Chinese Chemical Letters  2018, Vol. 29 Issue (6): 865-868   PDF    
Two-dimensional cobalt metal-organic frameworks for efficient C3H6/CH4 and C3H8/CH4 hydrocarbon separation
Weidong Fan, Yutong Wang, Zhenyu Xiao, Zhaodi Huang, Fangna Dai, Rongming Wang, Daofeng Sun    
State Key Laboratory of Heavy Oil Processing, College of Science, China University of Petroleum(East China), Qingdao 266580, China
Abstract: A Co-based two-dimensional (2D) microporous metal-organic frameworks, [Co2(TMTA)(DMF)2(H2O)2] NO3- DMF (UPC-32) has been synthesized based on 4, 4', 4"-(2, 4, 6-trimethylbenzene-1, 3, 5-triyl) tribenzoic acid (H3TMTA). UPC-32 features a 2D microporous framework exhibits high adsorption of H2 (118.2 cm3/g, 1.05 wt%, at 77 K), and adsorption heat (Qst) of CO2 (34-46 kJ/mol). UPC-32 with narrow distance between layers (3.8 Å) exhibits high selectivity of C3H6/CH4 (31.46) and C3H8/CH4 (28.04) at 298 K and 1 bar. It is the first 2D Co-MOF that showed selective separation of C3 hydrocarbon from CH4.
Key words: Metal-organic frameworks     Microporous     Two-dimensional     Selective separation    

C1 to C3 light hydrocarbons (CH4, C2H2, C2H4, C2H6, C3H6 and C3H8) are important energy resources and raw chemicals. For example, natural gases, whose main component is CH4, have been considered as the most promising alternative fuel for future vehicle transportation, while C3H8 are important chemical for the manufacture of fine chemicals such as acrylonitrile, acrylic acid, allyl alcohol, propylene oxide, isopropanol, glycerol as well as other organic chemicals. In order to fully utilize these light hydrocarbons, high quality and purity of them are essential [1-3], thus, separations of these light hydrocarbons, especially the separation of C2 and C3 hydrocarbons from CH4 are very important industrial processes [4, 5]. The traditional separation technology such as the cryogenic distillation and the pressure swing adsorption, which are based on their different vapor pressures and boiling points, are energy-intensive. Compared to the traditional separation technologies, adsorptive separation is one of the most promising low-cost and energy-efficient route. However, the traditional adsorbents, such as zeolites, have low separation coefficient owing to their less-varying porosity and low surface area. Hence, developing new-type and more effective adsorbents is critical for the effective separation of light hydrocarbons [6-9].

Metal-organic frameworks (MOFs), which are comprised of metal ions or clusters coordinated with organic ligands, have been regarded as a novel class of crystalline microporous materials with periodic network structures [10, 11]. Owing to the pores within such porous MOFs can be adjusted to maximize their size-selective sieving effects and the pore surfaces can be functionalized to direct specific recognition of small molecules, porous MOFs have been intensively investigated for applications in traditional gas storage and separation, such as N2, H2, CO2, and so on [12-16]. Recently, MOFs start to show ability in separation of light hydrocarbons [17-20]. For example, FJI-C1 [21] and FJI-C4 [22] have been reported to display high C3H8/CH4 separation performance (78.7 and 293.4). The sizes and shapes of the pores in MOFs sorbents are the foremost for their separation performance. The pore size of an adsorbent which is comparable to or slightly larger than the kinetic diameters of the adsorbate will significantly promote the separation selectivity of these light hydrocarbons. Hence, the design and synthesis of MOFs with narrow pores close to 4.0 Å are crucial for the separation of light hydrocarbons (C1-C3) as their kinetic diameters range from 3.3 Å to 4.4 Å. In addition, based on the adsorbate-surface interactions, tailoring pore surface function, such as the immobilization of polar functional groups is another effective strategy to improve the MOFs' separation.

In this communication, we present an unprecedented 2D cobalt MOF, [Co2(TMTA)(DMF)2(H2O)2]·NO3-·DMF (denoted as UPC-32) by using a functional 1, 3, 5-tris(4-carboxyphenyl)benzene ligands H3TMTA. UPC-32 exhibits 2D layered structure with permanent porosity, high adsorption of H2, and high adsorption heat (Qst) of CO2. In addition, UPC-32 shows porosity-dependent C3H6 and C3H8 uptakes and selective C3H6 and C3H8 adsorption over CH4.

The purple block-shaped crystals of UPC-32 were synthesized by the solvent thermal reaction of Co(NO3)2·6H2O and H3TMTA. The synthesis details are shown in Supporting information. Single crystal X-ray diffraction (SCXRD) of UPC-32 revealed that it crystallizes in the monoclinic system with the P21/c space group. There are two Co2+ ions (Co1 and Co2), one TMTA3- ligand, two coordinated DMF molecules and two coordinated water molecules in the asymmetric unit of UPC-32. Co1 and Co2 are connected by three carboxylic groups to form a Co2(COO)3 paddle-wheel cluster (Fig. 1a). The Co1 is coordinated by three oxygen atoms from three different carboxylic groups and one oxygen atom from water molecule with the average Co1-O bond length of 2.024 Å. The Co2 is coordinated by three oxygen atoms originating from three TMTA3- ligands, two oxygen atoms from coordinated DMF molecules and one oxygen atom from coordinated water molecule with the average Co2-O bond length of 2.039 Å. The Co2(COO)3 clusters are connected by three carboxylic groups of TMTA3- ligands to form a 2D double-layer framework. The single neighboring layer of UPC-32 stack on top of each other by π-π interactions (3.8 Å) (Fig. 1b) to form a 3D supramolecular architecture with large pores (15.8 × 15.8 × 3.8 Å3) (Fig. 1c). The potential void calculated by the PLATON [23] software is 53.7% of the total volume (2395.5 Å3 out of the 4456.3 Å3 unit cell volume). The phase purity for the bulk materials of UPC-32 has been confirmed by the PXRD analysis (Fig. S1 in Supporting information). Thermogravimetric analysis (TGA) curve display approximately 30.1% weight loss in the temperature range from 40 ℃ to 308 ℃ (Fig. S2 in Supporting information), which are assigned to the release of two coordinated DMF molecules, two coordinated water molecules and one guest DMF molecule for UPC-32. Above 308 ℃, the frameworks start to decompose.

Download:
Fig. 1. Structure of UPC-32: (a) The TMTA3- ligand and Co cluster. (b) The distance between the two layers. (c) ABAB stacking of 2D layers.

The establishment of permanent porosity is one of the important goals in MOF research [24]. The as-synthesized crystals of UPC-32 were solvent exchanged three times with dry acetone, then the samples were degassed at 298 K for one night and at 353 K for 12 h with the outgas rate of 5 mmHg/min to produce the activated samples for the gas adsorption measurements. The active phases are highly crystalline, and remain almost the same as its assynthesized phase (Fig. S1 in Supporting information). The N2 gas adsorption curve of UPC-32 at 77 K is recorded to check their porosity (Fig. 2). UPC-32 show type Ⅰ N2 adsorption isotherms, suggesting permanent micro-porosity. The N2 gas uptake is 332 cm3/g at 77 K and 1 bar. The pore volume calculated is 0.41 cm3/g, which is smaller than the theoretical pore volume (0.50 cm3/g) owing to the structural contractions during the activation. The Brunauer-Emmett-Teller (BET) surface area calculated is 1345 cm2/g.

Download:
Fig. 2. (a) The N2 sorption isotherms at 77 K, and (b) pore size distribution for UPC-32.

The H2 adsorption experiments show 118.2 cm3/g (1.05 wt%) and 83.6 cm3/g (0.75 wt%) uptakes around 1 bar at 77 and 87 K, respectively (Fig. 3a). The moderate H2 uptake for UPC-32 at 77 K is comparable to PCN-131 (0.84 wt%) and PCN-19 (0.95 wt%) [25, 26]. Moreover, the adsorption heat of H2 calculated by the Clausius-Clapeyron equation is 8.5 kJ/mol at zero coverage and decreases slowly with increasing H2 loading (Fig. 3c). These values are higher than those of some famous MOF materials, such as HKUST-1 (6.6 kJ/mol) [27], MOF-5 (5.2 kJ/mol) [28], and NOTT-122 (6.0 kJ/mol) [29].

Download:
Fig. 3. (a) The H2 sorption isotherms at 77 K and 87 K; (b) The CO2 sorption isotherms at 273 K and 298 K; (c) The adsorption heat (Qst) of H2 and CO2 for UPC-32, calculated by the Clausius-Clapeyron equation; (d) The CO2/CH4 (v/v: 50/50 and 10/90, respectively) selectivity at 298 K, calculated by the IAST method.

Since CO2 is a dominant component of greenhouse gas and a main contaminant of natural gas, it is meaningful to investigate the capacity for CO2 and selectivity of CO2/CH4. Low pressure CO2 adsorption isotherms are also measured at 273 K and 298 K (Fig. 3b). The amount of CO2 uptake for UPC-32 is 102.8 cm3/g (20.2 wt%) and 65.7 cm3/g (12.9 wt%) under 1 bar at 273 K and 298 K, respectively. Compared with the enormous MOFs, the amount of MOFs which exhibit over 20.0 wt% CO2 uptakes at 273 K and 1 bar is relatively small [30, 31]. The Qst for CO2 in UPC-32 is 46 kJ/mol calculated by the Clausius-Clapeyron equation (Fig. 3c and Table S1 in Supporting information). The predicted CO2/CH4 selectivity (for equimolar gas-phase mixtures) by IAST at 298 K under 1 bar are calculated as 6.6 for UPC-32 (Fig. 3d and Table S2 in Supporting information). It should be noted that these values are lower than Mg-MOF-74 (CO2/CH4: 105), UTSA-16 (CO2/CH4: 30) [32] and SIFSIX-3-Zn (CO2/CH4: 231) [33], but still comparable to ZIF-79 (CO2/CH4: 5.4) [34], SIFSIX-2-Cu (CO2/CH4: 5.3) and PCN-88 (CO2/CH4: 5.3) [35], making it qualified considerable candidate for CO2 capture and separation from natural gas.

Considering the small pore size and intrinsic permanent porosity of UPC-32, we have investigated its potential application for light hydrocarbons adsorption and separation. Although many MOFs with excellent gas sorption capacity (such as H2, CO2 and CH4) have been reported, only a few MOFs show high adsorption capacity and selectivity toward light hydrocarbons. To examine the adsorption and separation of light hydrocarbons for UPC-32, single component gas adsorption isotherms of UPC-32 for various light hydrocarbons (CH4, C2H2, C2H4, C2H6, C3H6, and C3H8) are performed at both 273 K and 298 K. As expected, UPC-32 can uptake a high amount of C3H8 (104.3 cm3/g), C3H6 (110.1 cm3/g), C2H6 (80.2 cm3/g), C2H4 (74.1 cm3/g), and C2H2 (85.0 cm3/g), but a relatively lower amount of CH4 (31.3 cm3/g) at 273 K and 1 bar (Fig. 4a). It should be noted that the sorption capacity of UPC-32 for C3H8 (84.9 cm3/g), C3H6 (98.2 cm3/g), C2H6 (66.1 cm3/g), C2H4 (57.2 cm3/g), C2H2 (60.6 cm3/g), and CH4 (18.9 cm3/g) at 298 K and 1 bar are higher than those of UTSA-35a and UTSA-36a (Fig. 4b and Fig. S5 in Supporting information) [36, 37]. The magnitude of the adsorption heats reveals the affinity of the pore surface toward adsorbents, which plays a significant part in determining the adsorptive selectivity [38]. To evaluate the affinity of such light hydrocarbons in UPC-32, the adsorption heats are calculated by the Clausius-Clapeyron equation. The Qsts for CH4, C2H2, C2H4, C2H6, C3H6, and C3H8 are 13.1, 6.3, 8.3, 6.6, 13.3, and 13.8 kJ/mol at zero coverage, respectively (Fig. 4c).

Download:
Fig. 4. The CH4, C2H6, C2H4, C2H2, C3H8 and C3H6 adsorption isotherms at 273 K (a) and 298 K (b) for UPC-32; (c) The Qst for CH4, C2H6, C2H4, C2H2, C3H8 and C3H6; (d) The C2H6/CH4, C2H4/CH4, C2H2/CH4, C3H8/CH4 and C3H6/CH4 (v/v: 50/50 and 10/90, respectively) selectivity at 298 K, calculated by the IAST method.

The C3 light hydrocarbons with higher adsorption heat may provide stronger affinity with a skeleton, which results in these gases being preferentially adsorbed on a skeleton of UPC-32. Thus, it may have high selectivity of C3 light hydrocarbons with respect to CH4. Therefore, the potential for separation of CH4 from C3 light hydrocarbons has been appraised by ideal solution adsorbed theory (IAST) for binary quimolar components (Fig. 4d). At 1 bar and 298 K, the selectivities of C3H8 and C3H6 with respect to CH4 are 28.0 and 31.4, which are higher than C2H2, C2H4, and C2H6 with respect to CH4 for 5.2, 4.3, and 5.8, respectively. The results indicate that UPC-32 is a prospective absorbent for effectively selective absorptive separation of CH4 from C3 hydrocarbons at room temperature. The high adsorption selectivity of C3/CH4 could be attributed to the narrow distance between layers, which is match well with the kinetic diameters of C3 light hydrocarbons.

In summary, we have developed and characterized a new microporous 2D Co-MOF (UPC-32) based on a predesigned functional ligand. UPC-32 exhibits high adsorption of H2 and high adsorption heat of CO2, it has the right pore size to maximize interactions between gases and frameworks, so it exerts high separation selectivity for C3 light hydrocarbons with respect to CH4, as shown by single component gas adsorption and selectivity calculations. These results indicate that UPC-32 could be a promising candidate for fuel gas purification and separation of light hydrocarbons in the future.

Acknowledgments

We are grateful for financial support from the National Natural Science Foundation of China (Nos. 21771191, 21571187), Taishan Scholar Foundation (No. ts201511019), the Applied Basic Research Projects of Qingdao (No. 16-5-1-95-jch) and the Fundamental Research Funds for the Central Universities (Nos. 16CX05015A, 14CX02213A).

Appendix A. Supplementary data

Supplementary data associated with this article can be found, in the online version, at https://doi.org/10.1016/j.cclet.2017.11.020.

References
[1]
A.H. Assen, Y. Belmabkhout, K. Adil, et al., Angew. Chem. Int. Ed. 54 (2015) 14353-14358. DOI:10.1002/anie.201506345
[2]
J.F. Cai, J.C. Yu, H.L. Wang, et al., Cryst. Growth Des. 15 (2015) 4071-4074. DOI:10.1021/acs.cgd.5b00675
[3]
H.M. Wen, B. Li, H.L. Wang, et al., Chem. Commun. 51 (2015) 5610-5613. DOI:10.1039/C4CC09999K
[4]
S.H. Yang, A.G. Ramirez-Cuesta, R. Newby, et al., Nat. Chem. 7 (2015) 121-129. DOI:10.1038/nchem.2114
[5]
Z.R. Jiang, J. Ge, Y.X. Zhou, et al., NPG Asia Mater. 8 (2016) 1-8.
[6]
Z.R. Herm, B.M. Wiers, J.A. Mason, et al., Science 340 (2013) 960-964. DOI:10.1126/science.1234071
[7]
A. Cadiau, K. Adil, P.M. Bhatt, et al., Science 353 (2016) 137-140. DOI:10.1126/science.aaf6323
[8]
X. Duan, C.D. Wu, S.C. Xiang, et al., Inorg. Chem. 54 (2015) 4377-4381.
[9]
B.L. Chen, S.C. Xiang, G.D. Qian, Acc. Chem. Res. 43 (2010) 1115-1124. DOI:10.1021/ar100023y
[10]
O. Benson, I. da Silva, S.P. Argent, et al., J. Am. Chem. Soc. 138 (2016) 14828-14831. DOI:10.1021/jacs.6b08059
[11]
Y. Wang, W.S. Wu, M.L. Huang, Chin. Chem. Lett. 27 (2016) 423-427. DOI:10.1016/j.cclet.2015.12.006
[12]
Y. Chen, J. Xiao, D. Lv, et al., Chem. Eng. Sci. 158 (2017) 539-544. DOI:10.1016/j.ces.2016.11.009
[13]
O. Alduhaish, B. Li, H. Arman, et al., Chin. Chem. Lett. 28 (2017) 1653-1658. DOI:10.1016/j.cclet.2017.04.025
[14]
J. Wang, D. Xie, Z. Zhang, et al., AIChE J. 63 (2017) 2165-2175. DOI:10.1002/aic.v63.6
[15]
Y.L. Hu, W.M. Verdegaal, S.H. Yu, H.L. Jiang, ChemSusChem 7 (2014) 734-737. DOI:10.1002/cssc.201301163
[16]
J.Q. Liu, W.J. Wang, Z.D. Luo, B.H. Li, D.Q. Yuan, Inorg. Chem. 56 (2017) 10215-10219. DOI:10.1021/acs.inorgchem.7b00851
[17]
E.D. Bloch, W.L. Queen, R. Krishna, et al., Science 335 (2012) 1606-1610. DOI:10.1126/science.1217544
[18]
T.L. Hu, H. Wang, B. Li, et al., Nat. Commun. 6 (2015) 7328. DOI:10.1038/ncomms8328
[19]
Y.B. He, R. Krishna, B.L. Chen, Energy Environ. Sci. 5 (2012) 9107-9120. DOI:10.1039/c2ee22858k
[20]
H. Xu, J.F. Cai, S.C. Xiang, et al., J. Mater. Chem. A 1 (2013) 9916-9921. DOI:10.1039/c3ta12086d
[21]
Y.B. Huang, Z.J. Lin, H.R. Fu, et al., ChemSusChem 7 (2014) 2647-2653. DOI:10.1002/cssc.201402206
[22]
L. Li, X.S. Wang, J. Liang, et al., ACS Appl. Mater. Interfaces 8 (2016) 9777-9781. DOI:10.1021/acsami.6b00706
[23]
A. L. Spek, PLATON-A multipurpose crystallographic tool, Utrecht University, The Netherlands, 2001.
[24]
R.A. Smaldone, R.S. Forgan, H. Furukawa, et al., Angew. Chem. Int. Ed. 49 (2010) 8630-8634. DOI:10.1002/anie.201002343
[25]
D.X. Xue, Y. Belmabkhout, O. Shekhah, et al., J. Am. Chem. Soc. 137 (2015) 5034-5040. DOI:10.1021/ja5131403
[26]
S.S. Kaye, J.R. Long, J. Am. Chem. Soc. 127 (2005) 6506-6507. DOI:10.1021/ja051168t
[27]
S.Q. Ma, H.C. Zhou, J. Am. Chem. Soc. 128 (2006) 11734-11735. DOI:10.1021/ja063538z
[28]
Q. Gao, Y.B. Xie, J.R. Li, et al., Cryst. Growth Des. 12 (2012) 281-288. DOI:10.1021/cg201059d
[29]
Y. Yan, M. Suyetin, E. Bichoutskaia, et al., Chem. Sci. 4 (2013) 1731-1736. DOI:10.1039/c3sc21769h
[30]
K. Sumida, D.L. Rogow, J.A. Mason, et al., Chem. Rev. 112 (2012) 724-781. DOI:10.1021/cr2003272
[31]
D.M. D'Alessandro, B. Smit, J.R. Long, Angew. Chem. Int. Ed. 49 (2010) 6058-6082. DOI:10.1002/anie.201000431
[32]
S. Xiang, Y. He, Z. Zhang, et al., Nat. Commun. 3 (2012) 954. DOI:10.1038/ncomms1956
[33]
P. Nugent, Y. Belmabkhout, S.D. Burd, et al., Nature. 495 (2013) 80-84. DOI:10.1038/nature11893
[34]
A. Phan, C.J. Doonan, F.J. Uribe-Romo, et al., Acc. Chem. Res. 43 (2010) 58-67. DOI:10.1021/ar900116g
[35]
J.R. Li, J. Yu, W. Lu, et al., Nat. Commun. 4 (2013) 1538. DOI:10.1038/ncomms2552
[36]
Y.B. He, Z.J. Zhang, S.C. Xiang, et al., Chem. Commun. 48 (2012) 6493-6495. DOI:10.1039/c2cc31792c
[37]
Y.B. He, Z.J. Zhang, S.C. Xiang, et al., Chem.-Eur. J. 18 (2012) 613-619. DOI:10.1002/chem.v18.2
[38]
Y.Y. Yang, Z.J. Lin, T.T. Liu, J. Liang, R. Cao, CrystEngComm 17 (2015) 1381-1388. DOI:10.1039/C4CE02163K