (c) Liu, Q.; Cong, H.; Deng, H. Deciphering the Spatial Arrangement of Metals and Correlation to Reactivity in Multivariate Metal-Organic Frameworks. J. Am. Chem. Soc. 2016, 138, 13822−13825. (8) (a) Hou, L.; Zhang, J. P.; Chen, X. M. Two Metal-Carboxylate Frameworks Featuring Uncommon 2D + 3D and 3-Fold-Interpenetration: (3,5)-Connected Isomeric hms and gra Nets. Cryst. Growth Des. 2009, 9, 2415−2419. (b) Qin, L.; Hu, J. S.; Li, Y. Z.; Zheng, H. G. Three New Coordination Polymers Based on One Reduced Symmetry Tripodal Linker. Cryst. Growth Des. 2011, 11, 3115−3121. (c) Li, F.; Zhao, X. L.; Cao, R. Synthesis, Crystal Structure and Photoluminescent Properties of Two Lanthanide Coordination Polymers with the Rigid Ligand of 5′-Carboxyl[1,1′:3′,1″-Terphenyl]-4,4″-Dicarboxylic Acid. Inorg. Chem. Commun. 2012, 21, 118−121.

(9) Blatov, V. A. TOPOS, A Multipurpose Crystallochemical Analysis with the Program Package; Samara State University: Samara Oblast, Russia, 2009. (10) Sarkisov, L.; Harrison, A. Computational structure characterisation tools in application to ordered and disordered porous materials. Mol. Simul. 2011, 37, 1248−1257. (11) (a) Katsenis, A. D.; Puškarić, A.; Štrukil, V.; Mottillo, C.; Julien, P. A.; Užarević, K.; Pham, M.-H.; Do, T.-O.; Kimber, S. A. J.; Lazić, P.; Magdysyuk, O.; Dinnebier, R. E.; Halasz, I.; Frišcǐ ć, T. In Situ X-ray Diffraction Monitoring of a Mechanochemical Reaction Reveals a Unique Topology Metal-Organic Framework. Nat. Commun. 2015, 6, 6662. (b) Chen, S.; Lucier, B. E. G.; Boyle, P. D.; Huang, Y. Understanding The Fascinating Origins of CO2 Adsorption and Dynamics in MOFs. Chem. Mater. 2016, 28, 5829−5846. (c) Rossini, A. J.; Zagdoun, A.; Lelli, M.; Canivet, J.; Aguado, S.; Ouari, O.; Tordo, P.; Rosay, M.; Maas, W. E.; Copéret, C.; Farrusseng, D.; Emsley, L.; Lesage, A. Dynamic Nuclear Polarization Enhanced Solid-State NMR Spectroscopy of Functionalized Metal-Organic Frameworks. Angew. Chem., Int. Ed. 2012, 51, 123−127. (12) (a) Tan, K.; Nijem, N.; Canepa, P.; Gong, Q.; Li, J.; Thonhauser, T.; Chabal, Y. J. Stability and Hydrolyzation of Metal Organic Frameworks with Paddle-Wheel SBUs upon Hydration. Chem. Mater. 2012, 24, 3153−3167. (b) Gadipelli, S.; Guo, Z. Postsynthesis Annealing of MOF-5 Remarkably Enhances the Framework Structural Stability and CO2 Uptake. Chem. Mater. 2014, 26, 6333−6338. (13) (a) Deria, P.; Mondloch, J. E.; Karagiaridi, O.; Bury, W.; Hupp, J. T.; Farha, O. K. Beyond Post-Synthesis Modification: Evolution of Metal-Organic Frameworks via Building Block Replacement. Chem. Soc. Rev. 2014, 43, 5896−5912. (b) Li, L.; Xue, H.; Wang, Y.; Zhao, P.; Zhu, D.; Jiang, M.; Zhao, X. Solvothermal Metal Metathesis on a Metal-Organic Framework with Constricted Pores and the Study of Gas Separation. ACS Appl. Mater. Interfaces 2015, 7, 25402−25412. (c) Deng, M.; Pan, Y.; Zhu, J.; Chen, Z.; Sun, Z.; Sun, J.; Ling, Y.; Zhou, Y.; Feng, P. Cation-Exchange Approach to Tuning the Flexibility of a Metal-Organic Framework for Gated Adsorption. Inorg. Chem. 2017, 56, 5069−5075. (14) (a) Song, X. K.; Kim, T. K.; Kim, H.; Kim, D.; Jeong, S.; Moon, H. R.; Lah, M. S. Post-Synthetic Modifications of Framework Metal Ions in Isostructural Metal-Organic Frameworks: Core-Shell Heterostructures via Selective Transmetalations. Chem. Mater. 2012, 24, 3065−3073. (b) Han, Y.; Li, J.-R.; Xie, Y.; Guo, G. Substitution Reactions in Metal-Organic Frameworks and Metal-Organic Polyhedral. Chem. Soc. Rev. 2014, 43, 5952−5981. (15) (a) Karagiaridi, O.; Bury, W.; Fairen-Jimenez, D.; Wilmer, C. E.; Sarjeant, A. A.; Hupp, J. T.; Farha, O. K. Enhanced Gas Sorption Properties and Unique Behavior toward Liquid Water in a PillaredPaddlewheel Metal-Organic Framework Transmetalated with Ni(II). Inorg. Chem. 2014, 53, 10432−10436. (b) Zheng, Z. K.; Opilik, L.; Schiffmann, F.; Liu, W.; Bergamini, G.; Ceroni, P.; Lee, L.; Schutz, A.; Sakamoto, J. J.; Zenobi, R.; Vandevondele, J.; Schlüter, A. D. Synthesis of Two-Dimensional Analogues of Copolymers by Site-to-Site Transmetalation of Organometallic Monolayer Sheets. J. Am. Chem. Soc. 2014, 136, 6103−6110. (16) (a) Prasad, T. K.; Hong, D. H.; Suh, M. P. High Gas Sorption and Metal-Ion Exchange of Microporous Metal-Organic Frameworks with Incorporated Imide Groups. Chem. - Eur. J. 2010, 16, 14043− 14050. (b) Li, C. P.; Chen, J.; Liu, C. S.; Du, M. Dynamic Structural Transformations of Coordination Supramolecular Systems upon Exogenous Stimulation. Chem. Commun. 2015, 51, 2768−2781. (17) (a) Shi, Z.; Pan, Z.; Jia, H.; Chen, S.; Qin, L.; Zheng, H. Zn(II)/ Cd(II) Terephthalate Coordination Polymers Incorporating Bi-, Tri-, and Tetratopic Phenylamine Derivatives: Crystal Structures and Photoluminescent Properties. Cryst. Growth Des. 2016, 16, 2747− 2755. (b) Chen, S.; Sheng, L.; Zhao, Y.; Liu, Z.; Qiao, R.; Yang, S. Syntheses, Structures, and Properties of a Series of Polyazaheteroaromatic Core-Based Zn(II) Coordination Polymers Together with Carboxylate Auxiliary Ligands. Cryst. Growth Des. 2016, 16, 229−241. (18) (a) Roque-Malherbe, R.; Uwakweh, O. N. C.; Lozano, C.; Polanco, R.; Hernandez-Maldonado, A.; Fierro, P.; Lugo, F.; PrimeraF

DOI: 10.1021/acs.inorgchem.7b02235 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry Pedrozo, J. N. Structural Effects and Interactions of Carbon Dioxide Molecules Adsorbed on Ni, Zn, and Cd Nitroprussides. J. Phys. Chem. C 2011, 115, 15555−15569. (b) Zhang, W. X.; Xue, W.; Chen, X. M. Flexible Mixed-Spin Kagomé Coordination Polymers with Reversible Magnetism Triggered by Dehydration and Rehydration. Inorg. Chem. 2011, 50, 309−316. (19) (a) Uemura, K.; Yamasaki, Y.; Onishi, F.; Kita, H.; Ebihara, M. Two-Step Adsorption on Jungle-Gym-Type Porous Coordination Polymers: Dependence on Hydrogen-Bonding Capability of Adsorbates, Ligand-Substituent Effect, and Temperature. Inorg. Chem. 2010, 49, 10133−10143. (b) Chen, L.; Mowat, J. P. S.; Fairen-Jimenez, D.; Morrison, C. A.; Thompson, S. P.; Wright, P. A.; Duren, T. Elucidating the Breathing of the Metal-Organic Framework MIL-53(Sc) with ab Initio Molecular Dynamics Simulations and in Situ X-ray Powder Diffraction Experiments. J. Am. Chem. Soc. 2013, 135, 15763−15773. (20) (a) Song, X. K.; Oh, M.; Lah, M. S. Hybrid Bimetallic MetalOrganic Frameworks: Modulation of the Framework Stability and Ultralarge CO2 Uptake Capacity. Inorg. Chem. 2013, 52, 10869− 10876. (b) Ma, Y.-X.; Li, Z.-J.; Wei, L.; Ding, S.-Y.; Zhang, Y.-B.; Wang, W. A Dynamic Three-Dimensional Covalent Organic Framework. J. Am. Chem. Soc. 2017, 139, 4995−4998. (c) Razavi, S. A. A.; Masoomi, M. Y.; Islamoglu, T.; Morsali, A.; Xu, Y.; Hupp, J. T.; Farha, O. K.; Wang, J.; Junk, P. C. Improvement of Methane-Framework Interaction by Controlling Pore Size and Functionality of Pillared MOFs. Inorg. Chem. 2017, 56, 2581−2588. (21) (a) Foo, M. L.; Matsuda, R.; Kitagawa, S. Functional Hybrid Porous Coordination Polymers. Chem. Mater. 2014, 26, 310−322. (b) Witman, M.; Ling, S.; Jawahery, S.; Boyd, P. G.; Haranczyk, M.; Slater, B.; Smit, B. The Influence of Intrinsic Framework Flexibility on Adsorption in Nanoporous Materials. J. Am. Chem. Soc. 2017, 139, 5547−5557.

G

DOI: 10.1021/acs.inorgchem.7b02235 Inorg. Chem. XXXX, XXX, XXX−XXX