Next Article in Journal
NMR, RP-HPLC-PDA-ESI-MSn, and RP-HPLC-FD Characterization of Green and Oolong Teas (Camellia sinensis L.)
Next Article in Special Issue
Fluorescein Hydrazide-Appended Metal–Organic Framework as a Chromogenic and Fluorogenic Chemosensor for Mercury Ions
Previous Article in Journal
Carotenoids Do Not Protect Bacteriochlorophylls in Isolated Light-Harvesting LH2 Complexes of Photosynthetic Bacteria from Destructive Interactions with Singlet Oxygen
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Gram-Scale Synthesis of an Ultrastable Microporous Metal-Organic Framework for Efficient Adsorptive Separation of C2H2/CO2 and C2H2/CH4

1
Key Laboratory of the Ministry of Education for Advanced Catalysis Materials, College of Chemistry and Life Sciences, Zhejiang Normal University, Jinhua 321004, China
2
Department of Chemistry, Yuquan Campus, Zhejiang University, 38 Zheda Road, Hangzhou 310027, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2021, 26(17), 5121; https://doi.org/10.3390/molecules26175121
Submission received: 28 July 2021 / Revised: 21 August 2021 / Accepted: 23 August 2021 / Published: 24 August 2021
(This article belongs to the Special Issue Preparation and Application of MOF Materials)

Abstract

:
A highly water and thermally stable metal-organic framework (MOF) Zn2(Pydc)(Ata)2 (1, H2Pydc = 3,5-pyridinedicarboxylic acid; HAta = 3-amino-1,2,4-triazole) was synthesized on a large scale using inexpensive commercially available ligands for efficient separation of C2H2 from CH4 and CO2. Compound 1 could take up 47.2 mL/g of C2H2 under ambient conditions but only 33.0 mL/g of CO2 and 19.1 mL/g of CH4. The calculated ideal absorbed solution theory (IAST) selectivities for equimolar C2H2/CO2 and C2H2/CH4 were 5.1 and 21.5, respectively, comparable to those many popular MOFs. The Qst values for C2H2, CO2, and CH4 at a near-zero loading in 1 were 43.1, 32.1, and 22.5 kJ mol−1, respectively. The practical separation performance for C2H2/CO2 mixtures was further confirmed by column breakthrough experiments.

Graphical Abstract

1. Introduction

Acetylene (C2H2) is one of most important fundamental chemicals in the petrochemical and electronic industries and is mainly produced by the cracking of petroleum or oxidative coupling of methane [1]. During such processes, some impurities like carbon dioxide (CO2) and methane (CH4) are cogenerated with C2H2 and must be removed to improve the quality of the C2H2 [2,3,4]. The traditional approaches for the separation of C2H2/CO2 and C2H2/CH4 based on cryogenic distillation or solvent extraction are either of high cost-/energy consumption or associated with pollution [5]. In this context, physisorptive separation using porous solid adsorbents has attracted particular interest due to the lower cost and energy penalty [6]. Among the diverse porous solid materials, metal-organic frameworks (MOFs) are of particular interest for such demands.
In the past decades, a large number of MOFs have been synthesized for various applications including but not limited to gas separation [7,8,9,10,11,12,13,14,15,16,17], carbon capture [18,19,20,21,22], pollutant removal [23,24], catalysis [25,26], sensing [27,28], energy devices [29,30], and water harvesting [31]. While many MOFs have shown remarkable separation selectivity for relatively distinct C2H2 and CH4, it is still very challenging to separate C2H2 and CO2 due to their similar polarity, molecular shape and geometrical dimensions (3.32 × 3.34 × 5.70 Å3 for C2H2 vs. 3.18 × 3.33 × 5.36 Å3 for CO2) [32,33,34,35]. Furthermore, most MOFs are sensitive to moisture, which renders them difficult to use under practical conditions where they are exposed to humidity [36]. Therefore, it is urgent to develop water and thermally stable MOFs to facilitate the application of MOFs in gas separation. Besides, another important question that most chemists have neglected is that whether the MOFs can be synthesized on a large scale while maintaining the separation performance [37]. While a few hundred milligram samples are enough for characterization and property measurement in the laboratory, one-pot gram scale synthesis is necessary for practical use.
Herein, we would like to report a highly water and thermal stable MOF Zn2(Pydc)(Ata)2 (1, H2Pydc = 3,5-pyridinedicarboxylic acid; HAta = 3-amino-1,2,4-triazole) that is constructed by inexpensive commercially available ligands and zinc ions for efficient separation of C2H2 from CH4 and CO2. In our work, 1 was synthesized in a one-pot reaction in a flask with production of 8.6 g. PXRD patterns confirmed its pure phase with identical structures to that of small-scale synthesis in a autoclave. Static single component gas adsorption isotherms showed that 1 could take up 54.9 cm3/g of C2H2 under 278 K and 1 bar, but only 43.1 cm3/g of CO2 and 27.8 cm3/g of CH4 under the same conditions. The selectivities calculated by ideal absorbed solution theory (IAST) were 5.1 and 21.5 for equimolar C2H2/CO2 and C2H2/CH4 mixtures, respectively. To confirm the practical separation performance of C2H2/CO2, column breakthrough experiments were carried out for a C2H2/CO2 (50/50) mixture. CO2 broke out from the bed packed with 1 at 79 min while C2H2 was retained in the column for 116 min, indicating the good dynamic separation performance of 1 for C2H2/CO2.

2. Results and Discussion

The synthesis and crystal structure of Zn2(Pydc)(Ata)2 (1) was first reported by Sun et al. [38]. Compound 1 features uncoordinated pyridyl and amino groups on the pore surface that can serve as recognition sites for interactions with guest molecules and is highly stable up to 400 °C and tolerant of acidic (pH ≥ 2) and basic (pH ≤ 14) aqueous solution conditions, which attracted our interest for the investigation of its application for selective gas separation. In the literature, 1 was synthesized on a very small scale (0.5 mmol) and the yield was specified. Although the stoichiometric ratio of Zn2+:Pydc2−:Ata in 1 is 2:1:2, the authors used a substrate ratio of 1:1:1 for the reaction. During the scale-up synthesis (34 mmol) of 1, we found that the reported 1:1:1 substrate ratio was not reliable, as it led to a large amount of residual H2Pydc and a huge reduction of yield. Once we used a Zn(NO3)2·6H2O:H2Pydc:HAta = 2:1:2 ratio, no H2Pydc residue was observed and the yield was increased to 89% (Scheme 1A). The purity was confirmed by comparing the PXRD patterns of the as-synthesized powder and the one simulated from the single crystal structure (Figure S2). The single crystal structure of 1 revealed that it belongs to the tetragonal space group I4/m, exhibiting a 3D microporous framework. Two carboxylate groups from Pydc2− are paired with two different Zn2+ ions, leaving the N sites free (Scheme 1B). For Ata, all three N atoms of the triazole ring are coordinated to three different Zn2+ ions while the −NH2 group is free (Scheme 1C). Looking at the coordination environment of the Zn2+ ion, one can find that a single Zn2+ ion is coordinated to two oxygen atoms and three nitrogen atoms (Scheme 1D).
Along the c axis, 1 features two different 1D channels characterized by Zn⋯Zn distances of 6.135 and 5.914 Å (Figure 1A). However, the small channels are not accessible by guest molecules if the flexibility of the ligands is not considered, as indicated by the Connolly surface analysis which gives a Connolly radius of 1.2 Å (Figure 1B). The average diameter of the large channels decorated with electronegative nitrogen sites is 3.6 Å (Figure 1B,C), which is slightly larger than the kinetic diameter of C2H2 (3.3 Å), CO2 (3.3 Å) but slightly smaller than that of CH4 (3.8 Å), thus suggesting its use a as a potential material for selective C2H2/CO2 and C2H2/CH4.
To explore the separation performance of 1, N2 adsorption and desorption isotherms were first collected at 77 K after activation of 1 under vacuum at 200 °C for 12 h. Figure 2A shows that 1 could take up 153 mL/g of N2 at P/P0 = 0.95 and the adsorption isotherm belonged to typical type I. The pore volume and Brunauer-Emmett-Teller (BET) surface area were calculated to be 0.24 cm3/g and 636 m2/g, respectively, close to reported values for crystals [38]. Then, C2H2, CO2, and CH4 adsorption/desorption isotherms in 1 at 278 K, 288 K, and 298 K were collected, which all exhibited type I isotherms with negligible hysteresis (Figure 2B–D). The capacities of C2H2, CO2, and CH4 in 1 at 278 K were 54.9, 43.1, and 27.8 mL/g, respectively, which decreased to 52.6, 35.3, and 23.7 mL/g at 288 K and 47.2, 33.0, and 19.1 mL/g at 298 K. The isotherms were fitted using the dual-site Langmuir–Freundlich equation (Tables S1–S3), and the IAST (ideal absorbed solution theory) selectivities for equimolar C2H2/CO2 and C2H2/CH4 mixtures were calculated and shown in Figure 3A–C. For the C2H2/CO2 mixture, the IAST selectivities were 5.1, 5.3, and 3.9 under 278, 288, and 298 K at 100 kPa, which are comparable to those of many popular MOFs such as BSF-1 (3.3, 298 K) [39], ZJNU-109 (3.6, 288 K) [40], BSF-2 (5.1, 298 K) [41], UTSA-68 (3.4, 298 K) [42] as well as many others shown in Table 1. The IAST selectivities for the C2H2/CH4 mixture were much higher and the values were 21.5, 19.6, and 15.9 under 278, 288, and 298 K at 100 kPa, respectively. These good selectivities of C2H2 over CO2 and CH4 originate from the higher affinity of 1 towards C2H2 by N⋯H-C hydrogen bonding. The IAST selectivities for C2H2/CO2 and C2H2/CH4 mixtures in 1 under 100 kPa with different C2H2 ratios were also calculated, which indicated that the composition had a larger influence on the C2H2/CH4 mixture, but whether the influence was positive or negative depended on the temperature. At 298 and 278 K, the increasing C2H2 ratio led to decreased selectivity while it decreased first and increased afterwards with the increase of the C2H2 ratio at 288 K. The isosteric heats of adsorption (Qst) were calculated using the Clausius–Clapeyron equation.
The Qst values for C2H2, CO2, and CH4 at a near-zero loading in 1 were calculated to be 43.1, 32.1, and 22.5 kJ mol−1, respectively (Figure 4A). These values were consistent with the adsorption isotherms showing that 1 accommodated C2H2 more favorably than CO2 and CH4.
To investigate its recyclability as well as regeneration conditions, cycling C2H2 adsorption-desorption experiments were conducted on 1. Figure 4B shows that 1 could be easily regenerated under vacuum at a mild temperature for 30 min. In detail, complete capacity could be realized under exposure to vacuum at temperatures above 50 °C for 30 min while >98% uptake was achieved under vacuum regeneration conditions at 25 °C for 30 min.
To further evaluate the dynamic separation of C2H2/CO2 gas mixture, breakthrough experiments were conducted in which an equimolar C2H2/CO2 mixture was flowed over a packed bed of 1 with a total flow of 2 mL/min at 298 K. Figure 5A shows the breakthrough curves of 1 for C2H2/CO2. Compared with C2H2, CO2 eluted first, which could be explained by the favorable affinity of 1 for C2H2, in agreement with the single-component adsorption isotherms in Figure 2D. The typical roll-up of the CO2 curves also revealed the weaker affinity of 1 for CO2. The regeneration of 1 with Ar purge was also studied. At 100 °C and with a Ar flowrate of 3 mL/min, nearly all of the adsorbed C2H2 and CO2 can be blown out within 3 h. These experiments confirmed the excellent potential of 1 for practical separation of C2H2/CO2 mixtures as well as the facile regeneration conditions of the material for consecutive use.

3. Materials and Methods

3.1. Materials

All the materials were used as received without further purification. Zinc nitrate hexahydrate [Zn(NO3)2·6H2O], (98% purity) was purchased from Sinopharm (Shanghai, China). 3,5-Pyridinedicarboxylic acid [H2Pydc] (98% purity) and 3-amino-1,2,4-triazole [HAta] (99% purity) were purchased from Energy Chemical (Shanghai, China). N,N-Dimethylformamide [HCON(CH3)2] (99.5% purity) was purchased from Chinasun Speciality Products Co. (Jinhua, China).

3.2. Synthesis of Zn2(Pydc)(Ata)2 1

The method for the small scale synthesis of 1 in a autoclave can be found in reference [38] and is summarized here. To a 25 mL autoclave were added 148.7 mg (0.5 mmol) of Zn(NO3)2·6H2O, 42.0 mg (0.5 mmol) of HAta, 83.6 mg (0.5 mmol) of H2Pydc, followed by a stir bar. Then 12 mL of DMF and 1 mL of H2O were added. The mixture was stirred under room temperature for 10 min whereupon almost all of the solid was dissolved. Then the stir bar was removed and the autoclave was heated to 100 °C and kept at this temperature for 72 h. When cooled to room temperature, slightly yellow crystals were obtained. Then the crystals were collected by filtration, washed with DMF (3 mL × 2) and dried under vacuum at 60 °C for 24 h. According to [38], the obtained solid product has a composition of Zn2(Pydc)(Ata)2·DMF·2H2O. Only by heating at 200 °C under high vacuum can the solvent guest molecules in the pores be removed.
For the 1 g synthesis of Zn2(Pydc)(Ata)2 the following procedure was performed: A 250 mL round-bottomed flask was charged with Zn(NO3)2·6H2O (1.785 g, 6 mmol), HAta (504.5 mg, 6 mmol) and H2Pydc (1.003 g, 6 mmol) and a stir bar. Then 144 mL of DMF and 12 mL of H2O were added and the mixture was stirred at 25 °C for about 10 min. Then, the flask bottle was equipped with a condenser, the stirring bar was taken out and the temperature of the oil bath was raised to 100 °C and the heating continued for 72 h. After that, the oil bath was removed, and the round-bottomed flask was allowed to cool to ambient temperature. The resulting powder was collected by filtration, washed with DMF (30 mL), and dried under vacuum at 60 °C for 24 h. The weight after workup was 1.0 g with a yield of 29% based on Zn(NO3)2·6H2O.
For the 10 g synthesis of Zn2(Pydc)(Ata)2 without increasing the solvent amount a 250 mL round-bottom flask was charged with Zn(NO3)2·6H2O (10.11 g, 34 mmol), HAta (2.86 g, 34 mmol) and H2Pydc (5.68 g, 34 mmol) and a stir bar. 144 mL of DMF and 12 mL of H2O were added, and the mixture was stirred at 25 °C for about 10 min. Then, the temperature of the oil bath was raised to 100 °C. The flask was equipped with a condenser and the heating was continued for 72 h. The stir bar was not removed for improving the reaction efficiency as the substrate concentrations on this scale were much higher. After that, the oil bath was removed and the flask was allowed to cool to ambient temperature. The resulting powder was collected by filtration and washed with DMF (15 mL × 6). During the washing process, two different kinds of solid with different density were observed and separated. One was white (lower density) and the other was slightly yellow (higher density). After analysis, the white one was found to be H2Pydc, while the slightly yellow one was product with > 95% purity (see Figure S2). Both solids were dried under vacuum at 60 °C for 24 h. The weights of recovered H2Pydc and yellow product were 2.96 g and 4.33 g, respectively. The yield was 45% based on Zn(NO3)2·6H2O.
Using the improved ratio of starting materials an optimized 10 g synthesis of Zn2(Pydc)(Ata)2 was performed as follows: A 250 mL round-bottomed flask was charged with Zn(NO3)2·6H2O (10.11 g, 34 mmol), HAta (2.86 g, 34 mmol) and H2Pydc (2.84 g, 17 mmol) and a stir bar. 144 mL of DMF and 12 mL of H2O were added, and the mixture was stirred at 25 °C for about 10 min. Then, the flask was equipped with a condenser, the temperature of the oil bath was raised to 100 °C and the heating was continued for 72 h. After that, the oil bath was removed, and the flask was allowed to cool to ambient temperature. The resulting powder was collected by filtration, washed with DMF (15 mL × 6). and dried under vacuum at 60 °C for 24 h. The weight after workup was 8.62 g with a yield of 89% based on Zn(NO3)2·6H2O.

3.3. Characterization

Powder X-ray diffraction (PXRD) data were collected on an AXS D8-Advance diffractometer (Bruker, Ettlingen, Germany) (Cu Kαλ = 1.540598 Ǻ) with an operating power of 40 KV, 30 mA and a scan speed of 4.0°/min. The range of 2θ was from 5° to 50°.

3.4. Adsorption Measurements

The gas adsorption measurements were performed on an Autosorb iQ instrument (Quantachrome, Florida, USA). Before gas adsorption measurements, fresh 1 was evacuated at 200 °C for 12 h until the pressure dropped below 7 μmHg. The adsorption isotherms of C2H2, CO2, and CH4 were all collected at 278, 288, and 298 K on activated samples.

3.5. Calculation for Adsorption Selectivity and Isosteric Heat of Adsorption

The adsorption isotherms in 1 were fitted using a dual-site Langmuir-Freundlich model:
q = q A ,   sat b A p v A 1 + b A p v A +   q B ,   sat b B p v B 1 + b B p v B
Here, p (unit: kPa) is the pressure of the bulk gas at equilibrium with the adsorbed phase, q (unit: L kg−1) is the adsorbed gas volume per mass of adsorbent, qA,sat an qB,sat (unit: L kg−1) are the saturation capacities of site A and B, bA and bB (unit: kPav) are the affinity coefficients of site A and B, and vA and vB represent the deviations from an ideal homogeneous surface.
The isosteric heat of adsorption, Qst, is calculated based on the Clausius-Clapeyron equation:
Q st =   RT 2   ( ln p T ) q
The IAST adsorption selectivity for two gases is defined as:
S ads =   q 1 / q 2 p 1 / p 2
where q1, and q2 are the equilibrium gas uptake from the adsorbed phase with partial pressures p1, and p2

3.6. Breakthrough Experiments

The breakthrough experiment was conducted at 298 K on a self-constructed separation setup equipped with a stainless steel column (Φ 4.6 mm × 100 mm). The weight of activated 1 packed in the column was 1.6913 g. The column packed with sample was first purged with a Ar flow (10 mL min−1) for 12 h at 100 °C for activation. A C2H2/CO2 = 1/1 (v/v) gas mixture was then introduced at 2 mL min−1. The flow rates of gases were regulated by mass flow controllers and outlet gas concentration was monitored by gas chromatography (GC-9860-5CNJ, Hope, Nanjing, China) using a thermal conductivity detector (TCD) with a detection limit of 100 ppm). After the breakthrough experiment, the column was regenerated with an Ar flow (3 mL min−1) at 100 °C for 5 h.

4. Conclusions

In conclusion, a large-scale synthetic route to Zn2(Pydc)(Ata)2 (1), a highly water and thermally stable MOF, was reported. The synthesized material was used in the separation of C2H2 from CH4 and CO2 with relatively high IAST selectivities for equimolar C2H2/CO2 (5.1) and C2H2/CH4 (21.5) gas mixtures, respectively, comparable to those of many popular MOFs. The practical separation performance for C2H2/CO2 mixture was confirmed by dynamic breakthrough experiments. In addition, 1 can be easily regenerated by vacuum or Ar purging, underscoring its potential use for C2H2/CO2 (5.1) and C2H2/CH4 (21.5) separation in industry.

Supplementary Materials

The following are available online. Figure S1: Photographs illustrating the reaction process, Figure S2: PXRD patterns comparison, Table S1: Langmuir-Freundlich parameters fit for C2H2, CO2, and CH4 in Zn2(Pydc)(Ata)2 at 298 K, Table S2: Langmuir-Freundlich parameters fit for C2H2, CO2, and CH4 in Zn2(Pydc)(Ata)2 at 288 K, Table S3: Langmuir-Freundlich parameters fit for C2H2, CO2, and CH4 in Zn2(Pydc)(Ata)2 at 278 K.

Author Contributions

Conceptualization, Y.Z.; Methodology, N.X., Y.J. (Yunjia Jiang), W.S.; Validation, J.L., Y.J. (Yujie Jin); Formal Analysis, N.X., Y.J. (Yunjia Jiang), W.S.; Writing-Original Draft Preparation, N.X., Y.Z.; Writing-Review & Editing, Y.J. (Yunjia Jiang), L.W., Y.Z., S.D.; Supervision, Y.Z., D.W., S.D.; Funding Acquisition, L.W., Y.Z., S.D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 21908193, 21871231, 21850410451), the Special Funds for Basic Scientific Research of Zhejiang University.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available in the Supplementary Materials of this article or from the corresponding author upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Stang, P.J.; Diederich, F. Modern Acetylene Chemistry; WileyVCH: Weinheim, Germany, 2008. [Google Scholar]
  2. Alduhaish, O.; Wang, H.; Li, B.; Hu, T.-L.; Arman, H.D.; Alfooty, K.; Chen, B. A Twofold Interpenetrated Metal–Organic Framework with High Performance in Selective Separation of C2H2/CH4. ChemPlusChem 2016, 81, 770–774. [Google Scholar] [CrossRef]
  3. Zhang, L.; Jiang, K.; Li, Y.; Zhao, D.; Yang, Y.; Cui, Y.; Chen, B.; Qian, G. Microporous Metal−Organic Framework with Exposed Amino Functional Group for High Acetylene Storage and Excellent C2H2/CO2 and C2H2/CH4 Separations. Cryst. Growth Des. 2017, 17, 2319–2322. [Google Scholar] [CrossRef]
  4. Hou, X.-J.; He, P.; Li, H.; Wang, X. Understanding the Adsorption Mechanism of C2H2, CO2, and CH4 in Isostructural Metal−Organic Frameworks with Coordinatively Unsaturated Metal Sites. J. Phys. Chem. C 2013, 117, 2824–2834. [Google Scholar] [CrossRef]
  5. Sholl, D.S.; Lively, R.P. Seven chemical separations to change the world. Nature 2016, 532, 435–438. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Hu, J.; Krishna, R.; Wang, L.; Yang, L.; Cui, X.; Duttwyler, S.; Xing, H. Rational Design of Microporous MOFs with Anionic Boron Cluster Functionality and Cooperative Dihydrogen Binding Sites for Highly Selective Capture of Acetylene. Angew. Chem. Int. Ed. 2020, 59, 17664–17669. [Google Scholar] [CrossRef] [PubMed]
  7. Zhao, X.; Wang, Y.; Li, D.S.; Bu, X.; Feng, P. Metal–Organic Frameworks for Separation. Adv. Mater. 2018, 30, 1705189. [Google Scholar] [CrossRef] [PubMed]
  8. Bao, Z.; Chang, G.; Xing, H.; Krishna, R.; Ren, Q.; Chen, B. Potential of microporous metal–organic frameworks for separation of hydrocarbon mixtures. Energy Environ. Sci. 2016, 9, 3612–3641. [Google Scholar] [CrossRef]
  9. Liao, P.-Q.; Huang, N.-Y.; Zhang, W.-X.; Zhang, J.-P.; Chen, X.-M. Controlling guest conformation for efficient purification of butadiene. Science 2017, 356, 1193–1196. [Google Scholar] [CrossRef] [Green Version]
  10. Li, Y.-P.; Wang, Y.; Xue, Y.-Y.; Li, H.-P.; Zhai, Q.-G.; Li, S.-N.; Jiang, Y.-C.; Hu, M.-C.; Bu, X. Ultramicroporous Building Units as a Path to Bi-microporous Metal-Organic Frameworks with High Acetylene Storage and Separation Performance. Angew. Chem. Int. Ed. 2019, 58, 13590–13595. [Google Scholar] [CrossRef]
  11. Lin, S.; Zhou, P.; Xu, T.; Fan, L.; Wang, X.; Yue, L.; Jiang, Z.; Zhang, Y.; Zhang, Z.; He, Y. Modulation of Topological Structures and Adsorption Properties of Copper-Tricarboxylate Frameworks Enabled by the Effect of the Functional Group and Its Position. Inorg. Chem. 2021, 60, 8111–8122. [Google Scholar] [CrossRef]
  12. Qiao, J.; Liu, X.; Liu, X.; Zhang, L.; Liu, Y. Two urea-functionalized pcu metal-organic frameworks based on a pillared-layer strategy for gas adsorption and separation. Inorg. Chem. Front. 2020, 7, 3500–3508. [Google Scholar] [CrossRef]
  13. Hua, G.-F.; Xie, X.-J.; Lu, W.-G.; Li, D. Optimizing supramolecular interactions in metal-organic frameworks for C2 separation. Dalton Trans. 2020, 49, 15548–15559. [Google Scholar] [CrossRef]
  14. Wang, Y.; Yuan, S.; Hu, Z.; Kundu, T.; Zhang, J.; Peh, S.B.; Cheng, Y.; Dong, J.; Yuan, D.; Zhou, H.-C.; et al. Pore Size Reduction in Zirconium Metal−Organic Frameworks for Ethylene/Ethane Separation. ACS Sustain. Chem. Eng. 2019, 7, 7118–7126. [Google Scholar] [CrossRef]
  15. Peng, Y.-L.; Pham, T.; Li, P.; Wang, T.; Chen, Y.; Chen, K.-J.; Forrest, K.A.; Space, B.; Cheng, P.; Zaworotko, M.J.; et al. Robust Ultramicroporous Metal-Organic Frameworks with Benchmark Affinity for Acetylene. Angew. Chem. Int. Ed. 2018, 57, 10971–10975. [Google Scholar] [CrossRef]
  16. Jiang, Z.; Zou, Y.; Xu, T.; Fan, L.; Zhou, P.; He, Y. A Hydrostable Cage-Based MOF with Open Metal Sites and Lewis Basic Sites Immobilized in Pore Surface for Efficient Separation and Purification of Natural Gas and C2H2. Dalton Trans. 2020, 49, 3553–3561. [Google Scholar] [CrossRef]
  17. Xu, T.; Fan, L.; Jiang, Z.; Zhou, P.; Li, Z.; Lu, H.; He, Y. Immobilization of N-Oxide Functionality into Nbo-Type Mofs for Significantly Enhanced C2H2/CH4 and CO2/CH4 Separations. Dalton Trans. 2020, 49, 7174–7181. [Google Scholar] [CrossRef]
  18. Li, N.; Chang, Z.; Huang, H.; Feng, R.; He, W.-W.; Zhong, M.; Madden, D.G.; Zaworotko, M.-J.; Bu, X.-H. Specific K+ Binding Sites as CO2 Traps in a Porous MOF for Enhanced CO2 Selective Sorption. Small 2019, 15, 1900426. [Google Scholar] [CrossRef]
  19. Kumar, A.; Madden, D.G.; Lusi, M.; Chen, K.-J.; Daniels, E.A.; Curtin, T.; Perry, J.J., IV; Zaworotko, M.J. Direct Air Capture of CO2 by Physisorbent Materials. Angew. Chem. Int. Ed. 2015, 54, 14372–14377. [Google Scholar] [CrossRef]
  20. Song, X.; Zhang, M.; Duan, J.; Bai, J. Constructing and finely tuning the CO2 traps of stable and various-pore-containing MOFs towards highly selective CO2 capture. Chem. Commun. 2019, 55, 3477–3480. [Google Scholar] [CrossRef]
  21. Ma, H.-Y.; Zhang, Y.-Z.; Yan, H.; Zhang, W.-J.; Li, Y.-W.; Wang, S.-N.; Li, D.-C.; Dou, J.-M.; Li, J.-R. Two microporous CoII-MOFs with dual active sites for highly selective adsorption of CO2/CH4 and CO2/N2. Dalton Trans. 2019, 48, 13541–13545. [Google Scholar] [CrossRef]
  22. Chen, Y.; Wu, H.; Xiao, Q.; Lv, D.; Li, F.; Li, Z.; Xia, Q. Rapid room temperature conversion of hydroxy double salt to MOF-505 for CO2 capture. Cryst. Eng. Comm. 2019, 21, 165–171. [Google Scholar] [CrossRef]
  23. Zhang, Y.; Cui, X.; Xing, H. Recent advances in the capture and abatement of toxic gases and vapors by metal-organic frameworks. Mater. Chem. Front. 2021, 5, 5970–6013. [Google Scholar] [CrossRef]
  24. Moon, S.-Y.; Liu, Y.; Hupp, J.T.; Farha, O.K. Instantaneous Hydrolysis of Nerve-Agent Simulants with a Six-Connected Zirconium-Based Metal-Organic Framework. Angew. Chem. Int. Ed. 2015, 54, 6795–6799. [Google Scholar] [CrossRef] [PubMed]
  25. Sabyrov, K.; Jiang, J.; Yaghi, O.M.; Somorjai, G.A. Hydroisomerization of n-Hexane Using Acidified Metal−Organic Framework and Platinum Nanoparticles. J. Am. Chem. Soc. 2017, 139, 12382–12385. [Google Scholar] [CrossRef] [PubMed]
  26. Niu, Z.; Zhang, W.; Lan, P.C.; Aguila, B.; Ma, S. Promoting Frustrated Lewis Pairs for Heterogeneous Chemoselective Hydrogenation via the Tailored Pore Environment within Metal-Organic Frameworks. Angew. Chem. Int. Ed. 2019, 58, 7420–7424. [Google Scholar] [CrossRef]
  27. Yue, D.; Zhao, D.; Zhang, J.; Zhang, L.; Jiang, K.; Zhang, X.; Cui, Y.; Yang, Y.; Chen, B.; Qian, G. A luminescent cerium metal-organic framework for the turn-on sensing of ascorbic acid. Chem. Commun. 2017, 53, 11221–11224. [Google Scholar] [CrossRef]
  28. Han, X.; Gu, C.; Ding, Y.; Yu, J.; Li, K.; Zhao, D.; Chen, B. Stable Eu3+/Cu2+-Functionalized Supramolecular Zinc(II) Complexes as Fluorescent Probes for Turn-On and Ratiometric Detection of Hydrogen Sulfide. ACS Appl. Mater. Interfaces 2021, 13, 20371–20379. [Google Scholar] [CrossRef]
  29. Jiang, J.; Furukawa, H.; Zhang, Y.-B.; Yaghi, O.M. High Methane Storage Working Capacity in Metal−Organic Frameworks with Acrylate Links. J. Am. Chem. Soc. 2016, 138, 10244–10251. [Google Scholar] [CrossRef] [PubMed]
  30. Wang, H.Y.; Ye, W.Q.; Yang, Y.; Zhong, Y.; Hu, Y. Zn-ion hybrid supercapacitors: Achievements, challenges and future perspectives. Nano Energy 2021, 85, 105942. [Google Scholar] [CrossRef]
  31. Hanikel, N.; Prevot, M.S.; Yaghi, O.M. MOF water harvesters. Nat. Nanotechnol. 2020, 15, 348–355. [Google Scholar] [CrossRef] [PubMed]
  32. Pei, J.; Shao, K.; Wang, J.-X.; Wen, H.-M.; Yang, Y.; Cui, Y.; Krishna, R.; Li, B.; Qian, G. A Chemically Stable Hofmann-Type Metal−Organic Framework with Sandwich-Like Binding Sites for Benchmark Acetylene Capture. Adv. Mater. 2020, 32, 1908275. [Google Scholar] [CrossRef]
  33. Niu, Z.; Cui, X.; Pham, T.; Verma, G.; Lan, P.C.; Shan, C.; Xing, H.; Forrest, K.A.; Suepaul, S.; Space, B.; et al. A MOF-based Ultra-Strong Acetylene Nano-trap for Highly Efficient C2H2/CO2 Separation. Angew. Chem. Int. Ed. 2021, 60, 5283–5288. [Google Scholar] [CrossRef]
  34. Wang, L.; Jiang, T.; Duttwyler, S.; Zhang, Y. Supramolecular Cu(II)-dipyridyl frameworks featuring weakly coordinating dodecaborate dianions for selective gas separation. Cryst. Eng. Comm. 2021, 23, 282–291. [Google Scholar] [CrossRef]
  35. Wang, L.; Sun, W.; Duttwyler, S.; Zhang, Y. Efficient adsorption separation of methane from CO2 and C2–C3 hydrocarbons in a microporous closo-dodecaborate [B12H12]2− pillared metal-organic framework. J. Solid State Chem. 2021, 299, 122167. [Google Scholar] [CrossRef]
  36. Yang, L.; Cui, X.; Zhang, Y.; Wang, Q.; Zhang, Z.; Suo, X.; Xing, H. Anion Pillared Metal−Organic Framework Embedded with Molecular Rotors for Size-Selective Capture of CO2 from CH4 and N2. ACS Sustain. Chem. Eng. 2019, 7, 3138–3144. [Google Scholar] [CrossRef]
  37. Lin, R.-B.; Li, L.; Zhou, H.-L.; Wu, H.; He, C.; Li, S.; Krishna, R.; Li, J.; Zhou, W.; Chen, B. Molecular sieving of ethylene from ethane using a rigid metal-organic framework. Nat. Mater. 2018, 17, 1128–1133. [Google Scholar] [CrossRef] [PubMed]
  38. Lan, J.; Qu, Y.; Zhang, X.; Ma, H.; Xu, P.; Sun, J. A novel water-stable MOF Zn(Py)(Atz) as heterogeneous catalyst for chemical conversion of CO2 with various epoxides under mild conditions. J. CO2 Util. 2020, 35, 216–224. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Yang, L.; Wang, L.; Duttwyler, S.; Xing, H. A Microporous Metal-Organic Framework Supramolecularly Assembled from a Cu(II) Dodecaborate Cluster Complex for Selective Gas Separation. Angew. Chem. Int. Ed. 2019, 58, 8145–8150. [Google Scholar] [CrossRef]
  40. Xu, T.; He, M.; Fan, L.; Zhou, P.; Jiang, Z.; He, Y. Engineering ligand conformation by substituent manipulation towards diverse copper-tricarboxylate frameworks with tuned gas adsorption properties. Dalton Trans. 2021, 50, 638–646. [Google Scholar] [CrossRef]
  41. Zhang, Y.; Yang, L.; Wang, L.; Cui, X.; Xing, H. Pillar iodination in functional boron cage hybrid supramolecular frameworks for high performance separation of light hydrocarbons. J. Mater. Chem. A 2019, 7, 27560–27566. [Google Scholar] [CrossRef]
  42. Chang, G.; Li, B.; Wang, H.; Hu, T.; Bao, Z.; Chen, B. Control of interpenetration in a microporous metal–organic framework for significantly enhanced C2H2/CO2 separation at room temperature. Chem. Commun. 2016, 52, 3494–3496. [Google Scholar] [CrossRef]
  43. Li, L.; Yin, Q.; Li, H.-F.; Liu, T.-F.; Cao, R. Rational design of phosphonocarboxylate metal–organic frameworks for light hydrocarbon separations. Mater. Chem. Front. 2018, 2, 1436–1440. [Google Scholar] [CrossRef]
  44. Fan, L.; Lin, S.; Wang, X.; Yue, L.; Xu, T.; Jiang, Z.; He, Y. A Series of Metal−Organic Framework Isomers Based on Pyridinedicarboxylate Ligands: Diversified Selective Gas Adsorption and the Positional Effect of Methyl Functionality. Inorg Chem. 2021, 60, 2704–2715. [Google Scholar] [CrossRef]
  45. Das, M.C.; Xu, H.; Xiang, S.; Zhang, Z.; Arman, H.D.; Qian, G.; Chen, B. A New Approach to Construct a Doubly Interpenetrated Microporous Metal–Organic Framework of Primitive Cubic Net for Highly Selective Sorption of Small Hydrocarbon Molecules. Chem. Eur. J. 2011, 17, 7817–7822. [Google Scholar] [CrossRef]
  46. Ding, T.; Zhang, S.; Zhang, W.; Zhang, G.; Gao, Z.W. Highly selective C2H2 and CO2 capture and magnetic properties of robust Co-chain based metal–organic framework. Dalton Trans. 2019, 48, 7938–7945. [Google Scholar] [CrossRef]
  47. Lin, R.-G.; Lin, R.-B.; Chen, B. A microporous metal–organic framework for selective C2H2 and CO2 Separation. J. Solid State Chem. 2017, 252, 138–141. [Google Scholar] [CrossRef]
  48. Huang, P.; Chen, C.; Wu, M.; Jiang, F.; Hong, M. An indium-organic framework for efficient storage of light hydrocarbons and selective removal of organic dyes. Dalton Trans. 2019, 48, 5527–5533. [Google Scholar] [CrossRef]
  49. Zeng, H.; Xie, M.; Huang, Y.L.; Zhao, Y.; Xie, X.J.; Bai, J.P.; Wan, M.Y.; Krishna, R.; Lu, W.; Li, D. Induced Fit of C2H2 in a Flexible MOF Through Cooperative Action of Open Metal Sites. Angew. Chem. Int. Ed. 2019, 58, 8515–8519. [Google Scholar] [CrossRef]
  50. Liu, R.; Liu, Q.Y.; Krishna, R.; Wang, W.; He, C.T.; Wang, Y.L. Water-Stable Europium 1,3,6,8-Tetrakis(4-carboxylphenyl)pyrene Framework for Efficient C2H2/CO2 Separation. Inorg. Chem. 2019, 58, 5089–5095. [Google Scholar] [CrossRef]
  51. Liu, L.; Yao, Z.; Ye, Y.; Chen, L.; Lin, Q.; Yang, Y.; Zhang, Z.; Xiang, S. Robustness, Selective Gas Separation, and Nitrobenzene Sensing on Two Isomers of Cadmium Metal–Organic Frameworks Containing Various Metal–O–Metal Chains. Inorg. Chem. 2018, 57, 12961–12968. [Google Scholar] [CrossRef]
  52. Wang, L.; Sun, W.; Zhang, Y.; Xu, N.; Krishna, R.; Hu, J.; Jiang, Y.; He, Y.; Xing, H. Interpenetration symmetry control within ultramicroporous robust boron cluster hybrid MOFs for benchmark purification of acetylene from carbon dioxide. Angew. Chem. Int. Ed. 2021. [Google Scholar] [CrossRef]
  53. Li, Y.T.; Zhang, J.W.; Lv, H.J.; Hu, M.C.; Li, S.N.; Jiang, Y.C.; Zhai, Q.G. Tailoring the Pore Environment of a Robust Ga-MOF by Deformed [Ga3O(COO)6] Cluster for Boosting C2H2 Uptake and Separation. Inorg. Chem. 2020, 59, 10368–10373. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. (A) Synthetic route towards Zn2(Pydc)(Ata)2. (B) Pydc2− coordination mode. (C) Ata coordination mode. (D) Zn2+ coordination environment.
Scheme 1. (A) Synthetic route towards Zn2(Pydc)(Ata)2. (B) Pydc2− coordination mode. (C) Ata coordination mode. (D) Zn2+ coordination environment.
Molecules 26 05121 sch001
Figure 1. (A) Structure of 1 showing the 1D channel with two different dimensions. (B,C) Connolly isosurface mapping on 1.
Figure 1. (A) Structure of 1 showing the 1D channel with two different dimensions. (B,C) Connolly isosurface mapping on 1.
Molecules 26 05121 g001
Figure 2. (A) N2 adsorption/desorption isotherms in 1 at 77 K. (BD) C2H2, CO2, and CH4 adsorption/desorption isotherms in 1 at 278 K, 288 K, and 298 K.
Figure 2. (A) N2 adsorption/desorption isotherms in 1 at 77 K. (BD) C2H2, CO2, and CH4 adsorption/desorption isotherms in 1 at 278 K, 288 K, and 298 K.
Molecules 26 05121 g002
Figure 3. (AC) IAST selectivities for equimolar C2H2/CO2 and C2H2/CH4 mixture in 1 at 278, 288, and 298 K. (D) IAST selectivities for C2H2/CO2 and C2H2/CH4 mixtures in 1 under 100 kPa with different C2H2 ratio at 278, 288, and 298 K.
Figure 3. (AC) IAST selectivities for equimolar C2H2/CO2 and C2H2/CH4 mixture in 1 at 278, 288, and 298 K. (D) IAST selectivities for C2H2/CO2 and C2H2/CH4 mixtures in 1 under 100 kPa with different C2H2 ratio at 278, 288, and 298 K.
Molecules 26 05121 g003
Figure 4. (A) Qst for C2H2, CO2 and CH4 in 1. (B) Cycling adsorption-desorption experiments. Before every measurement, 1 was regenerated at 200 (1st), 150 (2nd), 100 (3rd) 50 (4th), 25 (5th) °C, respectively, for 30 min.
Figure 4. (A) Qst for C2H2, CO2 and CH4 in 1. (B) Cycling adsorption-desorption experiments. Before every measurement, 1 was regenerated at 200 (1st), 150 (2nd), 100 (3rd) 50 (4th), 25 (5th) °C, respectively, for 30 min.
Molecules 26 05121 g004
Figure 5. (A) Column breakthrough experiments of C2H2/CO2 separation in 1 with a total flowrate of 2 mL/min at 25 °C. (B) Regeneration of 1 using an Ar purge of at 100 °C with a flow rate of 3 mL/min.
Figure 5. (A) Column breakthrough experiments of C2H2/CO2 separation in 1 with a total flowrate of 2 mL/min at 25 °C. (B) Regeneration of 1 using an Ar purge of at 100 °C with a flow rate of 3 mL/min.
Molecules 26 05121 g005
Table 1. Comparison of the gas adsorption performance of 1 and other MOFs.
Table 1. Comparison of the gas adsorption performance of 1 and other MOFs.
Uptake (cm3 g−1)
(298 K, 100 kPa)
IAST Selectivity
(298K, 100 kPa)
Ref
C2H2CO2CH4C2H2/CO2C2H2/CH4
FJI-C343.6--11.6--14.6[43]
ZJNU-2793.779.326.0--16.6[44]
USTA-3658.6 a--13.2 a--13.8 a[45]
[Co3(L)(OH)2(H2O)]·2DMF·2H2O107.362.017.7--13[46]
QMOF-141.524.63.9--13.5[47]
FJI-H2192.4--7.10--16.3[48]
JNU-16351--3--[49]
JXNU-5a55.934.8--5--[50]
UTSA-6870.1 a39.6 a--4 a--[42]
FJU-36a52.2 a35.5 a10.5 a2.8 a17.7 a[51]
BSF-152.539.710.53.346.9[39]
BSF-241.529.75.45.1324[41]
BSF-381.847.313.416.3205[6]
ZNU-1 (BSF-9)76.338.1--56.6--[52]
SNNU-6391.143.710.33.312.9[53]
ZJNU-109104.660.014.33.821.6[40]
Zn2(Pydc)(Ata)252.635.323.63.915.9This work
a at 296 K.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xu, N.; Jiang, Y.; Sun, W.; Li, J.; Wang, L.; Jin, Y.; Zhang, Y.; Wang, D.; Duttwyler, S. Gram-Scale Synthesis of an Ultrastable Microporous Metal-Organic Framework for Efficient Adsorptive Separation of C2H2/CO2 and C2H2/CH4. Molecules 2021, 26, 5121. https://doi.org/10.3390/molecules26175121

AMA Style

Xu N, Jiang Y, Sun W, Li J, Wang L, Jin Y, Zhang Y, Wang D, Duttwyler S. Gram-Scale Synthesis of an Ultrastable Microporous Metal-Organic Framework for Efficient Adsorptive Separation of C2H2/CO2 and C2H2/CH4. Molecules. 2021; 26(17):5121. https://doi.org/10.3390/molecules26175121

Chicago/Turabian Style

Xu, Nuo, Yunjia Jiang, Wanqi Sun, Jiahao Li, Lingyao Wang, Yujie Jin, Yuanbin Zhang, Dongmei Wang, and Simon Duttwyler. 2021. "Gram-Scale Synthesis of an Ultrastable Microporous Metal-Organic Framework for Efficient Adsorptive Separation of C2H2/CO2 and C2H2/CH4" Molecules 26, no. 17: 5121. https://doi.org/10.3390/molecules26175121

Article Metrics

Back to TopTop