Indium-Based Heterometal–Organic Frameworks ... - ACS Publications

We pay particular attention to In3+ ions because of its abundant and distinct ... Mg-MOF-74 exhibits the highest CO2 uptake capacity (228 cm3/g at 273...
0 downloads 0 Views 4MB Size
Article pubs.acs.org/crystal

Indium-Based Heterometal−Organic Frameworks with Different Nanoscale Cages: Syntheses, Structures, and Gas Adsorption Properties Published as part of a Crystal Growth and Design virtual special issue on Crystal Engineering of Nanoporous Materials for Gas Storage and Separation Yan-Jie Qi,† Dan Zhao,‡ Xin-Xiong Li,*,† Xiang Ma,† Wen-Xu Zheng,† and Shou-Tian Zheng*,† †

State Key Laboratory of Photocatalysis on Energy and Environment, College of Chemistry, Fuzhou University, Fuzhou, Fujian 350108, China ‡ Fuqing Branch of Fujian Normal University, Fuqing, Fujian 350300, China S Supporting Information *

ABSTRACT: A series of rare metal−organic frameworks based on In-M heterometallic clusters [(CH3)2NH2][In3M2(BTC)4(DMF)2(H2O)4Cl2]·solvent (1M, M = Ca, Sr, Ba, H3BTC = 1,3,5-benzenetricarboxylic acid, DMF = N,Ndimethylformamide) have been successfully synthesized. Structural analysis reveals that 1-M represent fascinating three-dimensional porous frameworks constructed from different nanoscale cages. What is more, these new materials exhibit interesting gas adsorption properties that can be tuned by encapsulating different alkaline earth metal ions.



atm),12 outshining its transition-metal analogues such as MOF74-Co and MOF-74-Ni and highlighting the unique significance of alkaline-earth metals in the design of high-performance gas sorption materials. The integration of alkaline-earth metals into indium-based MOFs provides an opportunity to construct novel materials with charming architectures and unique properties. However, up to now, limited work has been done by combining these two types of interesting metals to create heterometallic MOFs.13 Recently, we reported a heterometallic MOF (CPM-23)13a based on In3+ and Mg2+ ions which exhibits an enhanced CO2 capture capacity to a comparative degree. Inspired by this result, we decide to further explore this system. Herein, we report the syntheses, structures, and properties of a series of heterometallic MOFs [(CH3)2NH2][In3M2(BTC)4(DMF)2(H2O)4Cl2]·solvent, (1-M, M = Ca, Sr, Ba, H3BTC = 1,3,5-benzenetricarboxylic acid, DMF = N,N-dimethylformamide), which have the following interesting structural features: (1) 1-M show a series of rare MOFs based on the combination of In3+ with different alkaline-earth metals; (2) except for the integration of mixed metals, 1-M also consist of different inorganic building blocks with opposite charge properties within the same structure, [In(COO)4]− and [InM(COO)4]+; (3) 1-M exhibit intriguing three-dimensional (3D) porous frameworks built by different types of nanoscale heterometallic cages; (4) the gas uptake capacity of the isostructural materials

INTRODUCTION As a class of porous materials with well-defined structures, tunable pore sizes, and functionalized pore environments, metal−organic frameworks (MOFs) turn out to be a prospective platform for catalysis,1 proton conduction,2 chemical sensing,3 dye separation,4 and gas storage.5 Considering the existing MOFs,6 one notes that, though numerous homometallic structures have been reported, there has been relatively little progress in the synthesis of heterometallic MOFs. Heterometallic MOFs have exhibited great promise in molecular magnetism, electrochemistry, and gas absorption because of the charge transfer and synergistic effect between different metal centers.7 So far, most reported heterometallic MOFs are based on the combination of d- and f-block metals.8 In comparison, the study on the heterometallic MOFs built from mixed main group metals, such as the combination of pblock metals with p- or s-block metals, remain largely unexplored.9 Among the main-group elements, p-block In3+ ions and sblock alkaline-earth metal ions have been getting increasing attention. We pay particular attention to In3+ ions because of its abundant and distinct coordination geometry.10 In3+ is known to form various building blocks such as monomeric {In(O2CR)4}, trimeric {In3O(O2CR)6(H2O)3}, and chain-like {In(OH)∞.11 Additionally, alkaline-earth metals are another class of main-group elements that interests us. This is because alkaline-earth metals play a captivating role in CO2 adsorption. For example, Mg-MOF-74 exhibits the highest CO2 uptake capacity (228 cm3/g at 273 K, 180 cm3/g at 298 K and 1 © 2017 American Chemical Society

Received: October 20, 2016 Revised: January 23, 2017 Published: January 30, 2017 1159

DOI: 10.1021/acs.cgd.6b01538 Cryst. Growth Des. 2017, 17, 1159−1165

Crystal Growth & Design

Article

Table 1. Crystal Data and Structure Refinement for 1-M (M = Ca, Sr, Ba) name code

1-Ca

1-Sr

1-Ba

empirical formula fw (g/mol) temperature (K) cryst syst space group a, b (Å) c (Å) V (Å3) Z ρcalcd (g/cm3) F(000) crystal size (mm3) theta range (deg) limiting indices

C44H42O30N3Cl2Ca2In3 1588.33 150(2) tetragonal P42/ncm (No. 138) 19.5189(12) 24.2530(14) 9240.1(8) 4 1.109 3152 0.7 × 0.4 × 0.3 2.24−25.02 −15 ≤ h ≤ 23 −23 ≤ k ≤ 15 −11 ≤ l ≤ 28 0.0490 1.058 R1a = 0.0910 wR2b = 0.2610 R1a = 0.1093 wR2b = 0.2782

C44H42O30N3Cl2Sr2In3 1683.41 150(2) tetragonal P42/ncm (No. 138) 19.5615(10) 24.304(2) 9299.9(11) 4 1.202 3296 0.6 × 0.4 × 0.2 1.47−24.48 −22 ≤ h ≤ 8 −13 ≤ k ≤ 22 −29 ≤ l ≤ 26 0.0623 1.045 R1a = 0.0641 wR2b = 0.1560 R1a = 0.0831 wR2b = 0.1656

C44H42O30N3Cl2Ba2In3 1782.85 150(2) tetragonal P42/ncm (No. 138) 19.6511(15) 24.3346(19) 9397.2(10) 4 1.260 3332 0.6 × 0.4 × 0.3 1.68−25.24 −19 ≤ h ≤ 23 −19 ≤ k ≤ 23 −28 ≤ l ≤ 26 0.0563 1.073 R1a = 0.0536 wR2b = 0.1535 R1a = 0.0683 wR2b = 0.1627

Rint GOF R [I > 2σ] R (all data)

a R1 = Σ∥F0| − |Fc∥/Σ|F0|. bwR2 = [Σw(F02 − Fc2)2/w(F0)2]1/2, w = 1/[σ2(F02) + (xP)2 + yP], P = (F02 + 2Fc2)/3, where x = 0.102300, y = 95.068497 for 1-Ca; x = 0.080800, y = 89.889603 for 1-Sr; x = 0.093800, y = 31.315699 for 1-Ba.

room temperature and heated at 100 °C for 5 days under autogenous pressure. After the mixture was cooled naturally to room temperature, white block crystals were obtained after being washed three times with 5 mL of fresh DMF. Yield: 86% (based on H3BTC). Elemental analysis calcd (%) for C44H42O30N3Cl2Ba2In3 (Mr = 1782.85): C, 29.64; H, 2.37; N, 2.36. Found: C, 29.96; H, 2.97; N, 3.01. IR (solid ATR, ν/cm−1): 3398(m), 1703(w), 1612(s), 1554(s), 1438(s), 1360(s), 1108(w), 933(w), 758(m), 713(m), 538(w). Single-Crystal X-ray Crystallography. Crystals were selected for diffraction experiments on a Bruker APEX Duo II CCD detector at 150 K under nitrogen atmosphere with a Mo Kα radiation, λ = 0.71073 Å. The structures were solved through direct methods and refined by full-matrix least-squares refinements based on F2 using the SHELXTL software.14 The locations of disordered solvent molecules could not be confirmed, so the SQUEEZE method15 was used to dispel the contribution from guest molecules. The crystallographic data of 1-M are given in Table 1. CCDC-1494549−1494551 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/conts/retrieving. html. Powder X-ray Diffraction. Powder X-ray diffraction (PXRD) analyses were recorded on a Rigaku Ultima IV diffractometer with Cu Kα radiation (λ = 1.54051 Å). Elemental Analysis. Elemental analyses (EA) for C, H, and N were carried out on a Vario MICRO elemental analysis. Thermogravimetric Measurement. Thermal stability studies were performed on a Mettler Toledo TGA/SDTA 851e analyzer under an air-flow atmosphere with a heating rate of 10 °C/min at 30− 800 °C. IR Measurement. The IR was obtained on a Nicolet iS50 at room temperature. Gas Adsorption Analysis. Single-component gas measurements were performed with an Accelerated Surface Area and Porosimetry 2020 (ASAP 2020) surface area analyzer. All gases were used in the adsorption experiment of 99.999% purity or higher. Before the gas adsorption measurement, the samples of 1-M were completely exchanged by CH3CN after soaking 1-M in CH3CN under ambient conditions for a week with changing fresh CH3CN every day, and the CH3CN-exchanged sample was further activated by keep the sample at

1-M can be tuned by the introduction of different alkaline earth metals.



EXPERIMENTAL SECTION

Materials and Measurements. All the reactions were conducted in an autoclave under autogenous pressure. Reagent-grade-quality reactants and solvents from commercial sources were used without further purification. Synthesis of 1-Ca. In(NO3)3·4.5H2O (78.7 mg), CaCl2 (22.6 mg), H3BTC (42.4 mg), and N(n-C4H9)4Cl (28.7 mg) were dissolved in 8 mL of DMF in a 20 mL Teflon-lined stainless steel autoclave, and then the resulting solution was stirred for about 0.5 h at room temperature and heated at 100 °C for 5 days under autogenous pressure. After the mixture was cooled naturally to room temperature, white block crystals were obtained after being washed 3 times with 5 mL of fresh DMF. Yield: 75% (based on H3BTC). Elemental analysis calcd (%) for C44H42O30N3Cl2Ca2In3 (Mr = 1588.33): C, 33.27; H, 2.67; N, 2.65. Found: C, 33.46; H, 2.94; N, 2.69. IR (solid ATR, ν/ cm−1): 3405(m), 3185(m), 2932(w), 1644(s), 1567(m), 1444(s), 1366(s), 1250(m), 1108(m), 939(w), 765(m), 719(m), 668(w), 565(w), 461(w). Synthesis of 1-Sr. In(NO3)3·4.5H2O (78.7 mg), SrCl2·6H2O (53.3 mg), H3BTC (42.4 mg), and N(n-C4H9)4Cl (28.7 mg) were dissolved in 8 mL of DMF in a 20 mL Teflon-lined stainless steel autoclave, and then the resulting solution was stirred for about 0.5 h at room temperature and heated at 100 °C for 5 days under autogenous pressure. After the mixture was cooled naturally to room temperature, white block crystals were obtained after being washed three times with 5 mL of fresh DMF. Yield: 82% (based on H3BTC). Elemental analysis calcd (%) for C44H42O30N3Cl2Sr2In3 (Mr = 1683.41): C, 31.39; H, 2.51; N, 2.49. Found: C, 31.46; H, 2.82; N, 2.62. IR (solid ATR, ν/cm−1): 3415(m), 2902(w), 1637(s), 1562(m), 1441(s), 1362(s), 1246(m), 1128(m), 938(w), 763(m), 717(m), 667(w), 575(w), 458(w). Synthesis of 1-Ba. In(NO3)3·4.5H2O (78.7 mg), BaCl2·2H2O (49.4 mg), H3BTC (42.4 mg), and N(n-C4H9)4Cl (28.7 mg) were dissolved in 8 mL of DMF in a 20 mL Teflon-lined stainless steel autoclave, and then the resulting solution was stirred for about 0.5 h at 1160

DOI: 10.1021/acs.cgd.6b01538 Cryst. Growth Des. 2017, 17, 1159−1165

Crystal Growth & Design

Article

room temperature for 4 h and then at 40 °C for 8 h, keeping low vacuum conditions throughout the whole process. The cryogenic temperature of 77, 273 K for N2, H2, CO2, CH4 sorption measurements was controlled using liquid nitrogen or an ice−water bath. Virial Graph Analysis for the Enthalpy of CO2 Adsorption and Henry’s Law Selectivity.16 Isotherm data were analyzed using the virial equation: m

ln P = ln N + 1/T ∑ aiN i + i=o

n

∑ bjN j j=o

(I)

where P is pressure, N is the amount adsorbed (or uptake), T is temperature, and m and n determine the number of terms required to adequately describe the isotherm. ai and bj are parameters independent of temperature. Thus, the isosteric heat of adsorption is calculated according to m

Q st = − R ∑ aiN i i=0

(II)

where R is the universal gas constant. In this case, we adopt i = 0, 1, 2, 3 and j = 0, 1. 2. The Henry’s law constant (KH) is equal to

KH = e−a0T − b0

(III)

and the Henry’s law selectivity was calculated as

αA,B = KH,A /KH,B

(IV)

where A and B represent the two components.



RESULTS AND DISCUSSION Single-crystal X-ray diffraction analyses reveal that 1-M crystallize in the tetragonal space group P42/ncm and exhibit Figure 2. View of the diamondoid (a) and decahedral (b) cages in 1Ba; (c) the linking mode of decahedral cage with eight diamondoid cages; (d) the linking mode of diamondoid cage with eight decahedral cages. All hydrogen atoms are omitted for clarity.

Figure 1. (a−b) View of the coordination environment of In3+ and Ba2+ ions in 1-Ba; all hydrogen atoms were omitted for clarity.

an interesting 3D framework structures based on [In(COO)4]− and [InM(COO)4]+ inorganic building units (IBUs) and BTC3− linkers. Because of isostructural characteristic, 1-Ba was selected as example for structural description. There are two crystallographically independent In3+ ions (In1, In2) and one unique Ba2+ cation in the structure. The two crystallographically independent In3+ ions exhibit two different coordination geometries. In2 is chelated by four carboxylate groups from different BTC3− ligands, giving a distorted tetrahedral [In(COO)4]− unit (Figure 1a), whereas In1 is five-coordinated and defined by four O atoms from four BTC3− ligand and one terminal Cl atoms, showing a tetragonal pyramid configuration (Figure 1b). Ba2+ ion (Ba1) adopts an eight-coordinated square antiprismatic geometry and ligated by four O atoms from four different BTC3− ligands, three coordinated water ligands, and one O atom of a DMF molecule (Figure 1b). The In1 and Ba1 atoms are connected by sharing four carboxylate groups, forming a distorted heterometallic paddle-wheel IBU [InBa(COO)4]+. Each tridentate

Figure 3. View of the 3D framework of 1-Ba; all hydrogen atoms were omitted for clarity.

BTC3− ligand bonds to one [In(COO)4]− unit and two heterometallic [InBa(COO)4]+ IBUs via chelating and μ2-η2:η1 1161

DOI: 10.1021/acs.cgd.6b01538 Cryst. Growth Des. 2017, 17, 1159−1165

Crystal Growth & Design

Article

Figure 5. (a) N2 sorption isotherms at 77 K and (b) the pore size distribution of 1-M.

1.3 nm. The decahedral cage consists of six [In(COO)4]− units, six [InBa(COO)4]+ IBUs and 12 BTC3− ligands, and its edge length is about 2.3 nm (Figure 2b). In 1-Ba, each decahedral cage shares [In(COO)4]− units, [InBa(COO)4]+ IBUs, and ligands with its neighboring eight diamondoid cages (Figure 2c). Every diamondoid cage is also connected to eight decahedral cages through the sharing of its two [In(COO)4]− units, four [InBa(COO)4]+ IBUs, and four BTC3− ligands (Figure 2d). The linking of these two types of cages results in the formation of a 3D anionic framework (Figure 3). The framework possesses 1D channels with a diameter of about 0.5 nm along the c axis, which are filled with coordinated DMF molecules and water molecules. From the view of topology, the overall 3D framework can be rationalized as an unknown 3,4,4connected 3-nodal network with the Schlafli symbol of {62·82· 102}2{62·84}{63}4 by assigning the BTC3− ligands, paddle-wheel IBUs [InBa(COO)4]+, and tetrahedral [In(COO)4]− units as three-, four-, and four-connected nodes. It is noted that the M2+ ions play an important role in the formation of structures of 1−3, though the linkage of In1 and In2 ions via BTC3− ligands still can form the same 3D framework in the absence of M2+ ions from a structural view of point. The incorporation of different M2+ ions into the frameworks offers a significant route to modulate the gas uptake abilities of the final products.

Figure 4. PXRD patterns of 1-M (the simulated patterns based on single-crystal dates).

coordination modes, respectively (Figure S1). In 1-Ba, the In− O, In−Cl, and Ba−O bond lengths are in the range of 2.111(8)−2.398(12) Å, 2.391(4) Å, and 2.71(2)−2.83(2) Å, respectively. A prominent structural feature of 1-Ba is the presence of different nanoscale cages, including distorted diamondoid and decahedral cages. As shown in Figure 2a, the diamondoid cage is made up of two [In(COO)4]− units, four [InBa(COO)4]+ IBUs, and four BTC3− ligands; the overall edge length is about

Table 2. Pore volume, Surface Area, Thermodynamic CO2 Uptake Capacity and Selectivity of CO2/N2 (CO2/N2 = 15:85), CO2/ CH4 (CO2/CH4 = 50:50) for 1-M at 273 K surface areas, cm2/g 1-M

ρcrystal, g/cm3

Vp, cm3/g

BET

1-Ca 1-Sr 1-Ba

1.109 1.202 1.260

0.59 0.56 0.51

805 654 570

a

Langmuir 975 831 699

b

selectivityf

gas uptakes, cm3/g(cm3/mol) STP CO2c

N2c

d

CH4

94.33(149.83) 76.17(128.23) 68.65(122.39)

7.85(12.47) 4.37(7.36) 5.32(9.48)

12.75(20.25) 16.94(28.52) 7.22(12.87)

CO2e

Qst, kJ/mol 30.58 26.36 23.32

CO2/N2

CO2/CH4

34.9 31.1 33.6

19.3 7.1 12.2

a f

Calculated by the BET method. bCalculated by the Langmuir method. cAt 900 mmHg. dAt 760 mmHg. eCalculated by the virial method. Calculated by the Henry’s law. 1162

DOI: 10.1021/acs.cgd.6b01538 Cryst. Growth Des. 2017, 17, 1159−1165

Crystal Growth & Design

Article

Figure 6. Gas adsorption isotherms of 1-M at 273 K (left) and 298 K (right).

phase change indicated by PXRD. PLATON calculation shows that 1-Ba, 1-Sr, and 1-Ca respectively have 64.8, 65.5, and 66.1% potential guest-accessible volumes, which prompts us to explore their gas capture properties.17 Meanwhile, the estimated pores volumes of 1-Ca, 1-Sr, and 1-Ba are 0.59, 0.56, and 0.51 cm3/g (Table 2), respectively. Before the gas adsorption measurement, the samples of 1-M were completely exchanged by CH3CN after soaking 1-M in CH3CN under ambient conditions for a week with changing fresh CH3CN every day, and the CH3CN-exchanged samples was further activated by keeping the sample at room temperature for 4 h and then at 40 °C for 8 h. The PXRD patterns show that the desolvated samples retain their crystallinity well after the activation process (Figure 4). As shown in Figure 5a, the N2 sorption of 1-M exhibits an I-type adsorption isotherm typical of materials of permanent microporosity. The BET/Langmuir surface areas of 1-Ca, 1-Sr, and 1-Ba are 805/975, 654/831, and 570/699 m2/g, respectively. The median pore sizes of 5.9, 6.4, and 6.7 Å for 1-Ca, 1-Sr, and 1-Ba are calculated by The nonlocal density functional theory (NLDFT), respectively (Figure 5b).

Figure 7. CO2 adsorption enthalpies of 1-M.

The in situ variable-temperature PXRD patterns (Figure 4) and the thermogravimetric analysis (Figure S3) indicate that the phase of as-synthesized samples purity and the frameworks of 1-Ba, 1-Sr, and 1-Ca remain intact up to 140 °C, 240 °C, and 200 °C, respectively. The lowest stability of 1-Ba should be because the removal of its guest solvent molecules triggers the 1163

DOI: 10.1021/acs.cgd.6b01538 Cryst. Growth Des. 2017, 17, 1159−1165

Crystal Growth & Design



The N2, H2, CH4, and CO2 uptake capacities of 1-M were further measured at 273 K/298 K and shown in Figure 6. For CO2, the uptake amount of 1-Ba reaches 122.39 cm3/mol (68.65 cm3/g, 3.06 mmol/g) at 273 K and 74.02 cm3/mol (41.52 cm3/g, 1.85 mmol/g) at 298 K under 900 mmHg, which are comparable to well-known MOF-5 (47.11 cm3/g, 2.10 mmol/g at 296 K under 1 atm) with larger pore volume and BET/Langmiur surface areas than 1-Ba.18 The CO2 uptake capacity of the framework increases when heavy Ba2+ is replaced by light Sr2+ or even lighter Ca2+. Especially, 1-Ca shows the highest CO2 adsorption capacity of 149.83 cm3/mol (94.33 cm3/g, 4.21 mmol/g) at 273 K and 99.11 cm3/mol (55.59 cm3/g, 2.48 mmol/g) at 298 K under 900 mmHg among 1-M, which is comparable to well-known MOF HKUST-1, HAT-CTF-450 (4.4 mmol/g at 273 K/1.0 bar), COF-6 (85 cm3/g at 273 K/1.0 bar), and HOF-8 (57.3 cm3/g at 298 K/1.0 atm).19 The CO2 uptake capacities, pore volumes, and surface areas of 1-M exhibit a growing tendency from Ba2+, Sr2+, to Ca2+. Such a tendency is also observed in H2 adsorption analysis. The H2 uptake values of 1-Ca, 1-Sr, and 1-Ba at 77 K under 900 mmHg are 258.56 cm3/mol (162.79 cm3/g 7.26 mmol/g), 247.09 cm3/mol (146.78 cm3/g, 6.58 mmol/g), and 185.22 cm3/mol (103.89 cm3/g, 4.57 mmol/g), respectively (Figure S5). In order to have better insight into the influence of different alkaline earth ions on the interaction of the adsorbate with the frameworks, the coverage-dependent adsorption enthalpies of CO2 (Qst) are simulated by the virial method,20 a wellestablished and reliable methodology from fits of their adsorption isotherms at 273 and 298 K (Figure S6−S8). As shown in Figure 7, the adsorption enthalpies of 1-Ca, 1-Sr, and 1-Ba are 30.58, 26.36, and 23.32 kJ/mol at zero coverage, respectively. These results indicate that the frameworks’ affinity with CO2 is enhanced from 1-Ba, 1-Sr, to 1-Ca, which is attributed to the decreased pore sizes (Figure 5b).21 Notably, the Qst values of 1-M for CO2 are comparable to those of MIL53(Cr) (32 kJ/mol), HKUST-1 (hydrated) (30 kJ/mol), MAF2 (27 kJ/mol), JUC-132 (30 kJ/mol), JLU-Liu (30 kJ/mol), and NOTT-140 (25 kJ/mol), but higher than those of most “benchmark MOFs”, such as CuBTTri (21 kJ/mol), MOF-5 (17 kJ/mol), and UMCM-1 (12 kJ/mol).22 Interestingly, compared with the high CO2 uptake capacities, only a little CH4 is captured by 1-M at 273 and 298 K, especially, N2 and H2 are hardly adsorbed at all (Figure 6). We employ the Henry’s law to evaluate the CO2/N2 and CO2/CH4 gas separation selectivity, which is calculated from the single component isotherm data. The CO2/N2 (CO2/CH4) selectivity values for 1-Ca, 1-Sr, and 1-Ba are 34.9 (19.3), 31.1 (7.1), and 33.6 (12.2), respectively (Tables S1, 2).

Article

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.6b01538. Additional structural figures, additional characizations such as thermogravimetric analysis, IR spectra etc. (PDF) Accession Codes

CCDC 1494549−1494551 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

*(X.X.L.) E-mail: [email protected]. *(S.-T.Z.) E-mail: [email protected]. ORCID

Xin-Xiong Li: 0000-0002-9903-2699 Shou-Tian Zheng: 0000-0002-3365-9747 Funding

This work was financially supported by National Natural Science Foundations of China (Nos. 21303018, 21371033, and 21401195), the Natural Science Foundation For Young Scholars of Fujian Province (No. 2015J05041), and Projects from State Key Laboratory of Structural Chemistry of China (Nos. 20150001 and 20160020). Notes

The authors declare no competing financial interest.



REFERENCES

(1) (a) Li, Z. Y.; Schweitzer, N. M.; League, A. B.; Bernales, V.; Peters, A. W.; Getsoian, A. B.; Wang, T. C.; Miller, J. T.; Vjunov, A.; Fulton, J. L.; Lercher, J. A.; Cramer, C. J.; Gagliardi, L.; Hupp, J. T.; Farha, O. K. J. Am. Chem. Soc. 2016, 138, 1977−1982. (b) Cui, Y. J.; Li, B.; He, H. J.; Zhou, W.; Chen, B. L.; Qian, G. D. Acc. Chem. Res. 2016, 49, 483−493. (2) (a) Zhai, Q. G.; Mao, C. Y.; Zhao, X. X.; Lin, Q. P.; Bu, F.; Chen, X. T.; Bu, X. H.; Feng, P. Y. Angew. Chem., Int. Ed. 2015, 54, 7886− 7890. (b) Tang, Q.; Liu, Y. W.; Liu, S. X.; He, D. F.; Miao, J.; Wang, X. Q.; Yang, G. C.; Shi, Z.; Zheng, Z. P. J. Am. Chem. Soc. 2014, 136, 12444−12449. (c) Zhao, X.; Mao, C. Y.; Bu, X. H.; Feng, P. Y. Chem. Mater. 2014, 26, 2492−2495. (3) (a) Rao, X. T.; Song, T.; Gao, J. K.; Cui, Y. J.; Yang, Y.; Wu, C. D.; Chen, B. L.; Qian, G. D. J. Am. Chem. Soc. 2013, 135, 15559− 15564. (b) Zhang, M.; Feng, G. X.; Song, Z. G.; Zhou, Y. P.; Chao, H. Y.; Yuan, D. Q.; Tan, T. T. Y.; Guo, Z. G.; Hu, Z. G.; Tang, B. Z.; Liu, B.; Zhao, D. J. Am. Chem. Soc. 2014, 136, 7241−7244. (4) (a) Han, S. B.; Wei, Y. H.; Valente, C.; Lagzi, I.; Gassensmith, J. J.; Coskun, A.; Stoddart, J. F.; Grzybowski, B. A. J. Am. Chem. Soc. 2010, 132, 16358−16361. (b) Feng, D. W.; Liu, T. F.; Su, J.; Bosch, M.; Wei, Z. W.; Wan, W.; Yuan, D. Q.; Chen, Y. P.; Wang, X.; Wang, K. C.; Lian, X. Z.; Gu, Z. Y.; Park, J.; Zou, X. D.; Zhou, H. C. Nat. Commun. 2015, 6, 5979−5984. (c) Lan, Y. Q.; Jiang, H. L.; Li, S. L.; Xu, Q. Adv. Mater. 2011, 23, 5015−5020. (d) Ji, W. J.; Hu, M. C.; Li, S. N.; Jiang, Y. C.; Zhai, Q. G. CrystEngComm 2014, 16, 3474−3477. (5) (a) Li, J. R.; Sculley, J. L.; Zhou, H. C. Chem. Rev. 2012, 112, 869−932. (b) He, Y. B.; Zhou, W.; Qian, G. D.; Chen, B. L. Chem. Soc. Rev. 2014, 43, 5657−5678. (c) Qin, J. S.; Du, D. Y.; Li, W. L.; Zhang, J. P.; Li, S. L.; Su, Z. M.; Wang, X. L.; Xu, Q.; Shao, K. Z.; Lan, Y. Q. Chem. Sci. 2012, 3, 2114−2118. (d) Pang, J. D.; Jiang, F. L.; Wu, M. Y.; Liu, C. P.; Su, K. Z.; Lu, W. G.; Yuan, D. Q.; Hong, M. C. Nat. Commun. 2015, 6, 7575−7580.



CONCLUSION In summary, a series of rare indium-based heterometallic MOFs 1-M have been successfully prepared, representing a class of fascinating 3D porous frameworks constructed from different nanoscale cages. What is more, these new materials exhibit interesting gas adsorption properties that can be tuned by encapsulations of different alkaline earth metals. This work not only enriches the structural diversity of heterometallic MOFs but also confirms the potential for developing new classes of functional materials by a combination of various main group ions. 1164

DOI: 10.1021/acs.cgd.6b01538 Cryst. Growth Des. 2017, 17, 1159−1165

Crystal Growth & Design

Article

(6) (a) Rowsell, J. L. C.; Yaghi, O. M. J. Am. Chem. Soc. 2006, 128, 1304−1315. (b) Serre, C.; Millange, F.; Surblé, S.; Férey, G. Angew. Chem., Int. Ed. 2004, 43, 6285−6289. (c) Wang, F.; Liu, Z. S.; Yang, H.; Tan, Y. X.; Zhang, J. Angew. Chem., Int. Ed. 2011, 50, 450−453. (d) Lan, Y. Q.; Li, S. L.; Fu, Y. M.; Du, D. Y.; Zang, H. Y.; Shao, K. Z.; Su, Z. M.; Fu, Q. Cryst. Growth Des. 2009, 9, 1353−1360. (7) (a) Li, X. X.; Gong, Y. Q.; Zhao, H. X.; Wang, R. H. CrystEngComm 2014, 16, 8818−8824. (b) Cheng, J. W.; Zhang, J.; Zheng, S. T.; Zhang, M. B.; Yang, G. Y. Angew. Chem., Int. Ed. 2006, 45, 73−77. (c) Shi, P. F.; Xiong; Zhao, G. B.; Zhang, Z. Y.; Cheng, P. Chem. Commun. 2013, 49, 2338−2340. (8) (a) Zhu, X. W.; Zhou, X. P.; Li, D. Chem. Commun. 2016, 52, 6513−6516. (b) Cruz, C.; Spodine, E.; Vega, A.; Venegas-Yazigi, D.; Paredes-García, V. Cryst. Growth Des. 2016, 16, 2173−2182. (c) Zhang, H.; Yan, Z. H.; Luo, Y.; Zheng, X. Y.; Kong, X. J.; Long, L. S.; Zheng, L. S. CrystEngComm 2016, 18, 4142−4149. (d) Ding, B.; Liu, S. X.; Cheng, Y.; Guo, C.; Wu, X. X.; Guo, J. H.; Liu, Y. Y.; Li, Y. Inorg. Chem. 2016, 55, 4391−4402. (e) Saha, D.; Hazra, D. K.; Maity, T.; Koner, S. Inorg. Chem. 2016, 55, 5729−5731. (f) Wang, H.; Xu, J.; Zhang, D. S.; Chen, Q.; Wen, R. M.; Chang, Z.; Bu, X. H. Angew. Chem., Int. Ed. 2015, 54, 5966−5970. (9) (a) Yang, S. H.; Ramirez-Cuesta, A. J.; Newby, R.; Garcia-Sakai, V.; Manuel, P.; Callear, S. K.; Campbell, S. I.; Tang, C. C.; Schroder, M. Nat. Chem. 2015, 7, 121−129. (b) Alaerts, L.; Maes, M.; Giebeler, L.; Jacobs, P. A.; Martens, J. A.; Denayer, J. F. M.; Kirschhock, C. E. A.; De Vos, D. E. D. J. Am. Chem. Soc. 2008, 130, 14170−14178. (c) Banerjee, D.; Zhang, Z. J.; Plonka, A. M.; Li, J.; Parise, J. B. Cryst. Growth Des. 2012, 12, 2162−2165. (d) Mazaj, M.; Logar, N. Z. Cryst. Growth Des. 2015, 15, 617−624. (10) Schilling, L. H.; Reinsch, H.; Stock, N. In The Chemistry of Metal-Organic Frameworks: Synthesis, Characterization, and Applications; Kaskel, S., Ed.; Wiley-VCH: New York, 2016; Vol. 5, pp 105− 135. (11) (a) Zheng, S. T.; Wu, T.; Zuo, F.; Chou, C. S.; Feng, P. Y.; Bu, X. H. J. Am. Chem. Soc. 2012, 134, 1934−1937. (b) Zheng, S. T.; Mao, C. Y.; Wu, T.; Lee, S. Y.; Feng, P. Y.; Bu, X. H. J. Am. Chem. Soc. 2012, 134, 11936−11939. (c) Zheng, S. T.; Bu, J. J.; Wu, T.; Chou, C.; Feng, P. Y.; Bu, X. H. Angew. Chem., Int. Ed. 2011, 50, 8858−8862. (12) (a) Caskey, S. R.; Wong-Foy, A. G.; Matzger, A. J. J. Am. Chem. Soc. 2008, 130, 10870−10871. (b) Yang, D. A.; Cho, H. Y.; Kim, J.; Yang, S. T.; Ahn, W. S. Energy Environ. Sci. 2012, 5, 6465−6473. (13) (a) Zheng, S. T.; Wu, T.; Chou, C. S.; Fuhr, A.; Feng, P. Y.; Bu, X. H. J. Am. Chem. Soc. 2012, 134, 4517−4520. (b) Zhai, Q. G.; Bu, X. H.; Mao, C. Y.; Zhao, X.; Feng, P. Y. J. Am. Chem. Soc. 2016, 138, 2524−2527. (14) Sheldrick, G. M. SHELXTL 5.1; Bruker AXS Inc.: Madison, WI, 1997. (15) (a) Spek, A. L. J. J. Appl. Crystallogr. 2003, 36, 7−13. (b) van der Sluis, P. V. D.; Spek, A. L. Acta Crystallogr., Sect. A: Found. Crystallogr. 1990, 46, 194−201. (16) Lin, R. B.; Chen, D.; Lin, Y. Y.; Zhang, J. P.; Chen, X. M. Inorg. Chem. 2012, 51, 9950−9955. (17) Spek, A. L. PLATON; Utrecht University: The Netherlands, 2003. (18) Zhao, Z. X.; Li, Z.; Lin, Y. S. Ind. Eng. Chem. Res. 2009, 48, 10015−10020. (19) (a) Millward, A. R.; Yaghi, O. M. J. Am. Chem. Soc. 2005, 127, 17998−17999. (b) Zhu, X.; Tian, C. C.; Veith, G. M.; Abney, C. W.; Dehaudt, J.; Dai, S. J. Am. Chem. Soc. 2016, 138, 11497−11500. (c) Furukawa, H.; Yaghi, O. M. J. Am. Chem. Soc. 2009, 131, 8875− 8883. (d) Luo, X. Z.; Jia, X. J.; Deng, J. H.; Zhong, J. L.; Liu, H. J.; Wang, K. J.; Zhong, D. C. J. Am. Chem. Soc. 2013, 135, 11684−11687. (20) Lin, R. B.; Chen, D.; Lin, Y. Y.; Zhang, J. P.; Chen, X. M. Inorg. Chem. 2012, 51, 9950−9955. (21) Yang, Q.; Zhong, C.; Chen, J. F. J. Phys. Chem. C 2008, 112, 1562−1569. (22) (a) Demessence, A.; D’Alessandro, D. M.; Foo, M. L.; Long, J. R. J. Am. Chem. Soc. 2009, 131, 8784−8786. (b) Choi, J. S.; Son, W. J.; Kim, J.; Ahn, W. S. Microporous Mesoporous Mater. 2008, 116, 727−

731. (c) Mu, B.; Schoenecker, P. M.; Walton, K. S. J. Phys. Chem. C 2010, 114, 6464−6471. (d) Bourrelly, S.; Llewellyn, P. L.; Serre, C.; Millange, F.; Loiseau, T.; Férey, G. J. Am. Chem. Soc. 2005, 127, 13519−13521. (e) Liang, Z.; Marshall, M.; Chaffee, A. L. Energy Procedia 2009, 1, 1265−1271. (f) Zhang, J. P.; Chen, X. M. J. Am. Chem. Soc. 2009, 131, 5516−5521. (h) He, H.; Song, Y.; Zhang, C.; Sun, F.; Yuan, R.; Bian, Z.; Gao, L.; Zhu, G. Chem. Commun. 2015, 51, 9463−9466. (i) Wang, D.; Liu, B.; Yao, S.; Wang, T.; Li, G.; Huo, Q.; Liu, Y. Chem. Commun. 2015, 51, 15287−15289. (j) Tan, C.; Yang, S.; Champness, N. R.; Lin, X.; Blake, A. J.; Lewis, W.; Schröder, M. Chem. Commun. 2011, 47, 4487−4489.

1165

DOI: 10.1021/acs.cgd.6b01538 Cryst. Growth Des. 2017, 17, 1159−1165