Kinetic separation of C2H6/C2H4 in a cage-interconnected metal–organic framework: an interaction-screening mechanism

Mo Xie , Zhou Lu , Weigang Lu * and Dan Li *
College of Chemistry and Materials Science, Guangdong Provincial Key Laboratory of Functional Supramolecular Coordination Materials and Applications, Jinan University, Guangzhou, Guangdong 510632, People's Republic of China. E-mail: weiganglu@jnu.edu.cn; danli@jnu.edu.cn

Received 2nd March 2022 , Accepted 12th April 2022

First published on 12th April 2022


Abstract

Kinetic-based adsorptive separation is deemed as an energy-efficient approach for gas purification, yet its underlying mechanism is difficult to justify. Herein, we propose an intriguing interaction-screening mechanism with a cage-interconnected metal–organic framework (JNU-2) as a model via a multi-scale theoretical approach. Grand Canonical Monte Carlo (GCMC) simulations establish gas diffusion channels with the calculated C2H4 and C2H6 adsorptions comparable to the experimental ones. Molecular dynamic (MD) simulations reveal single-molecule passages along the diffusion channel and that the probability of C2H6 diffusing into the passage is nine times higher than that of C2H4. Density functional theory (DFT) calculations further confirm an overall preferential interaction with C2H6 passing through the single-molecule passage. This work has successfully demonstrated a theoretical methodology of multi-scale simulations and depicted a rarely observed interaction-screening mechanism in JNU-2 that corroborates its balanced adsorption capacity and C2H6/C2H4 adsorption selectivity. Such a methodology should be applicable to other well-defined structures for a better understanding of their gas adsorption/separation behaviours.


Introduction

Metal–organic frameworks (MOFs), a class of advanced nanoporous materials, show great application potential in the field of gas adsorption and separation due to their tunable pore sizes, large accessible surface areas, and chemical modifiability.1–3 MOFs are constructed through metal clusters/metal ions and organic linkers, which facilitate the introduction of desired structural elements targeting specific guest molecules.4 Taking light hydrocarbon separation as an example, suitable surface functionalization and matching pore size could selectively amplify the host–guest interaction, leading to excellent separation efficiency.5–9

Clarifying gas adsorption behaviours in the existing MOFs would be of great benefit for the further design and applications of new MOFs, such as gas storage and purification.10 Studies on thermodynamics-dominant gas separation of MOF materials have been developed maturely by combining the experimental data and theoretical calculations. MOF-74 is one of the representatives with open metal sites (OMSs) to show excellent gas adsorption and separation behaviours by thermodynamic interactions, in which the OMSs provide strong binding sites to unsaturated C–C bonds and thus produce higher selectivity of olefins/alkynes over alkanes.11,12 Kinetic sieving is another efficient strategy that has been widely applied owing to its excellent separation capability and easy desorption.13,14 The classic size-exclusion mechanism of kinetic sieving leads to selective adsorption of the small-sized one but fails to explain the reversed selectivity. In contrast to enormous experimental results and evidence, theoretical simulation has witnessed development lag likely due to the difficulties in determining the precise locations of loaded gas molecules and predicting the gas diffusion process. Meanwhile, traditional static models with loaded gas molecules, some from X-ray diffraction determination, are not conducive to revealing of a kinetic-based adsorption and selectivity mechanism.15,16 In this regard, molecular dynamic (MD) simulation in combination with Grand Canonical Monte Carlo (GCMC) is a powerful tool to probe the dynamic behaviours of gas molecules in MOFs.17–19 A comprehensive multi-scale simulation would be essential for painting a full picture of the gas adsorption behaviour, locally and globally, kinetically and thermodynamically.

Recently, our research group reported a microporous MOF (JNU-2) featuring large adsorption capacity and high C2H6/C2H4 selectivity.20 Single-component equilibrium adsorption and binding enthalpies for C2H6 and C2H4 indicate the similarity between their absorption behaviours, but the mixed-gas breakthrough experiments reveal an excellent C2H6/C2H4 separation (Fig. S1 and S2). It was suggested that kinetics might play a crucial part in this, while this C2H6-favoured adsorption did not conform to the classic molecular sieving mechanism. Herein, we carried out a multi-scale simulation study on JNU-2 to clarify the underlying kinetic mechanism for C2H6/C2H4 separation. GCMC simulations suggest that the largest cage (Cage C) is not accessible to C2H6 and C2H4 due to the small size of the opening windows. MD simulations further confirm that channel I connecting the two smaller cages (Cage A and Cage B) is the only gas diffusion pathway and Cage A is a single-molecule passage, while Cage B functions as a gas adsorption and storage chamber. DFT calculations demonstrate a negligible thermodynamic effect of Cage B on C2H6 and C2H4 but an overall more favourable interaction energy pathway for C2H6 diffusion through Cage A. The above multi-scale simulations and calculations enable us to establish an interaction-screening-based kinetic separation mechanism in JNU-2 for C2H6/C2H4 separation. This work demonstrates a generalizable theoretical methodology for the in-depth understanding of gas adsorptions and separations in MOF-based materials.

Results and discussion

The desired material JNU-2 was reproduced by previous reports; adsorption enthalpy (Qst) of C2H6 and C2H4 in JNU-2 is provided in Fig. S1 and breakthrough curves for the C2H6/C2H4 (10/90) mixture through JNU-2 are shown in Fig. S2. For discussion convenience, the three cage-like cavities in the crystal structure of JNU-2, from small to large, are referred to as Cage A, B, and C (Fig. 1a). Each two of them are linearly interconnected into one-dimensional channels in the directions perpendicular to the (100), (101), and (111) crystal planes, labelled as Channel I, II, and III, respectively (Fig. 1b).
image file: d2qi00465h-f1.tif
Fig. 1 (a) An orthogonal stacking of the three cages in the crystal structure of JNU-2, where Cage A, B, and C are highlighted in coloured spheres. H atoms are omitted for clarity. Colour representation: gold, Cu; dark blue, Zn; blue, N; grey, C; red, O. (b) Three possible gas diffusion channels in JNU-2. (c) GCMC simulated adsorption isotherms of C2H6 and C2H4 in reference to the experimental data. (d) The corrected GCMC simulated adsorption isotherms of C2H6 and C2H4 in reference to the experimental data.

GCMC simulation

GCMC simulation is a powerful tool to study gas adsorption in porous materials including MOFs. It can provide not only adsorption isotherms but also gas distribution statistics inside the frameworks accordingly, allowing us to locate the strong adsorption sites. In this manner, we carried out GCMC simulations of the adsorption of C2H6 and C2H4 on JNU-2, respectively. The adsorption isotherms (Fig. 1c) show that the simulated C2H6 adsorption is higher than that of C2H4 at all pressure points, which is consistent with the experimental data. However, a huge discrepancy is observed between the simulation and experiment for both C2H6 and C2H4, suggesting that the simulations do not fully reflect their real adsorption situations.

To figure out the reason for the simulation/experiment discrepancy, the simulated gas distribution of C2H6 and C2H4 inside JNU-2 was analyzed. The adsorptions of C2H6 and C2H4 appear quite similar based on their distribution density maps; both are evenly distributed in all three cavities (Cage A, B, and C) (Fig. S3), suggesting that the discrepancy is not caused by an overestimation of adsorption sites. We subsequently calculated the adsorption isotherms by replacing the UFF/TraPPE force field with the Dreiding21/OPLS-AA22 force field and using the Gasteiger charge in simulations. As shown in Fig. S4, the Dreiding/OPLS-AA simulation results are almost the same as before. Thus, we can rule out that the simulation parameters are the cause of the simulation/experiment discrepancy. JNU-2 is a rigid framework as demonstrated in our experimental report,20 in which the flexibility should not be the reason for the discrepancy either. A further look into the GCMC simulations showed that the insertion, translation, and rotation of the adsorbates inside the framework were allowed, but ignoring their kinetic behaviors could overestimate the adsorption in some cavities with small apertures in the above GCMC simulations. Therefore, we speculated that C2H6 and C2H4 molecules may not be able to enter some of the cavities in JNU-2, which was overlooked in adsorption simulations. The gas accessibility to the three cages was examined through the volume and limiting diameter analysis. The probe radius used in the calculations was set to be 2.8 Å, leading to non-smooth spherical accessible dimensions with slight overflows, which were further estimated to be 133 Å3, 1873 Å3, and 3134 Å3 for Cage A, B, and C, respectively (Fig. 2). Every two of them are interconnected through a window (aperture), resulting in a total of three different kinds of windows in JNU-2. A cross-section can be obtained if we cut a plane at the narrowest part of the window, and the size of the thus-obtained cross-section should be the limiting size of the window. By performing this, we can see that Window 1 has a round cross-section with a diameter of 4.37 Å, Window 2 has a triangular cross-section with a maximum distance of 2.55 Å, and Window 3 has an elliptical cross-section with maximum and minimum diameters of 3.66 Å and 2.26 Å, respectively. Considering the kinetic diameters of C2H6 and C2H4 molecules, it seems that Window 1 serves as the only passable aperture for both gas molecules. The other two windows, both connecting to Cage C, are too small to allow either of the gas molecules to pass through, indicating that Cage C should be inaccessible to C2H6 and C2H4.


image file: d2qi00465h-f2.tif
Fig. 2 The calculated volume (yellow sphere) and window size of the three cages along with the size of gas molecules (orange value). Window 1 connects Cage A and Cage B, Window 2 connects Cage B and Cage C, and Window 3 connects Cage A and Cage C.

To correct the GCMC simulations, we performed the GCMC simulations using another commonly-used software RASPA 2.023 by employing the same force field and atomic charge calculation method (UFF/TraPPE, Qeq charge). To block the adsorption in Cage C, a solid sphere was placed at the center of Cage C and its radius was set to 8 nm; the schematic diagram of the blocking sphere in JNU-2 is shown in Fig. S5. The RASPA-simulated adsorption (Fig. 1d) after correction is in good agreement with the experimental data. The slightly higher uptake for both C2H6 and C2H4 could be attributed to the irregularity of Cage C and it is not well represented by the blocking sphere in RASPA calculations. Overall, the corrected adsorption is consistent with the experimental data, which strongly supports our assumption that Cage C is inaccessible to C2H6 and C2H4.

MD simulation

To study the dynamic adsorption behavior of C2H6/C2H4 inside JNU-2, MD simulations were conducted. Owing to the symmetry and rigidity of JNU-2, the directions of three possible gas diffusion pathways (Channel I, Channel II, and Channel III) are perpendicular to the (100), (110), and (111) crystal planes of JNU-2. We accordingly set the (100), (110), or (111) crystal face from the unit cell of JNU-2 as the interface with the gas phase, so that the gas molecules can move along the directions of three channels. As shown in Fig. 3, all gas molecules diffused into the cages of Channel I after 20 ns, and the rigidity of the framework was well maintained. However, only a negligible amount of gas molecules diffused through the first cage of Channel II or Channel III, further verifying our assumption that Cage C was not accessible to either C2H4 or C2H6. The snapshots of gas distribution and concentration profiles of C2H4 and C2H6 in the z-axis direction and x/y-axis direction within the 20 ns simulations were further tracked and are shown in Fig. 3 and Fig. S6, respectively. Interestingly, the concentrations of both gas molecules in Cage A are lower than 2 from the yz view (Fig. 3a), and 1.2 from the xz and xy view (Fig. S6). Subtracting those being close to interconnected open pores, the number of gas molecules inside Cage A should be less than or equal to 1, suggesting that Cage A can only accommodate one gas molecule. It is worth noting that there are partially opened cavities on the interface. Without restricting the direction of gas diffusion, only a few gas molecules were observed passing through the interface and entering into the framework layer. This further confirms that Channel I is the only gas diffusion pathway for C2H6 and C2H4. From the concentration profiles in Cage B and Cage A along Channel I, it can be suspected that Cage B provides space for high adsorption capacity, while Cage A is a diffusion-limiting single-molecule passage, which may be the site to promote the kinetic selectivity of C2H6 over C2H4.
image file: d2qi00465h-f3.tif
Fig. 3 MD simulation of the C2H4/C2H6 (1[thin space (1/6-em)]:[thin space (1/6-em)]1) mixture passing through (a) Channel I, (b) Channel II, and (c) Channel III of JNU-2 along the z-axis. Top: snapshots of the gas distribution at the initial (0 ns) and final (20 ns) stages. Bottom: gas concentration profiles of C2H6 and C2H4 along the z coordinate. H atoms in the models are omitted and the C2H6 and C2H4 molecules are highlighted in lake blue and rose red for clarity.

Given the essential role of Cage A in selective adsorption, the following MD simulations were performed to probe the free diffusion of gas molecules into Cage A. Ten parallel 500 ps MD simulations were carried out by modeling a discrete Cage A in the middle of a box filled with a C2H6/C2H4 (1[thin space (1/6-em)]:[thin space (1/6-em)]1) mixture that was randomly generated and annealed. As shown in Fig. 4, in 9 out of 10 simulations, it was the C2H6 molecule that diffused into Cage A and remained in it until the end of the 500 ps simulation. In the only simulation where the C2H4 molecule diffused into Cage A, it was, later on, exchanged out by the C2H6 molecule at 380 ps. The results reveal that Cage A is indeed a single-molecule passage that can preferentially take in the C2H6 molecule, which is consistent with our speculation that Cage A is the origin of the kinetic selectivity.


image file: d2qi00465h-f4.tif
Fig. 4 Snapshots of gas molecules entering into Cage A in 10 parallel 500 ps MD simulations in a box (30 Å × 30 Å × 30 Å) filled with a C2H6/C2H4 (1[thin space (1/6-em)]:[thin space (1/6-em)]1) mixture. The C2H6 and C2H4 molecules entering Cage A are coloured in lake blue and rose red respectively. The corresponding time bar represents the gas residence time in Cage A within 500 ps.

DFT calculations

To quantify the host–guest interaction between adsorbate molecules and JNU-2 and further elucidate the C2H6/C2H4 selective adsorption mechanism, we first performed DFT calculations on discrete Cage B with random addition of C2H6 or C2H4 molecules one by one. Five C2H6 or C2H4 molecules were introduced successfully, and their optimized conformations inside Cage B (5C2H6@Cage B and 5C2H4@Cage B) are shown in Fig. 5. For the five C2H6 molecules in Cage B, two are at Window 1 (site iv and v), two are close to Window 2 (site ii and iii), and one is hovering over a carboxyl group (site i). All five C2H6 molecules interact with Cage B by forming multiple C–H⋯O interactions. In the case of 5C2H4@Cage B, four C2H4 molecules are located nearby the Zn metal, forming weak metal-π-complexation with an interaction distance of about 3.3 Å, and the last one lies above Window 2 (site ii). The interaction mode and adsorption sites for C2H6 molecules inside Cage B are rather different from C2H4 molecules, indicating that these two gas molecules do not necessarily compete for adsorption sites inside Cage B. The total interaction energy was calculated to be −30.97 kcal mol−1 for C2H6 and −32.11 kcal mol−1 for C2H4, which can be broken down into electrostatic interaction, Pauli repulsion, orbital interaction, and dispersion items based on energy decomposition analysis (EDA) (Table S1). Although the electrostatic energy of 5C2H4@Cage B appears to be higher, which can be attributed to the interaction between the π-electrons of C2H4 and metal cations, the overall adsorptions of C2H4 and C2H6 in Cage B are not much different in terms of thermodynamics. The results suggest that Cage B is the chamber for large adsorption, and may not be accountable for the selectivity of C2H6 over C2H4.
image file: d2qi00465h-f5.tif
Fig. 5 The interaction configurations of (a) 5C2H6@Cage B and (b) 5C2H4@Cage B; all adsorption sites are highlighted and distances are in Å.

To simulate the dynamic behavior of C2H6 and C2H4 in Cage A, we built a model of complete Cage A with parts of connecting Cage B and Cage C. Relaxed scanning was performed to produce a potential energy curve (PEV) of adsorbate molecules entering (from Cage B to Cage A) and exiting (from Cage A to Cage B) Cage A (see Fig. 6a). The corresponding interaction energies between adsorbate molecules and JNU-2 and PEV at each configuration point were calculated and depicted in Fig. 6b and S10.


image file: d2qi00465h-f6.tif
Fig. 6 (a) The model of Cage A with partial Cage B on both ends for DFT calculation and the definition of Dcc. (b) The interaction energy profile of gas molecules moving along the axis of Dcc. Dcc is displayed as negative when gas molecules are moving away from the center of Cage A. (c) The representative interaction configurations of C2H6@Cage A (top) and C2H4@Cage A (bottom); hydrogen bonds are labelled with green dashed lines.

The whole process of the C2H6/C2H4 molecule passing through Cage A can be divided into five stages: Cage B edge, Window 1, Cage A, Window 1, and Cage B edge. The calculated interaction energies vary from −5.0 to −8.0 kcal mol−1, indicating an energetically favourable pathway for both C2H6 and C2H4 (Fig. 6b). In detail, the overall interaction energy shows a trough-like curve with an obvious barrier at Window 1, which can be attributed to steric hindrance and electrostatic repulsion at the narrowest part of Window 1. The representative interaction configurations (Fig. 6c) show that the C2H6 molecule enters and exits Cage A smoothly with some molecular rotation along the movement path. Except on the Cage B edge, the C2H6 molecule maintains at least four weak hydrogen bonds with Cage A throughout the entire entering and exiting process. At the center of Cage A (Dcc = −0.54 Å, Fig. 6c), C2H6 is in a diagonal-like configuration and interacts with heteroatoms on both Window 1, resulting in five hydrogen bonds in total. As for C2H4, the energy barrier at Window1 is relatively larger, indicating that it is less favourable for C2H4 to enter into Cage A, which is consistent with the MD simulations. In addition, the C2H4 molecule does not pass through Cage A in a parallel configuration (Fig. 6c and Fig. S12). A flip was observed for the C2H4 molecule to turn to the side of Cage A and maintain four hydrogen-bonding interactions. Even so, the interaction between C2H4 and Cage A is weaker than that between C2H6 and Cage A at its center.

Overall, Cage A is a single-molecule passage for gas molecules to enter into Cage B, which is a gas storage chamber accounting for the large adsorption of C2H6 and C2H4. C2H6 can maintain a steady configuration and strong interaction throughout the whole diffusion process, especially at the center of Cage A. However, C2H4 has to adjust its molecular configuration to maintain strong interaction with Cage A, resulting in a less favourable interaction pathway for C2H4 entering and exiting Cage A. It should be pointed out that the gas molecule has to enter Cage A to reach Cage B; considering the fact that the possibility of C2H6 entering into Cage A is 9 times higher than that of C2H4, a kinetic selectivity of C2H6 over C2H4 can thus be rationalized by the multi-scale simulation study.

Conclusions

In summary, we carried out a multi-scale simulation study on JNU-2 to explore its adsorption and separation behaviour of C2H6 and C2H4. The results justify a C2H6-favoured kinetic separation mechanism that has rarely been observed in adsorption separation in porous materials. The adsorption isotherms obtained from GCMC simulations on JNU-2 with Cage C blocked are consistent with the experimental ones, suggesting that these gas molecules are only adsorbed in the two smaller cages (Cage A and Cage B). MD simulations confirm that the only gas diffusion channel is the one (Channel I) that connects Cage A and Cage B, and the probability of C2H6 diffusing into Cage A is 9 times higher than that of C2H4. DFT calculations further clarify that Cage B provides space for large adsorption of both C2H6 and C2H4 with little difference in terms of adsorption heat and no competition for strong adsorption sites. Meanwhile, the single-molecule passage Cage A can provide multiple hydrogen-bonding interactions with both C2H6 and C2H4, and the overall energy diagram turns out to be more conducive to the adsorption and diffusion of C2H6, resulting in a kinetic selectivity of C2H6 over C2H4. This work successfully illustrated the underlying kinetic separation mechanism of C2H6 over C2H4 on JNU-2 by adopting multi-scale simulations, demonstrating a new kinetic separation mechanism of interaction screening and providing an effective theoretical methodology for better understanding the gas adsorption and separation in MOFs.

Computational details

GCMC simulation

Grand Canonical Monte Carlo (GCMC) calculations were performed with the Sorption module embedded in Materials Studio 2018 and RASPA 2.023 to simulate the C2H6/C2H4 adsorption properties of JNU-2. In GCMC simulations, the JNU-2 structure was taken from the experimental crystallographic data in our previous work, and the conventional cubic unit cell of JNU-2 (a = b = c = 43.55 Å, α = β = γ = 90°) was utilized; the periodic boundary conditions were applied in all the three directions. The JNU-2 structure was kept rigid by constructing the atoms in the JNU-2 structure in simulations. All GCMC simulations including 2.5 × 106 equilibration cycles followed by 2.5 × 106 production cycles were carried out at 298 K and various pressure points from 0.01 bar to 1.0 bar. The intermolecular interactions were represented using a Lennard-Jones (LJ) potential which is defined as follows:

The LJ parameters for atoms in JNU-2 were all taken from the universal force field (UFF),24–26 and the LJ parameters for ethane and ethylene were taken from the TraPPE force field.27–29 The combined UFF/TraPPE force field is widely used to predict adsorption properties in the MOF research field.30 The Lorentz–Berthelot mixing rules31 were applied in describing the cross interactions between different atom types. The Ewald summation was used to calculate the electrostatic interactions. The charge equilibration (Qeq) method32 was applied to compute the atomic partial charges for JNU-2, and the atomic charges for methane/ethylene were calculated by employing the density functional theory (DFT) at the B3LYP33/def2-TZVP34 theoretical level. Table S2 lists the atomic charges of methane and ethylene. A cutoff of 12.5 Å in the interaction distance was used in all GCMC simulations.

Pore volume and window diameter analysis

The isolated cage structures of Cage A, B, and C were intercepted from the crystal structure of JNU-2. The cavity volumes of the cages were estimated using the VIODOO program35 using a 2.8 Å probe radius. Cage A was visualized by using the ChimeraX 1.0 program.36 The cross-sections of the accessible volume and the window position of each cage were made by utilizing the “slab” tool in ChimeraX, so that the window size could be obtained.

MD simulation

The simulation models are composed of a mixed gas layer, framework layer, and vacuum layer. The z-direction of the triclinic simulation box (α = β = γ = 90°) is fixed to 200 Å. The framework layers are cleaved from the crystal structure of JNU-2 including at least one group of adjacent cavities in each channel. The size of the framework layer along the z-direction is controlled at about 50 Å. The broken chemical bonds in the cleaved surface are saturated with hydrogen or methyl. The two frame layers on both sides are surface symmetrical so that the two interfaces in contact with the gas molecules are the same. The mixed gas layers are set to be about 50 Å in the z-direction and randomly filled with C2H6 and C2H4 molecules with the same molar ratio maintaining the gas densities in the three models at about 0.08 g cm−3. The vacuum layers on both sides are set to be about 25 Å in the z-direction. Periodic boundary conditions (PBC) are applied in the x- and y-directions, so that the gas molecules can diffuse along the z-direction into the framework layer and finally reach the vacuum layer in the presence of the gas pressure difference.

In MD simulations, a flexible JNU-2 model was adopted for the framework layer. The structural parameter of the UFF4MOF force field25,26 was used for Cu and Zn ions, which was specially made for MOFs, while the other atoms of JNU-2 adopting the UFF force field and the TraPPE force field was used for C2H6 and C2H4. The charge calculation and equilibration method here was consistent with the GCMC simulation. The Lorentz–Berthelot combination rules were applied to obtain the LJ cross potential parameters for intermolecular interactions. The Ewald summation was used to calculate the electrostatic interactions. In addition, a discrete Cage A structure from JNU-2 was placed at the center of a box (30 Å × 30 Å × 30 Å) filled with a C2H6/C2H4 (1[thin space (1/6-em)]:[thin space (1/6-em)]1) mixture that was randomly generated and annealed. Ten parallel 500 ps MD simulations were carried out by employing the same parameters as above.

All the MD simulations were performed with the Forcite module in the Materials Studio 2018 program. The framework layers were first optimized to a convergence tolerance of an energy difference less than 0.001 kcal mol−1 and force less than 0.5 kcal mol−1 Å−1. The mixed gas molecules are then added into the mixed gas layer in the middle of the simulation box. Three-channel models were simulated for 20 ns with a time step of 2.0 fs using a canonical ensemble (NVT) with an Anderson thermostat at 298 K. The concentrations of gas molecules in all three models were analyzed based on the trajectory of MD simulation in the specified x-, y- and z-directions.

DFT calculations

Density functional theory (DFT) calculations were performed to assess the behavior of gas adsorption in Cage B and energy changes of C2H4 and C2H6 when passing through Cage A. The cluster models of Cage A and Cage B used in DFT calculations were taken from the crystal structure of JNU-2; the truncated chemical bonds were saturated with hydrogen or methyl. The Cage B–absorbate interaction model was constructed by fixing the geometry of Cage B and randomly adding the gas molecules one by one, and the final interaction configurations were obtained by optimization. As for Cage A, we conducted a relaxed potential surface scanning on the energy of the gas crossing process using modredundant calculations. The distance between the center of the gas molecule and the center of Cage A (Dcc) was set as the scan variable, making the gas molecule move along the z-axis and optimize the structure of the gas molecule and calculate the energy, so that the whole crossing process, i.e.Cage BWindow 1Cage AWindow 1Cage B, was included in the scanning. Furthermore, we calculated the interaction energies between the gas molecule and Cage A in each scanning step and the basis set superposition error (BSSE) was considered herein. Optimization and energy calculations were performed using Gaussian09 program37 employing the B3LYP functional33 with Grimme's dispersion correction38–40 (B3LYP–D3(BJ)). The effective core potential LanL2DZ41 and the corresponding basis set were used for the Zn and Cu atom, and the double zeta basis set 6-31G(d)42 was used for other atoms. The energy decomposition analysis (EDA)43 calculations were performed using Amsterdam Density Functional (ADF) 2019 suit of the program44 at the B3LYP–D3(BJ)/TZP theoretical level without a frozen core. The EDA scheme divides the total interaction energy into the following items:
ΔEtot = ΔEele + ΔErep + ΔEorb + ΔEdisp
where ele, rep, orb, and disp denote the electrostatic interaction, Pauli repulsion, orbital interaction, and dispersion, respectively.

Author contributions

M. Xie and Z. Lu performed all calculations, data curation, and visualization and wrote the original draft. M. Xie and W. Lu conceived and conceptualized the research. W. Lu and D. Li supervised the research and reviewed and polished the manuscript.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China (No. 21731002, 21975104, 22150004, and 22101099), the Guangdong Major Project of Basic and Applied Research (No. 2019B030302009), and the Guangdong Basic and Applied Basic Research Foundation (No. 2020A1515011005). We thank the high-performance public computing service platform of Jinan University for providing computational resources. We also appreciate Xiao-Jing Xie (Jinan University) and Dr. Weijie Zhang (Univ. of North Texas) for their help and discussion.

Notes and references

  1. J.-R. Li, J. Sculley and H.-C. Zhou, Metal–Organic Frameworks for Separations, Chem. Rev., 2012, 112, 869–932 CrossRef CAS PubMed.
  2. K. Sumida, D. L. Rogow, J. A. Mason, T. M. McDonald, E. D. Bloch, Z. R. Herm, T.-H. Bae and J. R. Long, Carbon Dioxide Capture in Metal–Organic Frameworks, Chem. Rev., 2012, 112, 724–781 CrossRef CAS PubMed.
  3. Y. He, W. Zhou, G. Qian and B. Chen, Methane Storage in Metal–Organic Frameworks, Chem. Soc. Rev., 2014, 43, 5657–5678 RSC.
  4. H. Furukawa, K. E. Cordova, M. O'Keeffe and O. M. Yaghi, The Chemistry and Applications of Metal-Organic Frameworks, Science, 2013, 341, 1230444 CrossRef PubMed.
  5. W.-G. Cui, T.-L. Hu and X.-H. Bu, Metal–Organic Framework Materials for the Separation and Purification of Light Hydrocarbons, Adv. Mater., 2020, 32, 1806445 CrossRef CAS PubMed.
  6. R.-B. Lin, S. Xiang, W. Zhou and B. Chen, Microporous Metal-Organic Framework Materials for Gas Separation, Chem, 2020, 6, 337–363 CAS.
  7. A. J. Rieth, A. M. Wright and M. Dincă, Kinetic Stability of Metal–Organic Frameworks for Corrosive and Coordinating Gas Capture, Nat. Rev. Mater., 2019, 4, 708–725 CrossRef CAS.
  8. H. Zeng, M. Xie, T. Wang, R.-J. Wei, X.-J. Xie, Y. Zhao, W. Lu and D. Li, Orthogonal-Array Dynamic Molecular Sieving of Propylene/Propane Mixtures, Nature, 2021, 595, 542–548 CrossRef CAS PubMed.
  9. J.-R. Li, R. J. Kuppler and H.-C. Zhou, Selective Gas Adsorption and Separation in Metal–Organic Frameworks, Chem. Soc. Rev., 2009, 38, 1477–1504 RSC.
  10. L. Yang, S. Qian, X. Wang, X. Cui, B. Chen and H. Xing, Energy-Efficient Separation Alternatives: Metal–Organic Frameworks and Membranes for Hydrocarbon Separation, Chem. Soc. Rev., 2020, 49, 5359–5406 RSC.
  11. Ü. Kökçam-Demir, A. Goldman, L. Esrafili, M. Gharib, A. Morsali, O. Weingart and C. Janiak, Coordinatively Unsaturated Metal Sites (Open Metal Sites) in Metal–Organic Frameworks: Design and Applications, Chem. Soc. Rev., 2020, 49, 2751–2798 RSC.
  12. E. D. Bloch, W. L. Queen, R. Krishna, J. M. Zadrozny, C. M. Brown and J. R. Long, Hydrocarbon Separations in a Metal-Organic Framework with Open Iron(II) Coordination Sites, Science, 2012, 335, 1606–1610 CrossRef CAS PubMed.
  13. R.-B. Lin, L. Li, H.-L. Zhou, H. Wu, C. He, S. Li, R. Krishna, J. Li, W. Zhou and B. Chen, Molecular Sieving of Ethylene from Ethane Using a Rigid Metal–Organic Framework, Nat. Mater., 2018, 17, 1128–1133 CrossRef CAS PubMed.
  14. D.-D. Zhou, P. Chen, C. Wang, S.-S. Wang, Y. Du, H. Yan, Z.-M. Ye, C.-T. He, R.-K. Huang, Z.-W. Mo, N.-Y. Huang and J.-P. Zhang, Intermediate-Sized Molecular Sieving of Styrene from Larger and Smaller Analogues, Nat. Mater., 2019, 18, 994–998 CrossRef CAS PubMed.
  15. A. L. Dzubak, L.-C. Lin, J. Kim, J. A. Swisher, R. Poloni, S. N. Maximoff, B. Smit and L. Gagliardi, Ab initio Carbon Capture in Open-Site Metal–Organic Frameworks, Nat. Chem., 2012, 4, 810–816 CrossRef CAS PubMed.
  16. W. Chen, L. Huang, X. Yi and A. Zheng, Lithium Doping on 2D Squaraine-Bridged Covalent Organic Polymers for Enhancing Adsorption Properties: a Theoretical Study, Phys. Chem. Chem. Phys., 2018, 20, 6487–6499 RSC.
  17. D. J. Vogel, J. M. Rimsza and T. M. Nenoff, Prediction of Reactive Nitrous Acid Formation in Rare–Earth MOFs via ab[thin space (1/6-em)]initio Molecular Dynamics, Angew. Chem., Int. Ed., 2021, 60, 11514–11522 CrossRef CAS PubMed.
  18. Y. Ying, Z. Zhang, S. B. Peh, A. Karmakar, Y. Cheng, J. Zhang, L. Xi, C. Boothroyd, Y. M. Lam, C. Zhong and D. Zhao, Pressure–Responsive Two-Dimensional Metal–Organic Framework Composite Membranes for CO2 Separation, Angew. Chem., Int. Ed., 2021, 60, 11318–11325 CrossRef CAS PubMed.
  19. W.-Q. Lin, X.-L. Xiong, H. Liang and G.-H. Chen, Multiscale Computational Screening of Metal–Organic Frameworks for Kr/Xe Adsorption Separation: A Structure–Property Relationship-Based Screening Strategy, ACS Appl. Mater. Interfaces, 2021, 13, 17998–18009 CrossRef CAS PubMed.
  20. H. Zeng, X.-J. Xie, M. Xie, Y.-L. Huang, D. Luo, T. Wang, Y. Zhao, W. Lu and D. Li, Cage-Interconnected Metal–Organic Framework with Tailored Apertures for Efficient C2H6/C2H4 Separation under Humid Conditions, J. Am. Chem. Soc., 2019, 141, 20390–20396 CrossRef CAS PubMed.
  21. S. L. Mayo, B. D. Olafson and W. A. Goddard, DREIDING: A Generic Force Field for Molecular Simulations, J. Phys. Chem., 1990, 94, 8897–8909 CrossRef CAS.
  22. W. L. Jorgensen, D. S. Maxwell and J. Tirado-Rives, Development and Testing of the OPLS All-Atom Force Field on Conformational Energetics and Properties of Organic Liquids, J. Am. Chem. Soc., 1996, 118, 11225–11236 CrossRef CAS.
  23. D. Dubbeldam, S. Calero, D. E. Ellis and R. Q. Snurr, RASPA: Molecular Simulation Software for Adsorption and Diffusion in Flexible Nanoporous Materials, Mol. Simul., 2016, 42, 81–101 CrossRef CAS.
  24. A. K. Rappe, C. J. Casewit, K. S. Colwell, W. A. Goddard and W. M. Skiff, UFF, a Full Periodic Table Force Field for Molecular Mechanics and Molecular Dynamics Simulations, J. Am. Chem. Soc., 1992, 114, 10024–10035 CrossRef CAS.
  25. M. A. Addicoat, N. Vankova, I. F. Akter and T. Heine, Extension of the Universal Force Field to Metal–Organic Frameworks, J. Chem. Theory Comput., 2014, 10, 880–891 CrossRef CAS PubMed.
  26. D. E. Coupry, M. A. Addicoat and T. Heine, Extension of the Universal Force Field for Metal–Organic Frameworks, J. Chem. Theory Comput., 2016, 12, 5215–5225 CrossRef CAS PubMed.
  27. B. Chen and J. I. Siepmann, Transferable Potentials for Phase Equilibria. 3. Explicit-Hydrogen Description of Normal Alkanes, J. Phys. Chem. B, 1999, 103, 5370–5379 CrossRef CAS.
  28. M. G. Martin and J. I. Siepmann, Transferable Potentials for Phase Equilibria. 1. United-Atom Description of n-Alkanes, J. Phys. Chem. B, 1998, 102, 2569–2577 CrossRef CAS.
  29. C. D. Wick, M. G. Martin and J. I. Siepmann, Transferable Potentials for Phase Equilibria. 4. United-Atom Description of Linear and Branched Alkenes and Alkylbenzenes, J. Phys. Chem. B, 2000, 104, 8008–8016 CrossRef CAS.
  30. J. G. McDaniel, S. Li, E. Tylianakis, R. Q. Snurr and J. R. Schmidt, Evaluation of Force Field Performance for High-Throughput Screening of Gas Uptake in Metal–Organic Frameworks, J. Phys. Chem. C, 2015, 119, 3143–3152 CrossRef CAS.
  31. C. L. Kong, Combining Rules for Intermolecular Potential Parameters. II. Rules for the Lennard-Jones (12–6) Potential and the Morse Potential, J. Chem. Phys., 1973, 59, 2464–2467 CrossRef CAS.
  32. A. K. Rappe and W. A. Goddard, Charge Equilibration for Molecular Dynamics Simulations, J. Phys. Chem., 1991, 95, 3358–3363 CrossRef CAS.
  33. A. D. Becke, Density–Functional Thermochemistry. III. The Role of Exact Exchange, J. Chem. Phys., 1993, 98, 5648–5652 CrossRef CAS.
  34. F. Weigend, Accurate Coulomb-Fitting Basis Sets for H to Rn, Phys. Chem. Chem. Phys., 2006, 8, 1057–1065 RSC.
  35. G. J. Kleywegt and T. A. Jones, Detection, Delineation, Measurement and Display of Cavities in Macromolecular Structures, Acta Crystallogr., Sect. D: Biol. Crystallogr., 1994, 50, 178–185 CrossRef CAS PubMed.
  36. T. D. Goddard, C. C. Huang, E. C. Meng, E. F. Pettersen, G. S. Couch, J. H. Morris and T. E. Ferrin, UCSF ChimeraX: Meeting Modern Challenges in Visualization and Analysis, Protein Sci., 2018, 27, 14–25 CrossRef CAS PubMed.
  37. G. W. Trucks, M. J. Frisch, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta, F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin, V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A. Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels, Ö. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski and D. J. Fox, Gaussian09, Inc., Wallingford CT, 2009 Search PubMed.
  38. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, A Consistent and Accurate ab initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
  39. S. Grimme, Accurate Description of van der Waals Complexes by Density Functional Theory Including Empirical Corrections, J. Comput. Chem., 2004, 25, 1463–1473 CrossRef CAS PubMed.
  40. S. Grimme, Semiempirical GGA-Type Density Functional Constructed with a Long-Range Dispersion Correction., J. Comput. Chem., 2006, 27, 1787–1799 CrossRef CAS PubMed.
  41. L. E. Roy, P. J. Hay and R. L. Martin, Revised Basis Sets for the LANL Effective Core Potentials, J. Chem. Theory Comput., 2008, 4, 1029–1031 CrossRef CAS PubMed.
  42. P. C. Hariharan and J. A. Pople, The Influence of Polarization Functions on Molecular Orbital Hydrogenation Energies, Theor. Chim. Acta, 1973, 28, 213–222 CrossRef CAS.
  43. L. Zhao, M. von Hopffgarten, D. M. Andrada and G. Frenking, Energy Decomposition Analysis, Wiley Interdiscip. Rev.: Comput. Mol. Sci., 2017, 8, e1345 Search PubMed.
  44. T. Ziegler, E. J. Baerends, J. Autschbach, D. Bashford, A. Bérces, F. M. Bickelhaupt, C. Bo, P. M. Boerrigter, L. Cavallo, D. P. Chong, L. Deng, R. M. Dickson, D. E. Ellis, M. van Faassen, L. Fan, T. H. Fischer, C. Fonseca Guerra, M. Franchini, A. Ghysels, A. Giammona, S. J. A. van Gisbergen, A. W. Götz, J. A. Groeneveld, O. V. Gritsenko, M. Grüning, S. Gusarov, F. E. Harris, P. van den Hoek, C. R. Jacob, H. Jacobsen, L. Jensen, J. W. Kaminski, G. van Kessel, F. Kootstra, A. Kovalenko, M. V. Krykunov, E. van Lenthe, D. A. McCormack, A. Michalak, M. Mitoraj, S. M. Morton, J. Neugebauer, V. P. Nicu, L. Noodleman, V. P. Osinga, S. Patchkovskii, M. Pavanello, P. H. T. Philipsen, D. Post, C. C. Pye, W. Ravenek, J. I. Rodríguez, P. Ros, P. R. T. Schipper, H. van Schoot, G. Schreckenbach, J. S. Seldenthuis, M. Seth, J. G. Snijders, M. Solà, M. Swart, D. Swerhone, G. te Velde, P. Vernooijs, L. Versluis, L. Visscher, O. Visser, F. Wang, T. A. Wesolowski, E. M. van Wezenbeek, G. Wiesenekker, S. K. Wolff, T. K. Woo and A. L. Yakovlev, ADF 2018, SCM, Theoretical Chemistry, Vrije Universiteit, Amsterdam, The Netherlands, URL:https://www.scm.com Search PubMed.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2qi00465h
Present address: Department of Chemistry, University of North Texas, Denton, Texas 76203, USA.

This journal is © the Partner Organisations 2022