Next Article in Journal
Mineral Acid Co-Extraction in Reactive Extraction of Lactic Acid Using a Thymol-Menthol Deep Eutectic Solvent as a Green Modifier
Next Article in Special Issue
A Polyzwitterionic@MOF Hydrogel with Exceptionally High Water Vapor Uptake for Efficient Atmospheric Water Harvesting
Previous Article in Journal
The Piezocatalytic Degradation of Sulfadiazine by Lanthanum-Doped Barium Titanate
Previous Article in Special Issue
Zeolite Properties, Methods of Synthesis, and Selected Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Study of CHF3/CH2F2 Adsorption Separation in TIFSIX-2-Cu-i

1
School of Chemistry and Chemical Engineering, Shandong University of Technology, Zibo 255049, China
2
Shandong Dongyue Organosilicon Materials Co., Ltd., Zibo 256401, China
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(8), 1721; https://doi.org/10.3390/molecules29081721
Submission received: 2 March 2024 / Revised: 3 April 2024 / Accepted: 9 April 2024 / Published: 11 April 2024
(This article belongs to the Special Issue Zeolites and Porous Materials: Synthesis, Properties and Applications)

Abstract

:
Hydrofluorocarbons (HFCs) have important applications in different industries; however, they are environmentally unfriendly due to their high global warming potential (GWP). Hence, reclamation of used hydrofluorocarbons via energy-efficient adsorption-based separation will greatly contribute to reducing their impact on the environment. In particular, the separation of azeotropic refrigerants remains challenging, such as typical mixtures of CH2F2 (HFC-23) and CHF3 (HFC-32), due to a lack of adsorptive mechanisms. Metal–organic frameworks (MOFs) can provide a promising solution for the separation of CHF3–CH2F2 mixtures. In this study, the adsorption mechanism of CHF3–CH2F2 mixtures in TIFSIX-2-Cu-i was revealed at the microscopic level by combining static pure-component adsorption experiments, molecular simulations, and density-functional theory (DFT) calculations. The adsorption separation selectivity of CH2F2/CHF3 in TIFSIX-2-Cu-i is 3.17 at 3 bar under 308 K. The existence of similar TiF62− binding sites for CH2F2 or CHF3 was revealed in TIFSIX-2-Cu-i. Interactions between the fluorine atom of the framework and the hydrogen atom of the guest molecule were found to be responsible for determining the high adsorption separation selectivity of CH2F2/CHF3. This exploration is important for the design of highly selective adsorbents for the separation of azeotropic refrigerants.

1. Introduction

HFCs are third-generation fluorinated gases (F-gases), a class of synthetic compounds used primarily in refrigeration and air conditioning (RAC) [1,2,3]. HFCs are potent greenhouse gases. So, their production and application must be phased down to meet the emission reduction target according to the Montreal Protocol [4,5]. Depending on the actual production and use of refrigerants, many of the HFCs currently in use are azeotropic or near-azeotropic refrigerant mixtures [6]. The current mainstream refrigerants include R-444A, R-447A, and R-448A, which are blends of HFCs (R-32, R-125, R-23, R-134a, etc.) with hydrofluoroolefins (HFOs). Difluoromethane (CH2F2, GWP = 675) and trifluoromethane (CHF3, GWP = 14,800) are the most common components of refrigerant mixture currently used in the refrigeration and air conditioning industry, with very high GWP [7,8]. Therefore, to control HFC emissions, the first thing is to separate the components from the mixtures efficiently. However, due to the low efficiency of cryogenic distillation to separate refrigerant mixtures, the amount of refrigerant gas recovered remains low. There are very similar physical properties and molecular dynamics diameters for CHF3/CH2F2 molecules, which makes the search for alternative technologies to energy-intensive distillation processes very challenging. In this study, CHF3/CH2F2 was chosen as a sample to study the separation mechanism to provide a reference for the subsequent separation of HFCs. Selective adsorption technology has become an attractive solution for gas separation considering energy efficiency and environmental protection [9,10,11,12]. Metal–organic framework (MOF) materials show great promise for gas storage and separation applications due to their significant advantages, such as flexible framework, tunable pore size and structure, and ultra-high specific surface area [13,14,15,16,17,18,19].
However, the practical applications of some MOFs are limited by their poor structural stability due to strong dependence on solvent molecules. The framework structure will collapse if they are exposed to air, high-strength acids, and bases for a period of time. TIFSIX-2-Cu-i is easy to regenerate and thermally stable under air atmosphere [20,21]. The efficient separation of C2H2/C2H4 by TIFSIX-2-M-i has been demonstrated by previous studies [20,22,23]. Inspired by these findings, TIFSIX-2-Cu-i was chosen to study the separation of CHF3–CH2F2 mixtures. To our knowledge, previous work on the mechanism of CHF3/CH2F2 adsorption in TIFSIX-2-Cu-i is sparse, which is disadvantageous to understanding and predicting interactions between adsorbates and adsorbents. In this study, the feasibility of selective separation of CHF3/CH2F2 by TIFSIX-2-Cu-i was evaluated for the first time. The method of adding polarization to a generic forcefield was used to obtain simulated and experimentally consistent adsorption isotherms, ensuring the accuracy of the forcefield. In this work, TIFSIX-2-Cu-i exhibits preferential adsorption of CH2F2 over CHF3 with a high CH2F2 adsorption capacity (2.70 mmol/g at 298 K and 1 bar). Thermodynamic and kinetic analyses were carried out by a combination of adsorption experiments and molecular simulations. The adsorption selectivity, isosteric adsorption heat, and binding sites were investigated. In addition, these works are of great significance for exploring the adsorption separation of HFCs by fluorinated anion MOFs.

2. Results and Discussion

2.1. Adsorption Isotherm

To verify the accuracy of the forcefield, the present work compares the simulation adsorption isotherms for pure CHF3 or CH2F2 in TIFSIX-2-Cu-i at 288, 298, and 308 K with experimental adsorption isotherms (Figure 1). The simulation results under pressure values from 0 bar to 3 bar are in good agreement with the experimental data. Therefore, it can be inferred that the potential model and polarization forcefield parameters used are reliable for predicting the adsorption of CHF3 and CH2F2. For all three temperatures, the trends of the CHF3 and CH2F2 isotherms are similar. These adsorption isotherms were fitted with the Langmuir isotherm model. The Langmuir equation is defined as shown in Equation (1). The parameters of the isotherms for CHF3 and CH2F2 are summarized in Table 1.
qe = qmbp/(1 + bp)
where qe is the equilibrium adsorption capacity, qm is the maximum adsorption capacity, b is the adsorbate–adsorbent affinity coefficient, and p is the equilibrium pressure. From Table 1, parameter b decreases with increasing temperature at the same adsorbate; parameter b is consistently larger for CH2F2 than for CHF3 at all three temperatures. The interaction of CH2F2 with the framework was stronger compared to CHF3. Parameter qm shows a larger maximum adsorption capacity for CH2F2.
In detail, the uptake of CHF3 and CH2F2 reached 1.73 and 2.70 mmol/g at 1 bar and 298 K, respectively. At 308 K and 3 bar, the uptake of CH2F2 was 3.79 mmol/g, which is almost double the CHF3 uptake (1.99 mmol/g). At the same temperature and pressure, the adsorption capacity of CH2F2 was higher than that of CHF3. Thus, there was a greater adsorption affinity for CH2F2 than CHF3. The adsorption capacity of both CHF3 and CH2F2 increased significantly with increasing pressure. However, the adsorption capacity of CH2F2 increased faster compared to that of CHF3. These findings suggest that TIFSIX-2-Cu-i has stronger binding ability regarding CH2F2, indicating that TIFSIX-2-Cu-i is a potential material to separate CHF3/CH2F2 mixtures with high efficiency.

2.2. Adsorption Selectivity and Heat

In this section, thermodynamic adsorption selectivity and isosteric adsorption heat were explored, which were calculated from experimental measurements. In addition, molecular dynamics simulations were performed to explore the diffusivity of guest molecules in TIFSIX-2-Cu-i.
Myers and Praunitz developed ideal adsorbed solution theory (IAST). Here, the multicomponent adsorption equilibrium of CHF3 or CH2F2 was predicted using ideal adsorbed solution theory (IAST), which was calculated by the following equation [24]:
S = (x1/x2)/(y1/y2)
where S is the selectivity of a component versus another one (e.g., CH2F2/CHF3), x is the molar fraction in the adsorbed phase, and y is the molar fraction in the gas phase. The relatively high adsorption separation selectivity is shown in Figure 2. The selectivity of CH2F2/CHF3 is greater than 1.0 at all the adsorption isotherms, indicating that TIFSIX-2-Cu-i preferentially adsorbs CH2F2. At 298 K and 308 K, selectivity increases with increasing pressure, while, at 288 K, selectivity decreases slightly with increasing pressure. At 288 K, the adsorption amount of CH2F2 in the high-pressure zone flattens more as pressure rises compared to CHF3. We hypothesize that most of the adsorption sites of the framework were then occupied by CH2F2 molecules, making it difficult for the newly added CH2F2 molecules to find available adsorption sites, leading to slowdown of the adsorption rate.
To assess the interaction strength between the framework and gas molecules, we utilized single-component isotherms obtained at three distinct temperatures (Figure 1) to determine the isosteric adsorption heat (Qst) of CH2F2/CHF3 on TIFSIX-2-Cu-i. The isosteric adsorption heat was calculated indirectly using the Clausius–Clapeyron equation [25]:
d   l n P d T = q i   R T
where qi refers to the isosteric heat of adsorption, kJ/mol; P is the pressure, MPa; T is the temperature, K; and R is the gas constant, 8.314 J/(mol·K). As shown in Figure 3, the Qst values of CHF3 and CH2F2 were around 20 and 23 kJ/mol−1. The heat of adsorption of CH2F2 was always higher than the heat of adsorption of CHF3 in TIFSIX-2-Cu-i. Therefore, it is directly verified that TIFSIX-2-Cu-i interacts more strongly with CH2F2 than CHF3, which leads to greater adsorption of CH2F2 than CHF3.
In this part, the free diffusion behavior of CHF3 and CH2F2 in TIFSIX-2-Cu-i was explored. The corresponding mean square displacements (MSDs) obtained from the simulations are shown in Figure 4. The self-diffusion coefficients of CHF3 and CH2F2 in the TIFSIX-2-Cu-i were calculated using the Einstein relation as shown below:
D = 1 6 lim n d d t 1 N i i = 1 N i r i t r i 0 2
where the average is taken over time t for the mean square displacement of the center of mass position vectors r of all the molecules N in the system; ‹›indicates the overall average. The calculated results show that the self-diffusion coefficients of CHF3 and CH2F2 are 1.15 × 10−4 cm2/s and 2.18 × 10−4 cm2/s, respectively. This finding showed that TIFSIX-2-Cu-i exhibited high kinetic selectivity for CH2F2 over CHF3. TIFSIX-2-Cu-i is a doubly interpenetrated framework attributed to the much longer organic linker 4,4-dipyridylacetylene and slightly large pore sizes of about 5.2 × 5.2 Å. That makes it easier for the mixture to diffuse into the adsorption sites within the pores. The properties of this MOF can be characterized as pillared square lattice networks with a pcu topology, attributed to their pore surfaces with narrow pore sizes and highly electrostatic pore surfaces. These features combined provide exceptionally strong binding interactions with polarizable molecules, such as CHF3 and CH2F2 [26]. This also enabled the two guest molecules to possess high adsorption capacity.

2.3. Adsorption Sites

Figure 5 shows the optimal adsorption binding sites for CHF3 and CH2F2. The snapshots obtained detailed information about the adsorption of pure CHF3 and CH2F2 in TIFSIX-2-Cu-i at 298 K and 1 bar. In the doubly interpenetrated framework of TIFSIX-2-Cu-i, the H atom of CHF3 binds with the F atom from TiF62−. The distance of the C–H⋯F hydrogen bond was 2.372 Å (Figure 5b). The H⋯F distance of 2.372 Å obtained by the simulation was smaller than the sum of the van der Waals radii of H and F (2.55 Å) [27], confirming the existence of electrostatic interactions for Hδ+⋯Fδ−. This was consistent with the reported binding sites of TIFSIX-2-Cu-i to short-chain alkanes (C2H2, C2H4) [20,22,23]. CH2F2 has a similar binding site. The two H atoms of CH2F2 are bound at the F site by virtue of a synergistic hydrogen bonding interaction. The shortest length of the C–H⋯F bond between CH2F2 and the TIF62− site is 2.172 Å, which is shorter than the C–H⋯F between CHF3 and TIF62−. The TIFSIX-2-Cu-i interacts more strongly with CH2F2 than CHF3.
Radial Distribution Functions (RDFs) amount to one of the most common methods for determination of interatomic distances. The RDFs in 298 K describing the interactions between the individual atoms of the pure CHF3/CH2F2 and the TIFSIX-2-Cu-i framework are shown in Figure 6a,b. In Figure 6a, the framework interacts preferentially with H in CHF3. From the investigation of the RDF of H (CHF3) with each atom of the framework in Figure 6c, it was found that F (framework) interacts preferentially with H (CHF3), which is the same as the snapshot conclusion of Figure 5a. Similarly, according to Figure 6b,d, it is found that F (framework) interacts preferentially with H (CH2F2). Compared to CHF3, CH2F2 has more H–F bonds interacting with the framework at the same time. So, we believe that the reason for the higher adsorption affinity of CH2F2 than CHF3 is due to the binding geometry of CH2F2/CHF3 adsorbed in the supercage of TIFSIX-2-Cu-i.

2.4. Redistribution of Charge Density

In order to study the charge change in TIFSIX-2-Cu-i after adsorption, DFT calculations were carried out to investigate the redistribution of charge density in this system after adsorption of CHF3/CH2F2 molecules. As shown in Figure 7, the electrons of the H atoms of CHF3/CH2F2 migrate to the F atoms of the framework due to the strong electron-withdrawing ability of the F atoms, where “–” denotes a negative charge. As shown in Figure 8, the blue area in Figure 8b is larger and darker than in Figure 8a, and the charge density transfer between CH2F2 and the framework F atom is more pronounced. This may be because, compared to CHF3, CH2F2 has more interactions between H atoms and framework F atoms, which is also the reason for the relatively high adsorption separation selectivity for CH2F2 over CHF3.

3. Experiment Section

3.1. Preparation of Materials

All reagents and solvents were obtained commercially and used as received without further purification. Copper fluoroborate [Cu(BF4)2, 98.5%], methanol (CH3OH, 99%), and ammonium hexafluoro titanate ((NH4)2TiF6, 98%) were bought from Aladdin; 1,2-Di(pyridin-4yl)ethyne (C12H8N2, 99.35%) was purchased from Leyan Co., Ltd. (Shanghai, China); and helium (He, 99.999%) gas was purchased from BaiYan Co., Ltd. (Zibo, China); The methylene fluoride (CH2F2, 99.999%) and the methyl trifluoride (CHF3, 99.999%) were obtained from Dong Yue Co., Ltd. (Zibo, China).

3.2. Synthetic Procedures

Cu(BF4)2 (1 mmol), (NH4)2TiF6 (1 mmol), and 1,2-Di(pyridin-4yl) ethyne (2 mmol) were dissolved in 5 mL of water and 10 mL of methanol, and a blue slurry product was obtained after stirring at 338 K for 12 h. Then, the slurry was filtered and washed with 10 mL of methanol. The blue filter cake was heated at 393 K for 12 h under vacuum conditions to obtain TIFSIX-2-Cu-i material [26].

3.3. Characterization

The activated TIFSIX-2-Cu-i was subjected to X-ray diffraction characterization and scanned using a Bruker AXS D8 ADVANCE diffractometer under a CuKα radiation source operated at a voltage of 40 kV, a current of 20 mA, and a scattering angle in the range (2θ) of 5–40 degrees. The XRD pattern of TIFSIX-2-Cu-i was presented in Figure 9, which was compared to calculated patterns. ASAP 2460 (Micromeritics company, Shanghai, China) was employed in this experiment. Pore structure was characterized by the N2 adsorption method. The experimental temperature was 77 K. Before testing, the sample was treated by 12 h vacuumization at 393 K. Specific surface area was calculated using a multipoint Brunauer–Emmett–Teller model (BET). Table 2 shows the pore parameters of TIFSIX-2-Cu-i.

3.4. Single-Component Adsorption Measurements

The adsorption isotherms of CHF3 and CH2F2 were measured in the absolute pressure range of 1–3 bar in TIFSIX-2-Cu-i framework. The experimental temperatures were 288, 298, and 308 K. Excess adsorption experiments were performed using activated TIFSIX-2-Cu-i monomer. The temperature was controlled by an external circulating water bath. Before the measurements, TIFSIX-2-Cu-i was degassed at 393 K for 12 h under vacuum conditions. CHF3 and CH2F2 gas of purity 99.99% were used as adsorbates. The adsorption capacity of pure CHF3 or CH2F2 was calculated based on the pressure changes before and after adsorption. Figure 10 shows the diagram of adsorption measurements’ experimental apparatus. A standard volumetric method was used to measure pure gas adsorption isotherms [28]. The homemade apparatus is designed by the proposed method.

4. Models and Methods

4.1. Models

The interpenetrated polymorph, TIFSIX-2-Cu-i, is composed of doubly interpenetrated nets that are isostructural to the nets in TIFSIX-2-Cu. The independent nets are staggered with respect to one another, affording 5.2 Å pores [22,23]. The crystal cells used in the simulation were downloaded from the Cambridge Crystallographic Data Center (CCDC) as structural files. Optimized structure by DFT simulation was used for further calculations. TIFSIX-2-Cu-i is a variant of SIFSIX-2-Cu-i. Ti4+ has a higher polarizability relative to Si4+ [29]. So, TIFSIX-2-Cu-i has a higher thermal stability (decomposition temperature of 262 °C), which may be attributed to the relatively higher polarizability of Ti4+ [20]. TIFSIX-2-Cu-i atoms in the framework with different chemical properties are shown in Figure 11.

4.2. Density Functional Theory Calculations

The MOFs’ structure was optimized using ab initio density functional theory (DFT) as implemented in the Vienna Ab Initio Simulation Package version 2.2 (VASP) [30], with the overall energy converged to within 10−5 eV per atom. The Perdew–Burke–Ernzerhof (PBE) function of the generalized gradient approximation (GGA) [31] was used to represent the electron exchange correlation, and a cutoff energy of 500 eV was set for the plane wave. According to the Monkhorst–Pack methodology, the Brillouin zone was sampled with a series of K-point grids (2 × 2 × 4). After geometrical optimization, we obtained the electron charge density and then used Density Derived Electrostatic and Chemical (DDEC6) [32,33] to calculate the atomic charge of the net atomic charge framework for each MOF atom (Table 3). DFT simulations were used to explore the redistribution of charge density in this system after adsorption of CHF3 or CH2F2 molecules using the CP2K code [34]. The DZVP-MOLOPT-SR-GTH [35] basis set and the Goedecker–Teter–Hutter [36] pseudopotential were used, and the density generalization employed was a PBE with DFT-D3 [37] dispersion corrections.

4.3. Grand Canonical Monte Carlo Simulations

All GCMC simulations were performed with the RASPA [38] code to study CHF3/CH2F2 adsorption properties under different conditions. A grand canonical systematic (μVT) was used, where the system was under constant chemical potential, volume, and temperature. To eliminate periodic boundary conditions, we used supercell by 2 × 2 × 3 replicas of the unit cell for the calculations. The van der Waals interactions were truncated to a radius of 12 Å, and tail corrections were used to approximate the contributions beyond this truncation. In the simulations, 1 × 106 Monte Carlo steps were used for the equilibration and 1 × 107 Monte Carlo steps were used for production runs. The adsorbate molecules and the adsorbate framework were treated as rigid structures. CHF3 and CH2F2 were modeled as rigid tetrahedral molecules with five charged interaction sites. From previous simulation study, rigid body models were used to represent molecules as a collection of fixed geometric shapes that maintain a constant structure and orientation throughout the simulation. The host–guest and guest–guest interactions in the system were described by the short-range force and the electrostatic force, which are described by the Lennard-Jones and Coulomb potential functions (Equation (5)):
U i n t e r r i j = 4 ε i j σ i j r i j 12 σ i j r i j 6 + 1 4 π ε 0 q i q j r i j
where j σ i j and ε i j are the collision diameter and potential well depth, respectively, r i j is the distance between sites i and j , q i denotes the atomic charge on site i , and ε 0 is the permittivity of free space. The cross-interactions with other molecules and frameworks are obtained using the Lorentz–Berthelot mixing rule (as shown in Equations (6) and (7)):
σ i j = σ i i + σ j j 2
ε i j = ε i i · ε j j
A polarizable forcefield was employed to achieve an accurate description of the adsorption behavior of CHF3/CH2F2 in TIFSIX-2-Cu-i for molecular simulations. Back-polarization was neglected to achieve reasonable simulation times. To account for the implied polarization, we rescaled the Lennard-Jones [39] energy parameters according to the atomic polarizabilities. The Lennard-Jones energy parameters and charges of CHF3 and CH2F2 were taken from previous studies [40,41]. The Lennard-Jones parameters of TIFSIX-2-Cu-i for N, C, and H were taken from the OPLS-AA forcefield, and the rest of the atoms were taken from the UFF–Dreiding hybrid forcefield [42]. The equations used to adjust the parameters in this study are as follows:
ε i s c a l e d = ε i · 1 + λ α i α m a x 1 + λ α i α m a x · λ
αi means the polarizability of atomi, αmax means the max polarizability, and εi means the initial forcefield parameter. λ and ξ are scaling factors between 0 and 1, whose values depend on the discrepancy between the experimental data and the simulation results, used to rescale the Lennard-Jones energy parameters. The detailed methodology is described in Refs. [43,44]. In this study, by fitting the experimental data to the simulation results, λ was set to 0.2 and ξ to 0.9 for CH2F2, while λ was set to 0.9 and ξ to 0.01 for CHF3. Table 3 and Table 4 summarize all the forcefield parameters, atomic polarizabilities [45,46], and atomic charges.

4.4. Molecular Dynamics Simulations Details

In this paper, RASPA code was used to perform molecular dynamics simulations of CHF3/CH2F2 in TIFSIX-2-Cu-i. The polarized forcefield parameters from Table 3 and Table 4 were employed. All MD simulations were employed for 1 ns with a time step of 0.5 fs in the NVT ensemble to explore the diffusion of the equimolar CHF3/CH2F2 mixture in TIFSIX-2-Cu-i. The simulations were performed for 1 × 107 cycles, 2000 initialization cycles, and 20,000 equilibration cycles. We truncated the van der Waals interaction with a radius of 12.0 Å and used tail correction to approximate the contribution beyond this cutoff.

5. Conclusions

In this study, the adsorption mechanisms of CHF3 or CH2F2 in the TIFSIX-2-Cu-i framework were studied by combining single-component adsorption experiments and molecular simulations. In order to ensure consistency between the experimental data and the simulation results, a polarization forcefield was introduced. TIFSIX-2-Cu-i has excellent CH2F2 adsorption capacity (3.79 mmol/g) and CH2F2/CHF3 selectivity (3.17) at 3 bar and 308 K, making it a promising material to separate CHF3/CH2F2 mixtures. Regarding the competitive adsorption of CHF3–CH2F2 mixtures in TIFSIX-2-Cu-i, both the thermodynamic and kinetic selectivity of CH2F2 relative to CHF3 were observed to be relatively high. According to the combined effect of adsorption and diffusion, TIFSIX-2-Cu-i exhibits markedly preferential adsorption of CH2F2 for CHF3. The calculated heats of adsorption indicate relatively strong interactions between CH2F2 and TIFSIX-2-Cu-i. The snapshots show that CHF3/CH2F2 adsorption on TIFSIX-2-Cu-i involves multiple H⋯F interactions, where CHF3/CH2F2 interacts with TiF62− simultaneously. The typical binding sites of CH2F2 molecules in the TIFSIX-2-Cu-i channel are very similar to those of CHF3, and the F atoms in the TiF62− of the TIFSIX-2-Cu-i framework preferentially adsorb with H atoms of the CH2F2 molecule. CH2F2 has more H atoms, and TIFSIX-2-Cu-i shows stronger affinity for CH2F2 than CHF3.
In conclusion, this study reveals the adsorption mechanism of CHF3–CH2F2 mixtures in TIFSIX-2-Cu-i channels at the microscopic level. This research exploration provides an effective and superior strategy for the design and screening of highly selective adsorbents for the separation of CHF3–CH2F2 mixtures.

Author Contributions

Conceptualization, Q.F. and S.W.; methodology, S.W. and X.C.; software, L.Z. (Lei Zhou); validation, S.W., H.Q. and L.Z. (Lei Zhou); formal analysis, Z.D., H.L. and B.L.; investigation, S.W.; resources, L.Z. (Lei Zhou); data curation, S.W., Z.W. and L.Z. (Lina Zhang); writing—original draft preparation, S.W.; writing—review and editing, S.W. and Q.F.; visualization, S.W.; supervision, Q.F., X.C. and H.Q.; project administration, S.W.; funding acquisition, Q.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the Natural Science Foundation of Shandong Province in China (ZR2020MB121).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available upon request from the corresponding author.

Acknowledgments

The authors would like to thank all reviewers for their constructive advice.

Conflicts of Interest

Author Lei Zhou was employed by the company Shandong Dongyue Organosilicon Materials Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Sovacool, B.K.; Griffiths, S.; Kim, J.; Bazilian, M. Climate change and industrial F-gases: A critical and systematic review of developments, sociotechnical systems and policy options for reducing synthetic greenhouse gas emissions. Renew. Sustain. Energy Rev. 2021, 141, 110759. [Google Scholar] [CrossRef]
  2. Mota-Babiloni, A.; Navarro-Esbrí, J.; Barragán-Cervera, Á.; Molés, F.; Peris, B. Analysis based on EU Regulation No 517/2014 of new HFC/HFO mixtures as alternatives of high GWP refrigerants in refrigeration and HVAC systems. Int. J. Refrig. 2015, 52, 21–31. [Google Scholar] [CrossRef]
  3. Mota-Babiloni, A.; Makhnatch, P.; Khodabandeh, R. Recent investigations in HFCs substitution with lower GWP synthetic alternatives: Focus on energetic performance and environmental impact. Int. J. Refrig. 2017, 82, 288–301. [Google Scholar] [CrossRef]
  4. Velders, G.J.M.; Andersen, S.O.; Daniel, J.S.; Fahey, D.W.; McFarland, M. The importance of the Montreal Protocol in protecting climate. Proc. Natl. Acad. Sci. USA 2007, 104, 4814–4819. [Google Scholar] [CrossRef]
  5. Heath, E.A. Amendment to the Montreal Protocol on Substances that Deplete the Ozone Layer (Kigali Amendment). Int. Leg. Mater. 2017, 56, 193–205. [Google Scholar] [CrossRef]
  6. Benhadid-Dib, S.; Benzaoui, A. Refrigerants and their Environmental Impact Substitution of Hydro Chlorofluorocarbon HCFC and HFC Hydro Fluorocarbon. Search for an Adequate Refrigerant. Energy Procedia 2012, 18, 807–816. [Google Scholar] [CrossRef]
  7. Heredia-Aricapa, Y.; Belman-Flores, J.; Mota-Babiloni, A.; Serrano-Arellano, J.; García-Pabón, J.J. Overview of low GWP mixtures for the replacement of HFC refrigerants: R134a, R404A and R410A. Int. J. Refrig. 2020, 111, 113–123. [Google Scholar] [CrossRef]
  8. Hu, J.; Liu, C.; Liu, L.; Li, Q. Thermal energy storage of R1234yf, R1234ze, R134a and R32/MOF-74 nanofluids: A molecular simulation study. Materials 2018, 11, 1164. [Google Scholar] [CrossRef]
  9. Hu, G.; Wang, Z.; Zhang, W.; He, H.; Zhang, Y.; Deng, X.; Li, W. MIL-161 Metal–Organic Framework for Efficient Au(III) Recovery from Secondary Resources: Performance, Mechanism, and DFT Calculations. Molecules 2023, 28, 5459. [Google Scholar] [CrossRef]
  10. Apriliyanto, Y.B.; Faginas-Lago, N.; Evangelisti, S.; Bartolomei, M.; Leininger, T.; Pirani, F.; Pacifici, L.; Lombardi, A. Multilayer Graphtriyne Membranes for Separation and Storage of CO2: Molecular Dynamics Simulations of Post-Combustion Model Mixtures. Molecules 2022, 27, 5958. [Google Scholar] [CrossRef]
  11. Tagliabue, M.; Farrusseng, D.; Valencia, S.; Aguado, S.; Ravon, U.; Rizzo, C.; Corma, A.; Mirodatos, C. Natural gas treating by selective adsorption: Material science and chemical engineering interplay. Chem. Eng. J. 2009, 155, 553–566. [Google Scholar] [CrossRef]
  12. Wang, Q.; Yu, Y.; Li, Y.; Min, X.; Zhang, J.; Sun, T. Methane separation and capture from nitrogen rich gases by selective adsorption in microporous materials: A review. Sep. Purif. Technol. 2022, 283, 120206. [Google Scholar] [CrossRef]
  13. Yancey, A.D.; Terian, S.J.; Shaw, B.J.; Bish, T.M.; Corbin, D.R.; Shiflett, M.B. A review of fluorocarbon sorption on porous materials. Microporous Mesoporous Mater. 2022, 331, 111654. [Google Scholar] [CrossRef]
  14. Peng, J.; Fu, C.; Zhong, J.; Ye, B.; Xiao, J.; Duan, C.; Lv, D. Efficient Selective Capture of Carbon Dioxide from Nitrogen and Methane Using a Metal-Organic Framework-Based Nanotrap. Molecules 2023, 28, 7908. [Google Scholar] [CrossRef]
  15. Li, Y.; Bai, Y.; Wang, Z.; Gong, Q.; Li, M.; Bo, Y.; Xu, H.; Jiang, G.; Chi, K. Exquisitely Constructing a Robust MOF with Dual Pore Sizes for Efficient CO2 Capture. Molecules 2023, 28, 6276. [Google Scholar] [CrossRef] [PubMed]
  16. Xu, F.; Wu, Y.; Wu, J.; Lv, D.; Yan, J.; Wang, X.; Chen, X.; Liu, Z.; Peng, J. A Microporous Zn(bdc)(ted)0.5 with Super High Ethane Uptake for Efficient Selective Adsorption and Separation of Light Hydrocarbons. Molecules 2023, 28, 6000. [Google Scholar] [CrossRef] [PubMed]
  17. Ghazvini, M.F.; Vahedi, M.; Nobar, S.N.; Sabouri, F. Investigation of the MOF adsorbents and the gas adsorptive separation mechanisms. J. Environ. Chem. Eng. 2021, 9, 104790. [Google Scholar] [CrossRef]
  18. Wu, D.; Zhang, P.-F.; Yang, G.-P.; Hou, L.; Zhang, W.-Y.; Han, Y.-F.; Liu, P.; Wang, Y.-Y. Supramolecular control of MOF pore properties for the tailored guest adsorption/separation applications. Coord. Chem. Rev. 2021, 434, 213709. [Google Scholar] [CrossRef]
  19. Wanigarathna, D.K.; Gao, J.; Liu, B. Metal organic frameworks for adsorption-based separation of fluorocompounds: A review. Mater. Adv. 2020, 1, 310–320. [Google Scholar] [CrossRef]
  20. Jiang, M.; Cui, X.; Yang, L.; Yang, Q.; Zhang, Z.; Yang, Y.; Xing, H. A thermostable anion-pillared metal-organic framework for C2H2/C2H4 and C2H2/CO2 separations. Chem. Eng. J. 2018, 352, 803–810. [Google Scholar] [CrossRef]
  21. Yu, Y.; Yang, L.; Tan, B.; Hu, J.; Wang, Q.; Cui, X.; Xing, H. Remarkable separation of C5 olefins in anion-pillared hybrid porous materials. Nano Res. 2020, 14, 541–545. [Google Scholar] [CrossRef]
  22. Li, B.; Chen, B. Fine-Tuning Porous Metal-Organic Frameworks for Gas Separations at Will. Chem 2016, 1, 669–671. [Google Scholar] [CrossRef]
  23. Bajpai, A.; O’Nolan, D.; Madden, D.G.; Chen, K.-J.; Pham, T.; Kumar, A.; Lusi, M.; Perry, J.J.; Space, B.; Zaworotko, M.J. The effect of centred versus offset interpenetration on C2H2 sorption in hybrid ultramicroporous materials. Chem. Commun. 2017, 53, 11592–11595. [Google Scholar] [CrossRef]
  24. Zhang, Q.; Chen, X.; Fu, Q.; Qin, H.; Zhang, H.; Li, H.; Wang, S.; Dong, Z.; Zhang, Z.; Wang, M. Study on the Adsorption Separation of CHF3/N2 by an SIFSIX-3-Ni Metal–Organic Framework. Ind. Eng. Chem. Res. 2023, 62, 13627–13636. [Google Scholar] [CrossRef]
  25. Kołaczkiewicz, J.; Bauer, E. Clausius-Clapeyron equation analysis of two-dimensional vaporization. Surf. Sci. 1985, 155, 700–714. [Google Scholar] [CrossRef]
  26. Chen, K.-J.; Scott, H.S.; Madden, D.G.; Pham, T.; Kumar, A.; Bajpai, A.; Lusi, M.; Forrest, K.A.; Space, B.; Perry, J.J.; et al. Benchmark C2H2/CO2 and CO2/C2H2 Separation by Two Closely Related Hybrid Ultramicroporous Materials. Chem 2016, 1, 753–765. [Google Scholar] [CrossRef]
  27. Bokii, G. Kristallokhimiya, Izd; Nauka: Moskva, Russia, 1971. [Google Scholar]
  28. Zhou, L.; Zhou, Y.; Li, M.; Chen, P.; Wang, Y. Experimental and Modeling Study of the Adsorption of Supercritical Methane on a High Surface Activated Carbon. Langmuir 2000, 16, 5955–5959. [Google Scholar] [CrossRef]
  29. Nugent, P.; Rhodus, V.; Pham, T.; Tudor, B.; Forrest, K.; Wojtas, L.; Space, B.; Zaworotko, M. Enhancement of CO2 selectivity in a pillared pcu MOM platform through pillar substitution. Chem. Commun. 2013, 49, 1606–1608. [Google Scholar] [CrossRef]
  30. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  31. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  32. Limas, N.G.; Manz, T.A. Introducing DDEC6 atomic population analysis: Part 2. Computed results for a wide range of periodic and nonperiodic materials. RSC Adv. 2016, 6, 45727–45747. [Google Scholar] [CrossRef]
  33. Limas, N.G.; Manz, T.A. Introducing DDEC6 atomic population analysis: Part 4. Efficient parallel computation of net atomic charges, atomic spin moments, bond orders, and more. RSC Adv. 2018, 8, 2678–2707. [Google Scholar] [CrossRef] [PubMed]
  34. Hutter, J.; Iannuzzi, M.; Schiffmann, F.; VandeVondele, J. cp2k: Atomistic simulations of condensed matter systems. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2014, 4, 15–25. [Google Scholar] [CrossRef]
  35. VandeVondele, J.; Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007, 127, 114105. [Google Scholar] [CrossRef] [PubMed]
  36. Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703. [Google Scholar] [CrossRef] [PubMed]
  37. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. 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. [Google Scholar] [CrossRef] [PubMed]
  38. Dubbeldam, D.; Calero, S.; Ellis, D.E.; Snurr, R.Q. RASPA: Molecular simulation software for adsorption and diffusion in flexible nanoporous materials. Mol. Simul. 2016, 42, 81–101. [Google Scholar] [CrossRef]
  39. Schnabel, T.; Vrabec, J.; Hasse, H. Unlike Lennard–Jones parameters for vapor–liquid equilibria. J. Mol. Liq. 2007, 135, 170–178. [Google Scholar] [CrossRef]
  40. Fu, Q.; Tanaka, H.; Miyahara, M.T.; Qin, Y.; Shen, Y.; Zhang, D. CHF3–CHClF2 Binary Competitive Adsorption Equilibria in Graphitic Slit Pores: Monte Carlo Simulations and Breakthrough Curve Experiments. Ind. Eng. Chem. Res. 2018, 57, 6440–6450. [Google Scholar] [CrossRef]
  41. Yang, T.; Siepmann, J.I.; Wu, J. Phase Equilibria of Difluoromethane (R32), 1,1,1,2-Tetrafluoroethane (R134a), and trans-1,3,3,3-Tetrafluoro-1-propene (R1234ze(E)) Probed by Experimental Measurements and Monte Carlo Simulations. Ind. Eng. Chem. Res. 2020, 60, 739–752. [Google Scholar] [CrossRef]
  42. Rappe, A.K.; Casewit, C.J.; Colwell, K.S.; Goddard, W.A., III; Skiff, W.M. UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024–10035. [Google Scholar] [CrossRef]
  43. Becker, T.M.; Dubbeldam, D.; Lin, L.-C.; Vlugt, T.J.H. Investigating polarization effects of CO2 adsorption in MgMOF-74. J. Comput. Sci. 2016, 15, 86–94. [Google Scholar] [CrossRef]
  44. Becker, T.M.; Heinen, J.; Dubbeldam, D.; Lin, L.-C.; Vlugt, T.J.H. Polarizable Force Fields for CO2 and CH4 Adsorption in M-MOF-74. J. Phys. Chem. C 2017, 121, 4659–4673. [Google Scholar] [CrossRef] [PubMed]
  45. Shannon, R.D. Dielectric polarizabilities of ions in oxides and fluorides. J. Appl. Phys. 1993, 73, 348–366. [Google Scholar] [CrossRef]
  46. Van Duijnen, P.T.; Swart, M. Molecular and atomic polarizabilities: Thole’s model revisited. J. Phys. Chem. A 1998, 102, 2399–2407. [Google Scholar] [CrossRef]
Figure 1. The experimental and simulated adsorption isotherms for pure CHF3 (a) or CH2F2 (b) at 288, 298, and 308 K in TIFSIX-2-Cu-i.
Figure 1. The experimental and simulated adsorption isotherms for pure CHF3 (a) or CH2F2 (b) at 288, 298, and 308 K in TIFSIX-2-Cu-i.
Molecules 29 01721 g001
Figure 2. The adsorption separation selectivity for CH2F2 over CHF3 in TIFSIX-2-Cu-i at different temperatures and pressures, which correspond to the CH2F2-CHF3 mixture (CH2F2/CHF3, 50/50, v/v).
Figure 2. The adsorption separation selectivity for CH2F2 over CHF3 in TIFSIX-2-Cu-i at different temperatures and pressures, which correspond to the CH2F2-CHF3 mixture (CH2F2/CHF3, 50/50, v/v).
Molecules 29 01721 g002
Figure 3. The isosteric adsorption heat (Qst) of CH2F2/CHF3 on the TIFSIX-2-Cu-i.
Figure 3. The isosteric adsorption heat (Qst) of CH2F2/CHF3 on the TIFSIX-2-Cu-i.
Molecules 29 01721 g003
Figure 4. Mean-square displacements for CHF3–CH2F2 mixture (CHF3/CH2F2, 50/50, v/v) in TIFSIX-2-Cu-i at 288 K (a), 298 K (b), and 308 K (c).
Figure 4. Mean-square displacements for CHF3–CH2F2 mixture (CHF3/CH2F2, 50/50, v/v) in TIFSIX-2-Cu-i at 288 K (a), 298 K (b), and 308 K (c).
Molecules 29 01721 g004
Figure 5. The typical binding sites for CH2F2 (a) or CHF3 (b) in TIFSIX-2-Cu-i. The cyan, gray, and white spheres represent fluorine, carbon, and hydrogen atoms, respectively.
Figure 5. The typical binding sites for CH2F2 (a) or CHF3 (b) in TIFSIX-2-Cu-i. The cyan, gray, and white spheres represent fluorine, carbon, and hydrogen atoms, respectively.
Molecules 29 01721 g005
Figure 6. The RDF between the framework and each atom of CHF3 (a)/CH2F2 (b); the RDF between the representative atoms on the framework and hydrogen atom of CHF3 (c)/CH2F2 (d) in 298 K.
Figure 6. The RDF between the framework and each atom of CHF3 (a)/CH2F2 (b); the RDF between the representative atoms on the framework and hydrogen atom of CHF3 (c)/CH2F2 (d) in 298 K.
Molecules 29 01721 g006
Figure 7. The redistribution of charge density in TIFSIX-2-Cu-i after adsorbing CHF3 molecules (a) or CH2F2 molecules (b). Color code: brown, C; meat pink, H; silver, F; blue, Ti.
Figure 7. The redistribution of charge density in TIFSIX-2-Cu-i after adsorbing CHF3 molecules (a) or CH2F2 molecules (b). Color code: brown, C; meat pink, H; silver, F; blue, Ti.
Molecules 29 01721 g007
Figure 8. The slices of charge density redistribution for TIFSIX-2-Cu-i and CHF3 (a)/CH2F2 (b) after molecular adsorption, which correspond to the electron transfer between hydrogen atom of CHF3/CH2F2 and the fluorine atom of TIF62−.
Figure 8. The slices of charge density redistribution for TIFSIX-2-Cu-i and CHF3 (a)/CH2F2 (b) after molecular adsorption, which correspond to the electron transfer between hydrogen atom of CHF3/CH2F2 and the fluorine atom of TIF62−.
Molecules 29 01721 g008
Figure 9. XRD patterns of TIFSIX-2-Cu-i after activation (compared to calculated patterns).
Figure 9. XRD patterns of TIFSIX-2-Cu-i after activation (compared to calculated patterns).
Molecules 29 01721 g009
Figure 10. Diagram of adsorption measurements’ experimental apparatus.
Figure 10. Diagram of adsorption measurements’ experimental apparatus.
Molecules 29 01721 g010
Figure 11. The chemically different atoms in TIFSIX-2-Cu-i. Atom colors: C = gray; H = white; N = blue; F = green; Ti = silver; Cu = red.
Figure 11. The chemically different atoms in TIFSIX-2-Cu-i. Atom colors: C = gray; H = white; N = blue; F = green; Ti = silver; Cu = red.
Molecules 29 01721 g011
Table 1. Parameters of CHF3 and CH2F2 fitted by the Langmuir Adsorption Isotherm Model.
Table 1. Parameters of CHF3 and CH2F2 fitted by the Langmuir Adsorption Isotherm Model.
T (K)qm (mmol/g)b (bar−1)R-Squared
2886.224170.855260.9582
CH2F22987.063670.572360.97048
3086.977560.442330.9712
2884.925550.518970.976526
CHF32984.227340.471270.96737
3083.86380.381440.9911
Table 2. Pore parameters of TIFSIX-2-Cu-i.
Table 2. Pore parameters of TIFSIX-2-Cu-i.
SBET (m2/g)Sample Density (cm3/g)Total Pore Volume (cm3/g)t-Plot Micropore Volume (cm3/g)Mesopore Volume (cm3/g)Average Pore Size (Å)
3721.2110.2110.1310.0793.988
Table 3. Partial charges, forcefield parameters, and atomic polarizabilities corresponding to the atom types in CHF3 and CH2F2.
Table 3. Partial charges, forcefield parameters, and atomic polarizabilities corresponding to the atom types in CHF3 and CH2F2.
Atom Typesq (|e|)σ (Å)ε (K)α3)
C_CHF30.7193.52551.2886
H_CHF30.0162.625.20.4138
F_CHF3−0.2452.92250.44475
C_CH2F20.3853.46421.2886
H_CH2F20.0492.2290.4138
F_CH2F2−0.24152.95370.44475
Table 4. Forcefield parameters and the atomic polarizabilities corresponding to the atom types in the TIFSIX-2-Cu-i.
Table 4. Forcefield parameters and the atomic polarizabilities corresponding to the atom types in the TIFSIX-2-Cu-i.
Atom Typesσ (Å)ε (K)α (Å3)
N3.2585.54790.97157
C3.540.2581.2886
H2.4215.0970.4138
F3.093236.48340.44475
Cu3.1142.512.1963
Ti2.82868.554733.2428
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, S.; Zhou, L.; Qin, H.; Dong, Z.; Li, H.; Liu, B.; Wang, Z.; Zhang, L.; Fu, Q.; Chen, X. Study of CHF3/CH2F2 Adsorption Separation in TIFSIX-2-Cu-i. Molecules 2024, 29, 1721. https://doi.org/10.3390/molecules29081721

AMA Style

Wang S, Zhou L, Qin H, Dong Z, Li H, Liu B, Wang Z, Zhang L, Fu Q, Chen X. Study of CHF3/CH2F2 Adsorption Separation in TIFSIX-2-Cu-i. Molecules. 2024; 29(8):1721. https://doi.org/10.3390/molecules29081721

Chicago/Turabian Style

Wang, Shoudong, Lei Zhou, Hongyun Qin, Zixu Dong, Haoyuan Li, Bo Liu, Zhilu Wang, Lina Zhang, Qiang Fu, and Xia Chen. 2024. "Study of CHF3/CH2F2 Adsorption Separation in TIFSIX-2-Cu-i" Molecules 29, no. 8: 1721. https://doi.org/10.3390/molecules29081721

Article Metrics

Back to TopTop