Next Article in Journal
Identification of a Potential Inhibitor (MCULE-8777613195-0-12) of New Delhi Metallo-β-Lactamase-1 (NDM-1) Using In Silico and In Vitro Approaches
Next Article in Special Issue
Recent Progress in Metal Oxide-Based Photocatalysts for CO2 Reduction to Solar Fuels: A Review
Previous Article in Journal
The Antiviral Effects of 2-Deoxy-D-glucose (2-DG), a Dual D-Glucose and D-Mannose Mimetic, against SARS-CoV-2 and Other Highly Pathogenic Viruses
Previous Article in Special Issue
Synthesis of [B,Al]-EWT-Type Zeolite and Its Catalytic Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Customizing Pore System in a Microporous Metal–Organic Framework for Efficient C2H2 Separation from CO2 and C2H4

School of Materials Science and Engineering, National Institute for Advanced Materials, Nankai University, Tianjin 300350, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 18 August 2022 / Revised: 4 September 2022 / Accepted: 8 September 2022 / Published: 12 September 2022

Abstract

:
Selective-adsorption separation is an energy-efficient technology for the capture of acetylene (C2H2) from carbon dioxide (CO2) and ethylene (C2H4). However, it remains a critical challenge to effectively recognize C2H2 among CO2 and C2H4, owing to their analogous molecule sizes and physical properties. Herein, we report a new microporous metal–organic framework (NUM-14) possessing a carefully tailored pore system containing moderate pore size and nitro-functionalized channel surface for efficient separation of C2H2 from CO2 and C2H4. The activated NUM-14 (namely NUM-14a) exhibits sufficient pore space to acquire excellent C2H2 loading capacity (4.44 mmol g−1) under ambient conditions. In addition, it possesses dense nitro groups, acting as hydrogen bond acceptors, to selectively identify C2H2 molecules rather than CO2 and C2H4. The breakthrough experiments demonstrate the good actual separation ability of NUM-14a for C2H2/CO2 and C2H2/C2H4 mixtures. Furthermore, Grand Canonical Monte Carlo simulations indicate that the pore surface of the NUM-14a has a stronger affinity to preferentially bind C2H2 over CO2 and C2H4 via stronger C-H···O hydrogen bond interactions. This article provides some insights into customizing pore systems with desirable pore sizes and modifying groups in terms of MOF materials toward the capture of C2H2 from CO2 and C2H4 to promote the development of more MOF materials with excellent properties for gas adsorption and separation.

1. Introduction

Accompanied by worldwide economic advances, energy demand and supply have appeared cumulatively prominent; the thirst for high-purity raw materials, conveniently feasible technological streams, and desired final products has turned into even more unprecedented impendency. Industrial chemical separations occupy a large proportion of the quotient of global energy consumption, which has achieved a spectacular 10–15%, amounting to half of the usage amount of industrial energy in the United States [1]. Among the multitudinous commodity chemicals involving industrial interest, acetylene (C2H2) and ethylene (C2H4) are the two kinds of the most critical feedstocks for the electronic industry and polymerization manufacturing. Mature C2H2 fabrication methods in the petrochemical industry mainly rely on thermal cracking of hydrocarbons or partial oxidation of natural gas, and then obtained C2H2 production can be employed to cut/weld metals or manufacture diversiform high-value chemicals, such as vinyl chloride, acetaldehyde, acrylic acid derivatives, and synthetic fiber/rubber [2]. However, carbon dioxide (CO2) as an unwished impurity is inevitably mingled in the obtained C2H2 production during the aforementioned preparation processes, which will ultimately injure the quality of the downstream chemical products [3]. Therefore, the separation of C2H2/CO2 mixtures became of great importance in the petrochemical industry. In addition, C2H4, as an indispensable building block in the chemical synthesis industry with global annual production about 201 Mt by 2020, is generally used to manufacture polyethylene, vinyl chloride, ethylene oxide, etc. [4]. C2H4 is usually obtained by the catalytic cracking of hydrocarbons and steam cracking of naphtha; nevertheless, trace amounts of C2H2 as contaminants often inevitably coexist in C2H4 production. These C2H2 impurities about 1000–5000 ppm not only would significantly affect the quality of the resulting polyethylene but also can further form solid metal acetylide to block the fluid stream-triggering explosion [5]. Under this background, capturing trace C2H2 from C2H2/C2H4 mixture to purify C2H4 also became an imperative task. The similar geometric characteristics, including molecular size and shape, and alike physical properties mainly referring to the boiling point between C2H2 and CO2 or C2H4 molecules, render the separations of C2H2/CO2 and C2H2/C2H4 mixtures as high-challenge scientific problems [6,7,8]. Researchers in the petrochemical industry have devoted massive efforts to pursuing solutions for overcoming these separation difficulties. The current commercial pathway principally depends on cryogenic distillation, which has a high energy penalty and expensive economic cost. Porous materials utilizing selective physisorption with high energy efficiency, low investment cost, and environmentally friendly peculiarity might provide a promising alternative to separate these troublesome mixtures.
Metal–organic frameworks (MOFs) with large surface area, high porosity, fascinating modularity, and abundant functionality, as new-style crystalline porous materials, are constructed with metal ions/clusters and organic linkers. Because of their unique characteristics, MOFs have received wide attention in numerous application fields, such as gas storage and separation, heterogeneous catalysis, proton conduction, and fluorescence detection [9,10,11,12]. The most remarkable advantage of MOF materials over other traditional porous materials (such as zeolite, silica gel, and porous carbon) lies in the tunability of the pore system, predominantly referring to pore size/shape and pore surface chemistry. Plentiful adsorption-based MOFs were explored to address C2H2 separations challenges, which have obtained certain achievements [13,14]. Although a large number of research results indicated that incorporating strong binding sites (such as Lewis basic sites and open metal sites) into large-pore MOFs can effectively promote high C2H2 uptake capacities, most of them exhibited low separation selectivity and even terrible actual dynamic breakthrough properties due to their large pore size [15,16]. On the other side, the introduction of fluorinated anions (such as SiF62−, NbOF52−, and TiF62−) as hydrogen-bonding acceptors (HBAs) into ultra-microporous pillared hybrid MOFs can preferentially bind C2H2 molecules to dramatically enhance the separation selectivity [17,18]. However, the majority of ultra-microporous MOFs with excellent separation performance exhibited relatively low C2H2 loading capacity owing to their limited pore space. Recently, through the structure–performance relationship screening of 62 reported top-performing C2H2 adsorbents, Zhai et al. found that tuning the aperture of MOFs within the range of moderate pore size (5.0–7.5 Å), combined with accessible HBAs, is an effective route to break through the trade-off barrier between C2H2 storage and C2H2/CO2 separation [19]. Many studies have shown that embedding functional groups (such as -F, -NO2, -NH2, -OH) into MOF channels can also improve the separation selectivity of C2H2/C2H4 without a slight sacrifice in intrinsic moderate pore volumes or surface areas to adsorb a large amount of the objective gas molecules [20,21]. Therefore, if a porous MOF contains a suitable pore system that is simultaneously provided with both moderate pore size and high-density accessible functional groups like HBAs, it may synchronously possess satisfactory separation performance for C2H2/CO2 and C2H2/C2H4 mixtures. Although multiple design strategies, such as the isoreticular chemistry principle [22], pore–space partition strategy [23], and forming structural interpenetration [24], have been developed to guide the synthesis of desired MOF materials, difficult challenges still remain in customizing MOF materials with desirable pore systems. This is because even subtle effects can lead to undesirable results in the uncontrollable and complex self-assembly process of the MOFs.
Herein, we designed and synthesized a novel Ni-MOF {[Ni(TPT)(NPC)(H2O)]·solvent}n (NUM-14, TPT = 2,4,6-tri(4-pyridinyl)-1,3,5-triazine, H2NPC = 3-nitrophthalic acid), featuring a desired pore size of 5.8 Å and functionalized pore environment decorated by abundant nitro groups serving as hydrogen bond receptors, which can preferentially capture C2H2 for achieving efficient separation of C2H2/CO2 and C2H2/C2H4 mixtures. The single-component adsorption isotherms indicate that NUM-14a (activated NUM-14) has a good adsorption capacity for C2H2 with 4.44 mmol g−1, exceeding some cutting-edge MOF materials. The predicted IAST selectivities and practical breakthrough experiments for C2H2/C2H4 and C2H2/CO2 mixtures demonstrated that NUM-14a has favourable separation potential for these two kinds of mixed gases. Moreover, the values of Qst calculation quantitatively vindicated that NUM-14a owns the strongest interaction with C2H2, stronger than with CO2 and C2H4, which is also proved from another side by Grand Canonical Monte Carlo (GCMC) calculations. Both experimental and theoretical results show that NUM-14a can realize the efficient separation of C2H2 from CO2 and C2H4, and that this fine performance mainly comes from its suitable pore system.

2. Results and Discussion

2.1. Single Crystal X-ray Diffraction Structure

The solvothermal reaction of TPT, H2NPC, and Ni(NO3)2 in a DMF/H2O mixed solvent system at 100 °C for 72 h harvested jade-green blocky crystals of as-synthesized NUM-14. The crystallographic structure analysis manifested that NUM-14 crystallizes in the trigonal P3121 space group. The asymmetric unit of NUM-14 contains one Ni2+ atom, half of deprotonated NPC2− ligand, and a half of TPT ligand. Each Ni2+ atom has a six-coordinate mode to form slightly distorted octahedral geometry with three N atoms from three different TPT ligands and two carboxylate O atoms from two different NPC2− linkers, as well as one terminal water molecule. Each NPC2− ligand connected two Ni2+ atoms to form a 1D infinite helix chain spiraling counterclockwise along the c axis. Each TPT ligand also coordinate with two Ni2+ atoms to generate two kinds of infinite 1D helix chains along the c axis, in which one of them rotates clockwise and another inversely rotates counterclockwise (Figure S1). Then three types of spiral chains make up one channel column. The remaining end of TPT ligands as pore wall on each channel column is connected to the adjacent channel column to form two kinds of one-dimensional channel structures along the c axis, one of which is the triangular channel with an aperture of about 6 Å while the another is similar to the former except that the nitro groups on the NPC2− linkers toward its interior (Figure 1b–d). The coordinate mononuclear Ni2+ octahedron and TPT ligand can be simplified as 5- and 3-connected nodes, respectively (Figure 1a). Therefore, NUM-14 can be simplified as a 2-nodal 3,5-connected topology network with the point symbol of (4·72)(43·62·74·8) (Figure 1b) [25]. Calculation by PLATON revealed that the solvent-accessible volume in fully desolvated NUM-14 is 59.4%. The accessible channel surface in NUM-14a is mainly composed of the pyridine/triazine rings and abundant nitro groups in ligands, getting a very polar and rich hydrogen bond receptor pore environment.

2.2. Purity and Stabilities of NUM-14

The experimental and activated PXRD patterns are strongly consistent with the simulated model based on the single-crystal data confirming the high-phase purity and skeleton stability after the degassing of NUM-14 (Figure S2). The thermostability was demonstrated by TGA. As shown in Figure S3, it can be seen that the skeleton of NUM-14 remains stable up to 300 °C. As shown in Figure S4, through testing PXRD of NUM-14 samples soaked in various common solvents for 1 week, we found that NUM-14 acquits itself well in solvent stability under different conditions.

2.3. Gas Adsorption Properties of NUM-14a

The activated sample of NUM-14a was prepared by heating the CH2Cl2-exchanged sample under preset high vacuum conditions at 40 °C for the duration of 10 h. Then, the N2 sorption isotherm was recorded at 77 K to characterize and evaluate the permanent porosity of NUM-14a. As indicated in Figure 2a, it shows a reversible type-I adsorption curve with the saturated loading of 289.9 cm3 g−1 and is akin to multitudinous typical microporous MOF materials [26,27]. According to the N2 adsorption isotherm, the BET (Brunauer–Emmett–Teller) and Langmuir surface areas were calculated, which reached 1075.5 and 1194.2 m2 g−1, respectively. The Horvath–Kawazoe method was applied to calculate the pore size distribution and the result exhibited the main aperture concentrates upon 5.8 Å, which allows small gas molecules to pass through easily.
Combining with appropriate pore size and functionalized polar surface, NUM-14a is promising for evaluating C2H2 adsorption and separation performance from CO2 and C2H4. As shown in Figure 2b,c, single-component sorption isotherms of NUM-14a for C2H2, CO2, and C2H4 were measured at 273 and 298 K. Under 1 bar, these pure-component gas adsorption isotherms revealed that the C2H2 adsorption capacities of NUM-14a are 143.1 and 99.5 cm3 g−1 at 273 and 298 K, respectively, which is markedly higher than that of CO2 (102 and 50.2 cm3 g−1) and C2H4 (115.8 and 82.7 cm3 g−1), implying the distinctly stronger adsorption ability of NUM-14a for C2H2 than CO2 and C2H4. Noteworthily, the adsorption capacity of C2H2 (4.44 mmol g−1) at 298 K and 1 bar in NUM-14a is superior to most top-performing MOFs, like Zn-FBA (1.03 mmol g−1) [28], IPM-101 (2.55 mmol g−1) [29], ZNU-1 (3.41 mmol g−1) [30], ZNU-4 (3.58 mmol g−1) [31], and ZJU-74a (3.83 mmol g−1) [32], which is attributed to its high porosity. Moreover, the C2H2/CO2 uptake ratio at 298 K and 1 bar for NUM-14a reaches 1.98, while the uptake ratio of C2H2/C2H4 is only 1.20. Inspired by the distinctive uptake capacity and preferential binding of C2H2 for NUM-14a, we then calculated the isosteric enthalpies of adsorption (Qst), a crucial metric that quantifies the interaction strength between gas molecules and pores in MOFs, by the Virial-type equation to quantitatively estimate the binding affinity between gas molecules and host framework (Figures S5–S7). The calculated Qst at near-zero coverage for C2H2 (31.57 kJ mol−1) is higher than CO2 (22.89 kJ mol−1) and C2H4 (23.67 kJ mol−1), demonstrating the relatively stronger host–guest affinity for NUM-14a toward C2H2 in contrast to CO2 and C2H4 (Figure 2d). These results of Qst are adequately consistent with the gas adsorption behaviors as depicted in single-component isotherms and prove the feasibility of the C2H2 preferential adsorption in the framework of NUM-14a.

2.4. Gas Separation Performances of NUM-14a

Given the preferential capture and stronger affinity of C2H2 than CO2 and C2H4 in NUM-14a, IAST was further adopted to evaluate the separation performances of NUM-14a for C2H2/CO2 and C2H2/C2H4 mixtures. Single-component adsorption isotherms of C2H2 and C2H4 obtained from experimental determinations were fitted by the dual-site Langmuir–Freundlich model, while the isotherms of CO2 were fitted by the single-site Langmuir–Freundlich equation for pursuing the more accurate consistency between experimental data and theoretical model (Figures S8–S13). The fitting results were then used to predictably calculate adsorptive selectivities for equimolar C2H2/CO2 and C2H2/C2H4 with two different ratios of 50/50 or 1/99 at 273 and 298 K (Figure 3a,b). As revealed in Figure 3b, the calculated selectivity of NUM-14a for equimolar C2H2/CO2 (50:50, v/v) is 3.37 at 298 K and 100 kPa, which is lower than some famous MOF materials under the same conditions, such as CuI@UiO-66-(COOH)2 (185.00) [33], ATC-Cu (53.60) [34], and MOF-OH (25.00) [35], and slightly lower than that of FJU-118a (7.80) [36], BUT-85 (6.10) [37], but higher than many other well-known materials, such as [Ca(dtztp)0.5] (1.70) [38], [Ni(tzba)0.5(F)(bpy)] (2.20) [39], CAU-10H (2.50) [40] and SNNU-5-Sc (2.66) [41]. Due to the similar C2H2 and C2H4 adsorption behavior, although NUM-14a exhibits relatively low selectivities for C2H2/C2H4 mixtures with volume ratios of 1:99 (1.63) and 50:50 (1.61) at 298 K and 100 kPa, the result for C2H2/C2H4 (1:99, v/v) is still comparable to many consequences of reported MOFs, such as 1.13 for NUM-12a [42], 1.61 for Zn(ad)(int) [4], 1.59 for ZJNU-14 [43], 1.77 for ZJNU-7 [44], and 2.1 for UiO-67-(NH2)2 (Table S2) [45].
To further evaluate the separation performance of NUM-14a for C2H2/CO2 and C2H2/C2H4 mixtures, we performed dynamic breakthrough experiments in a packed tube using NUM-14a as physical adsorbent at 298 K under a total inlet gas flow rate of 2 mL min−1. As shown in Figure 4a, the breakthrough curve clearly proves that NUM-14a can effectively separate the C2H2/CO2 mixture. When the equimolar C2H2/CO2 gas mixture passes through the adsorption column, the CO2 gas first elutes due to its deficient uptake capacity. Then, C2H2 breaks through the packed column with a penetration time of 19 min g−1 and the adsorbent slowly tends to saturate soon afterward. Ultimately, the outlet gas flow reaches adsorption equilibrium with the same components as the imported stream at 55 min g−1. Similarly, the separation performance of NUM-14a for the equimolar C2H2/C2H4 mixture is described in Figure 4b. The experiment result displays that C2H2 gas can be more preferably adsorbed in the separation unit than C2H4, so the pure C2H4 can be obtained with a time interval of 5 min g−1, which is shorter than the breakthrough time of the C2H2/CO2 mixture because of the stronger adsorption capacity of C2H4 relative to CO2. These breakthrough tests undoubtedly testify to the ability of NUM-14a for separating C2H4 from mixtures containing CO2 or C2H4.

2.5. Adsorption Mechanism

To profoundly investigate the interesting gas adsorption behaviors between the preferential sites and adsorbed gas molecules within NUM-14a, GCMC calculations were conducted to probe the interaction between the gas molecules and the host framework at 298 K and 1 bar. As shown in Figure 5a, for adsorbed C2H2 in pores, three strong binding sites of hydrogen bonds were found, where two distances of C-H···O between C2H2 and two nitro oxygen atoms are 2.378 Å and 3.608 Å, and the distance of C-H···O between terminal hydrogen atom of C2H2 and one carbonyl oxygen atom is 2.611 Å. The CO2 molecule only interacts with the framework of NUM-14a through one weak C-H···O (3.562 Å) hydrogen bond interaction between acidic hydrogen atom on the pyridine ring and the basic oxygen atom at the end of CO2 molecule (Figure 5b). For the C2H4 molecule which has weaker acidic than C2H2 and therefore has a weaker affinity to basic sites, two hydrogen bond interactions between C2H4 with two nitroxides were sought with long C-H···O distances (2.487 and 3.074 Å) (Figure 5c). By comparing GCMC calculation results, we can find that the pore surface of NUM-14a shows a stronger recognition effect on C2H2 molecules than CO2 and C2H4, explaining the reason for outstanding capture and separation performance of C2H2 in NUM-14a observed in experiments.
According to the adsorption data, we can find that NUM-14a has a good adsorption capacity for C2H2, which is mainly attributed to its medium pore size and large porosity. Moreover, the combination of separation experiments and the simulation calculations shows that NUM-14a has stronger binding forces for C2H2 compared with CO2 and C2H4, which is possibly ascribed to the stronger synergistic hydrogen bond interactions between C2H2 and nitro group and the exposed oxygen atom. All of these prove that the customized fabrication of a pore system with moderate pore size and modified surface with HBAs in MOFs is beneficial to separate C2H2/CO2 and C2H2/C2H4 mixtures.

3. Materials and Methods

3.1. Materials and Characterization

All chemicals and reagents were purchased from commercial suppliers and used without further purification. Powder X-ray diffraction (PXRD) was measured on a Rigaku Miniflex 600 with Cu Kα radiation (λ = 1.5425 Å) under air conditions. Thermogravimetric analysis (TGA) was recorded on a Rigaku standard thermogravimetry-differential thermal analysis (TG-DTA) analyzer, utilizing an empty and clean Al2O3 crucible as reference (heating rate = 10 °C min−1 in Ar atmosphere).

3.2. Gas Sorption Measurements

Before the sorption measurement, the sample of NUM-14 was soaked in dichloromethane for 3 days to exchange solvent molecules in the channels. The degas procedure for the solvent-exchanged NUM-14 was conducted at 40 °C under a high vacuum (less than 10−5 Torr) for 10 h and led to the formation of activated sample NUM-14a. The N2 sorption isotherm measurement was carried out using a Micrometrics ASAP 2460 volumetric gas adsorption analyzer at 77 K in a liquid nitrogen bath. The C2H2, C2H4, and CO2 sorption isotherm measurements were carried out at 273 and 298 K respectively using a Micrometrics ASAP 2020M volumetric gas adsorption analyzer.

3.3. X-ray Crystallography

Single-crystal X-ray diffraction data of NUM-14 were collected on the Rigaku XtaLAB PRO MM007 DW diffractometer with Cu Kα radiation (λ = 1.54184 Å) at T = 99.99 (1) K. The structure was solved with the SHELXT program and refined by full-matrix least-squares against F2 using the SHELXL program [46,47]. Anisotropic thermal parameters were implemented to all non-hydrogen atoms, and all hydrogen atoms were placed in the calculated positions and refined with isotropic thermal parameters. The Solvent Mask in Olex2 software was employed to remove scattering contributions of the disordered solvent molecules and the generated solvent-free data of direction intensities were further refined [48]. Details of the crystal parameters, data collection, and refinement of NUM-14 are listed in Table S1.

3.4. Synthesis of NUM-14

A solvothermal reaction of Ni(NO3)2·6H2O (29.0 mg, 0.1 mmol), TPT (18.7 mg, 0.06 mmol), and H2NPC (21.1 mg, 0.1 mmol) in 2.5 mL of a mixed solvent of DMF/H2O (4:1, v/v) was kept at 100 °C for 3 days, and blocky crystals of NUM-14 were first time synthesized in 83% yield based on the TPT ligand. When the reaction temperature slowly cooled to room temperature, a fresh sample was collected by filtration and washed with fresh DMF several times.

3.5. Isosteric Enthalpy of Adsorption Calculations

The experimental adsorption data of C2H2, CO2, and C2H4 at 273 and 298 K in NUM-14a were fitted using a virial model (Equation (1)):
ln P = ln N + 1 T i = 0 m a i N i + j = 0 n b j N j
where P is the pressure in Torr, N is the adsorbed amount in mmol g−1, T is the temperature in K, and ai and bj are virial coefficients.
The isosteric enthalpies of adsorption (Qst) were calculated based on the fitted virial coefficients using the following equation (Equation (2)):
Q s t = R i = 0 m a i N i
Qst is the coverage-dependent isosteric heat of adsorption and R is the universal gas constant.

3.6. Adsorption Selectivity Calculations

Ideal adsorbed solution theory (IAST) was utilized to predict gas adsorption selectivity of binary mixtures from the experimental single-component isotherms of C2H2, CO2, and C2H4 [49]. The experimental pure-component adsorption isotherms of C2H2 and C2H4 were initially fitted by the dual-site Langmuir–Freundlich model (Equation (3)), while CO2 was fitted by the single-site Langmuir–Freundlich model. That is, only the first half of Equation (3) is used to fit the experimental data to obtain a perfect fitting degree:
q = q A , s a t b A p 1 / n 1 1 + b A p 1 / n 1   + q B , s a t b A p 1 / n 2 1 + b B p 1 / n 2
where p is the pressure in kPa, q is the adsorbed amount in mmol g−1, and qA,sat and qB,sat are the saturation capacities of two distinct adsorption sites A and B in mmol g−1. bA and bB are the affinity coefficients in kPa−1, and n1 and n2 represent the deviations from an ideal homogeneous surface.
Then the fitted parameters were used to calculate the selectivity (Equation (4)):
S A / B = X A / X B Y A / Y B
In which, Xi and Yi represent the mole fractions of component i in the adsorbed and bulk phases, respectively.

3.7. Breakthrough Experiments

The breakthrough experiments for C2H2/CO2 (50:50, v/v) and C2H2/C2H4 (50:50, v/v) mixtures were completed on the Multi-component Adsorption Breakthrough Curve Analyzer from Beishide Instrument Technology Co., Ltd. (Beijing, China). An activated crystalline powder sample (1.30 g) was packed into a breakthrough column (6 mm diameter and 4 mL volume) with a 3 cm length of the sample loading, which was purged with He flow (50 mL min−1), sustaining 120 min at 40 °C before each breakthrough experiment. Subsequently, the mixed gas flows of C2H2/CO2 and C2H2/C2H4 with a flow rate of 2 mL min−1 were respectively introduced into the adsorber at 298 K. At the same time, the compositions of effluents from the packed column were monitored and analyzed in real-time by online mass spectrometry.

3.8. Grand Canonical Monte Carlo Simulations

The GCMC simulations were performed for the adsorption of C2H2, CO2, and C2H4 in NUM-14a with the Material Studio 8.0. The optimal adsorption sites were simulated under 298 K with a pressure of 1.0 bar. We used 1.0 × 107 cycles for equilibration, and the production steps were set to 1.0 × 106. The framework of NUM-14a and adsorbate molecules were treated as a rigid structure. A standard Lennard–Jones and Coulomb model was used and the Lennard–Jones parameters for the framework atoms as well as adsorbate molecules were adopted from the universal force field (UFF). Ewald summation was used to calculate electrostatic interactions for both adsorbent–adsorbate and adsorbateadsorbate interactions.

4. Conclusions

In summary, we constructed a new Ni-based MOF NUM-14, which has a fascinating pore system featuring an appropriate pore size and pleasant pore environment, to separate C2H2 well from CO2 and C2H4. By a feat of its pore characteristic, NUM-14a can adsorb more C2H2 than CO2 and C2H4. With the advantages of a nitro-functionalized pore wall framework, NUM-14a expresses greater affinity to C2H2 through hydrogen bond interactions. Furthermore, the strong pure-component gas adsorption behavior and favorable C2H2/CO2, and C2H2/C2H4 separation performance are achieved by NUM-14a. In addition, the underlying selective adsorption mechanism and separation reason are revealed by GCMC simulations at the molecular level. This work can guide the designed synthesis of desirable MOFs with a tailor-made pore system and practical application of them for high-challenge gas separation problems.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/molecules27185929/s1. Table S1. Crystal data and structure refinement parameters for NUM-14. Table S2. Comparisons of C2H2 uptake and selectivities of C2H2/CO2 and C2H2/C2H4 for NUM-14a and other MOFs. Figure S1. Three kinds of helix chains constitute the channel column of NUM-14. Figure S2. Comparison of simulated, experimental, and activated PXRD patterns of NUM-14. Figure S3. TGA curve for NUM-14 under Ar atmosphere. Figure S4. The PXRD patterns for NUM-14 after immersed in common solvents a week. Figures S5–S7. The details of Virial equation fitting to the experimental C2H2, CO2, and C2H4 adsorption data for NUM-14a. Figures S8–S13. The details of dual/single-site Langmuir–Freundlich isotherm fitting to the experimental C2H2, CO2, and C2H4 adsorption data for NUM-14a at 273 and 298 K. Figures S14–S16. Density distributions of C2H2, CO2, and C2H4 in NUM-14a at 298 K and 1 bar.

Author Contributions

Q.Z.: Conceptualization, Methodology, Visualization, Investigation, Writing-original draft. G.-N.H.: Investigation. X.L.: Software, Formal analysis. S.-Q.Y.: Investigation. T.-L.H.: Supervision, Conceptualization, Writing—review & editing, Project administration, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the Natural Science Foundation of Tianjin (20JCYBJC01330), and the National Natural Science Foundation of China (21673120) for financial support of this work.

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

There are no conflicts to declare.

Sample Availability

Samples of the compound NUM-14 are available from the authors.

References

  1. Sholl, D.S.; Lively, R.P. Seven chemical separations to change the world. Nature 2016, 532, 435–437. [Google Scholar] [CrossRef] [PubMed]
  2. Pässler, P.; Hefner, W.; Buckl, K.; Meinass, H.; Meiswinkel, A.; Wernicke, H.-J.; Ebersberg, G.; Müller, R.; Bässler, J.; Behringer, H.; et al. Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH: Weinheim, Germany, 2000. [Google Scholar]
  3. Yang, S.-Q.; Hu, T.-L. Reverse-selective metal-organic framework materials for the efficient separation and purification of light hydrocarbons. Coord. Chem. Rev. 2022, 468, 214628. [Google Scholar] [CrossRef]
  4. Ding, Q.; Zhang, Z.; Liu, Y.; Chai, K.; Krishna, R.; Zhang, S. One-step ethylene purification from ternary mixtures in a metal-organic framework with customized pore chemistry and shape. Angew. Chem. Int. Ed. 2022, 61, e202208134. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, S.-Q.; Sun, F.-Z.; Liu, P.; Li, L.; Krishna, R.; Zhang, Y.-H.; Li, Q.; Zhou, L.; Hu, T.-L. Efficient purification of ethylene from C2 hydrocarbons with an C2H6/C2H2-selective metal-organic framework. ACS Appl. Mater. Interfaces 2021, 13, 962–969. [Google Scholar] [CrossRef] [PubMed]
  6. Chen, Y.; Xiong, Q.; Wang, Y.; Du, Y.; Wang, Y.; Yang, J.; Li, L.; Li, J. Boosting molecular recognition of acetylene in UiO-66 framework through pore environment functionalization. Chem. Eng. Sci. 2021, 237, 116572. [Google Scholar] [CrossRef]
  7. Mu, X.-B.; Xue, Y.-Y.; Hu, M.-C.; Zhang, P.; Wang, Y.; Li, H.-P.; Li, S.-N.; Zhai, Q.-G. Fine-tuning of pore-space-partitioned metal-organic frameworks for efficient C2H2/C2H4 and C2H2/CO2 separation. Chin. Chem. Lett. 2022. [Google Scholar] [CrossRef]
  8. Zhang, Q.; Zhou, L.; Liu, P.; Li, L.; Yang, S.-Q.; Li, Z.-F.; Hu, T.-L. Integrating tri-mural nanotraps into a microporous metal-organic framework for C2H2/CO2 and C2H2/C2H4 separation. Sep. Purif. Technol. 2022, 296, 121404. [Google Scholar] [CrossRef]
  9. Sahoo, R.; Das, M.C. C2s/C1 hydrocarbon separation: The major step towards natural gas purification by metal-organic frameworks (MOFs). Coord. Chem. Rev. 2021, 442, 213998. [Google Scholar] [CrossRef]
  10. Shao, Y.-R.; Zhou, L.; Yu, L.; Li, Z.-F.; Li, Y.-T.; Li, W.; Hu, T.-L. In situ construction of a Co/ZnO@C heterojunction catalyst for efficient hydrogenation of biomass derivative under mild conditions. ACS Appl. Mater. Interfaces 2022, 14, 17195–17207. [Google Scholar] [CrossRef]
  11. Pal, S.C.; Das, M.C. Superprotonic conductivity of MOFs and other crystalline platforms beyond 10−1 S cm−1. Adv. Funct. Mater. 2021, 31, 2101584. [Google Scholar] [CrossRef]
  12. Xu, Y.; Yang, S.-L.; Li, G.; Bu, R.; Liu, X.-Y.; Gao, E.-Q. Cooperative protonation underlying the acid response of metal-organic frameworks. Chem. Mater. 2022, 34, 5500–5510. [Google Scholar] [CrossRef]
  13. Qazvini, O.T.; Babarao, R.; Telfer, S.G. Multipurpose metal-organic framework for the adsorption of acetylene: Ethylene purification and carbon dioxide removal. Chem. Mater. 2019, 31, 4919–4926. [Google Scholar] [CrossRef]
  14. Verma, G.; Ren, J.; Kumar, S.; Ma, S. New paradigms in porous framework materials for acetylene storage and separation. Eur. J. Inorg. Chem. 2021, 44, 4498–4507. [Google Scholar] [CrossRef]
  15. Luna-Triguero, A.; Vicent-Luna, J.M.; Madero-Castro, R.M.; Gómez-Álvarez, P.; Calero, S. Acetylene storage and separation using metal-organic frameworks with open metal sites. ACS Appl. Mater. Interfaces 2019, 11, 31499–31507. [Google Scholar] [CrossRef]
  16. Wen, H.-M.; Wang, H.; Li, B.; Cui, Y.; Wang, H.; Qian, G.; Chen, B. A microporous metal-organic framework with lewis basic nitrogen sites for high C2H2 storage and significantly enhanced C2H2/CO2 separation at ambient conditions. Inorg. Chem. 2016, 55, 7214–7218. [Google Scholar] [CrossRef]
  17. Yang, L.; Jin, A.; Ge, L.; Cui, X.; Xing, H. A novel interpenetrated anion-pillared porous material with high water tolerance afforded efficient C2H2/C2H4 separation. Chem. Commun. 2019, 55, 5001–5004. [Google Scholar] [CrossRef]
  18. Cui, X.; Chen, K.; Xing, H.; Yang, Q.; Krishna, R.; Bao, Z.; Wu, H.; Zhou, W.; Dong, X.; Han, Y.; et al. Pore chemistry and size control in hybrid porous materials for acetylene capture from ethylene. Science 2016, 353, 141–144. [Google Scholar] [CrossRef]
  19. Xue, Y.-Y.; Bai, X.-Y.; Zhang, J.; Wang, Y.; Li, S.-N.; Jiang, Y.-C.; Hu, M.-C.; Zhai, Q.-G. Precise pore space partitions combined with high-density hydrogen-bonding acceptors within metal-organic frameworks for highly efficient acetylene storage and separation. Angew. Chem. Int. Ed. 2021, 60, 10122–10128. [Google Scholar] [CrossRef]
  20. Chen, F.; Bai, D.; Wang, X.; He, Y. A comparative study of the effect of functional groups on C2H2 adsorption in NbO-type metal-organic frameworks. Inorg. Chem. Front. 2017, 4, 960–967. [Google Scholar] [CrossRef]
  21. Qazvini, O.T.; Macreadie, L.K.; Telfer, S.G. Effect of ligand functionalization on the separation of small hydrocarbons and CO2 by a series of MUF-15 analogues. Chem. Mater. 2020, 32, 6744–6752. [Google Scholar] [CrossRef]
  22. Fan, W.; Zhang, X.; Kang, Z.; Liu, X.; Sun, D. Isoreticular chemistry within metal-organic frameworks for gas storage and separation. Coord. Chem. Rev. 2021, 443, 213968. [Google Scholar] [CrossRef]
  23. Hong, A.N.; Yang, H.; Bu, X.; Feng, P. Pore space partition of metal-organic frameworks for gas storage and separation. EnergyChem 2022, 4, 100080. [Google Scholar] [CrossRef]
  24. Wang, T.; Lin, E.; Peng, Y.-L.; Chen, Y.; Cheng, P.; Zhang, Z. Rational design and synthesis of ultramicroporous metal-organic frameworks for gas separation. Coord. Chem. Rev. 2020, 423, 213485. [Google Scholar] [CrossRef]
  25. Zhang, D.-S.; Chang, Z.; Li, Y.-F.; Jiang, Z.-Y.; Xuan, Z.-H.; Zhang, Y.-H.; Li, J.-R.; Chen, Q.; Hu, T.-L.; Bu, X.-H. Fluorous metal-organic frameworks with enhanced stability and high H2/CO2 storage capacities. Sci. Rep. 2013, 3, 3312. [Google Scholar]
  26. Sun, F.-Z.; Yang, S.-Q.; Krishna, R.; Zhang, Y.-H.; Xia, Y.-P.; Hu, T.-L. Microporous metal-organic framework with a completely reversed adsorption relationship for C2 hydrocarbons at room temperature. ACS Appl. Mater. Interfaces 2020, 12, 6105–6111. [Google Scholar] [CrossRef] [PubMed]
  27. Jiang, S.; Guo, L.; Chen, L.; Song, C.; Liu, B.; Yang, Q.; Zhang, Z.; Yang, Y.; Ren, Q.; Bao, Z. A strongly hydrophobic ethane-selective metal-organic framework for efficient ethane/ethylene separation. Chem. Eng. J. 2022, 442, 136152. [Google Scholar] [CrossRef]
  28. Yang, L.; Yan, L.; Niu, W.; Feng, Y.; Fu, Q.; Zhang, S.; Zhang, Y.; Li, L.; Gu, X.; Dai, P.; et al. Adsorption in reversed order of C2 hydrocarbons on an ultramicroporous fluorinated metal-organic framework. Angew. Chem. Int. Ed. 2022, 61, e202204046. [Google Scholar] [CrossRef]
  29. Sharma, S.; Mukherjee, S.; Desai, A.V.; Vandichel, M.; Dam, G.K.; Jadhav, A.; Kociok-Köhn, G.; Zaworotko, M.J.; Ghosh, S.K. Efficient capture of trace acetylene by an ultramicroporous metal-organic framework with purine binding sites. Chem. Mater. 2021, 33, 5800–5808. [Google Scholar] [CrossRef]
  30. 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, 60, 22865–22870. [Google Scholar] [CrossRef] [PubMed]
  31. Xu, N.; Hu, J.; Wang, L.; Luo, D.; Sun, W.; Hu, Y.; Wang, D.; Cui, X.; Xing, H.; Zhang, Y. A TIFSIX pillared MOF with unprecedented zsd topology for efficient separation of acetylene from quaternary mixtures. Chem. Eng. J. 2022, 450, 138034. [Google Scholar] [CrossRef]
  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] [PubMed]
  33. Zhang, L.; Jiang, K.; Yang, L.; Li, L.; Hu, E.; Yang, L.; Shao, K.; Xing, H.; Cui, Y.; Yang, Y.; et al. Benchmark C2H2/CO2 separation in an ultra-microporous metal-organic framework via copper(I)-alkynyl chemistry. Angew. Chem. Int. Ed. 2021, 60, 15995–16002. [Google Scholar] [CrossRef]
  34. 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]
  35. Gong, W.; Cui, H.; Xie, Y.; Li, Y.; Tang, X.; Liu, Y.; Cui, Y.; Chen, B. Efficient C2H2/CO2 separation in ultramicroporous metal-organic frameworks with record C2H2 storage density. J. Am. Chem. Soc. 2021, 143, 14869–14876. [Google Scholar] [CrossRef]
  36. Song, Q.; Yang, Y.; Yuan, F.; Zhu, S.; Wang, J.; Xiang, S.; Zhang, Z. Electrostatic force-driven lattice water bridging to stabilize a partially charged indium MOF for efficient separation of C2H2/CO2 mixtures. J. Mater. Chem. A 2022, 10, 9363–9369. [Google Scholar] [CrossRef]
  37. Si, G.-R.; Wu, W.; He, T.; Xu, Z.-C.; Wang, K.; Li, J.-R. Stable bimetallic metal-organic framework with dual-functional pyrazolate-carboxylate ligand: Rational construction and C2H2/CO2 separation. ACS Mater. Lett. 2022, 4, 1032–1036. [Google Scholar] [CrossRef]
  38. Wang, G.-D.; Li, Y.-Z.; Zhang, W.-F.; Hou, L.; Wang, Y.-Y.; Zhu, Z. Acetylene separation by a Ca-MOF containing accessible sites of open metal centers and organic groups. ACS Appl. Mater. Interfaces 2021, 13, 58862–58870. [Google Scholar] [CrossRef]
  39. Wang, G.-D.; Wang, H.-H.; Shi, W.-J.; Hou, L.; Wang, Y.-Y.; Zhu, Z. A highly stable MOF with F and N accessible sites for efficient capture and separation of acetylene from ternary mixtures. J. Mater. Chem. A 2021, 9, 24495–24502. [Google Scholar] [CrossRef]
  40. Ye, Y.; Xian, S.; Cui, H.; Tan, K.; Gong, L.; Liang, B.; Pham, T.; Pandey, H.; Krishna, R.; Lan, P.C.; et al. Metal-organic framework based hydrogen-bonding nanotrap for efficient acetylene storage and separation. J. Am. Chem. Soc. 2022, 144, 1681–1689. [Google Scholar] [CrossRef]
  41. Lv, H.-J.; Zhang, J.-W.; Jiang, Y.-C.; Li, S.-N.; Hu, M.-C.; Zhai, Q.-G. Micropore regulation in ultrastable [Sc3O]-organic frameworks for acetylene storage and purification. Inorg. Chem. 2022, 61, 3553–3562. [Google Scholar] [CrossRef]
  42. Zhang, Q.; Yang, S.-Q.; Zhou, L.; Yu, L.; Li, Z.-F.; Zhai, Y.-J.; Hu, T.-L. Pore-space partition through an embedding metal-carboxylate chain-induced topology upgrade strategy for the separation of acetylene/ethylene. Inorg. Chem. 2021, 60, 19328–19335. [Google Scholar] [CrossRef]
  43. Jiang, Z.; Fan, L.; Zhou, P.; Xu, T.; Chen, J.; Hu, S.; Chen, D.-L.; He, Y. An N-oxide-functionalized nanocage-based copper-tricarboxylate framework for the selective capture of C2H2. Dalton Trans. 2020, 49, 15672–15681. [Google Scholar] [CrossRef]
  44. Jiang, Z.; Fan, L.; Zhou, P.; Xu, T.; Hu, S.; Chen, J.; Chen, D.-L.; He, Y. An aromatic-rich cage-based MOF with inorganic chloride ions decorating the pore surface displaying the preferential adsorption of C2H2 and C2H6 over C2H4. Inorg. Chem. Front. 2021, 8, 1243–1252. [Google Scholar] [CrossRef]
  45. Gu, X.-W.; Wang, J.-X.; Wu, E.; Wu, H.; Zhou, W.; Qian, G.; Chen, B.; Li, B. Immobilization of lewis basic sites into a stable ethane-selective MOF enabling one-step separation of ethylene from a ternary mixture. J. Am. Chem. Soc. 2022, 144, 2614–2623. [Google Scholar] [CrossRef]
  46. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. A 2015, 71, 3–8. [Google Scholar] [CrossRef]
  47. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef]
  48. Bourhis, L.J.; Dolomanov, O.V.; Gildea, R.J.; Howard, J.A.; Puschmann, H. The anatomy of a comprehensive constrained, restrained refinement program for the modern computing environment—Olex2 dissected. Acta Cryst. A 2015, 71, 59–75. [Google Scholar] [CrossRef]
  49. Myers, A.L.; Prausnitz, J.M. Thermodynamics of mixed-gas adsorption. AlChE J. 1965, 11, 121–127. [Google Scholar] [CrossRef]
Figure 1. (a) Topology simplification of NiN3O3 octahedron and the ligands. (b) The triangular channel structure and topology simplification of 3D framework in NUM-14 along the c axis. (c) The Connolly surface void spaces of NUM-14. (d) The side view of channel structure in NUM-14 along the a-axis. Color code: Ni, sky blue; O, red; N, blue; C, light orange. Guest molecules and H atoms have been omitted for clarity.
Figure 1. (a) Topology simplification of NiN3O3 octahedron and the ligands. (b) The triangular channel structure and topology simplification of 3D framework in NUM-14 along the c axis. (c) The Connolly surface void spaces of NUM-14. (d) The side view of channel structure in NUM-14 along the a-axis. Color code: Ni, sky blue; O, red; N, blue; C, light orange. Guest molecules and H atoms have been omitted for clarity.
Molecules 27 05929 g001
Figure 2. (a) N2 sorption isotherm and pore size distribution of NUM-14a at 77 K. Single-component gas isotherms of C2H2, CO2, and C2H4 for NUM-14a at (b) 273 K and (c) 298 K (filled and open symbols represent adsorption and desorption curves, respectively). (d) Isosteric enthalpy of adsorption of C2H2, CO2, and C2H in NUM-14a.
Figure 2. (a) N2 sorption isotherm and pore size distribution of NUM-14a at 77 K. Single-component gas isotherms of C2H2, CO2, and C2H4 for NUM-14a at (b) 273 K and (c) 298 K (filled and open symbols represent adsorption and desorption curves, respectively). (d) Isosteric enthalpy of adsorption of C2H2, CO2, and C2H in NUM-14a.
Molecules 27 05929 g002
Figure 3. (a) The adsorption selectivities of NUM-14a, predicted from IAST for C2H2/CO2 (50:50, v/v), C2H2/C2H4 (1:99, v/v) and C2H2/C2H4 (50:50, v/v) at 273 K (a) and 298 K (b).
Figure 3. (a) The adsorption selectivities of NUM-14a, predicted from IAST for C2H2/CO2 (50:50, v/v), C2H2/C2H4 (1:99, v/v) and C2H2/C2H4 (50:50, v/v) at 273 K (a) and 298 K (b).
Molecules 27 05929 g003
Figure 4. The column breakthrough curves for mixture gases of (a) C2H2/CO2 (50:50, v/v), and (b) C2H2/C2H4 (50:50, v/v). The experiments were conducted at 298 K and the inlet gas flow rate was maintained at 2 mL min−1.
Figure 4. The column breakthrough curves for mixture gases of (a) C2H2/CO2 (50:50, v/v), and (b) C2H2/C2H4 (50:50, v/v). The experiments were conducted at 298 K and the inlet gas flow rate was maintained at 2 mL min−1.
Molecules 27 05929 g004
Figure 5. The GCMC calculated adsorption sites of C2H2 (a), CO2 (b), and C2H4 (c) in NUM-14a at 298 K and 1 bar.
Figure 5. The GCMC calculated adsorption sites of C2H2 (a), CO2 (b), and C2H4 (c) in NUM-14a at 298 K and 1 bar.
Molecules 27 05929 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, Q.; Han, G.-N.; Lian, X.; Yang, S.-Q.; Hu, T.-L. Customizing Pore System in a Microporous Metal–Organic Framework for Efficient C2H2 Separation from CO2 and C2H4. Molecules 2022, 27, 5929. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27185929

AMA Style

Zhang Q, Han G-N, Lian X, Yang S-Q, Hu T-L. Customizing Pore System in a Microporous Metal–Organic Framework for Efficient C2H2 Separation from CO2 and C2H4. Molecules. 2022; 27(18):5929. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27185929

Chicago/Turabian Style

Zhang, Qiang, Guan-Nan Han, Xin Lian, Shan-Qing Yang, and Tong-Liang Hu. 2022. "Customizing Pore System in a Microporous Metal–Organic Framework for Efficient C2H2 Separation from CO2 and C2H4" Molecules 27, no. 18: 5929. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules27185929

Article Metrics

Back to TopTop