Next Article in Journal
Are There Lipid Membrane-Domain Subtypes in Neurons with Different Roles in Calcium Signaling?
Next Article in Special Issue
Simulating Crystal Structure, Acidity, Proton Distribution, and IR Spectra of Acid Zeolite HSAPO-34: A High Accuracy Study
Previous Article in Journal
Determination of the Concentration of Heavy Metals in Artisanal Cheeses Produced in the Mexican States of Tabasco and Chiapas
Previous Article in Special Issue
Exquisitely Constructing a Robust MOF with Dual Pore Sizes for Efficient CO2 Capture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Selective Capture of Carbon Dioxide from Nitrogen and Methane Using a Metal-Organic Framework-Based Nanotrap

1
School of Environmental and Chemical Engineering, Foshan University, Foshan 528000, China
2
GAC R&D Center, Guangzhou Automobile Group Co., Ltd., 668 Jinshan Road East, Panyu District, Guangzhou 511434, China
3
Appraisal Center for Environment and Engineering, Ministry of Ecology and Environment, Beijing 100012, China
4
School of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510640, China
5
School of Materials Science and Hydrogen Engineering, Foshan University, Foshan 528000, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(23), 7908; https://doi.org/10.3390/molecules28237908
Submission received: 16 November 2023 / Revised: 30 November 2023 / Accepted: 30 November 2023 / Published: 2 December 2023
(This article belongs to the Special Issue Zeolites and Porous Materials: Synthesis, Properties and Applications)

Abstract

:
Selective carbon capture from exhaust gas and biogas, which mainly involves the separation of CO2/N2 and CO2/CH4 mixtures, is of paramount importance for environmental and industrial requirements. Herein, we propose an interesting metal-organic framework-based nanotrap, namely ZnAtzCO3 (Atz = 3-amino-1,2,4-triazolate, CO32− = carbonate), with a favorable ultramicroporous structure and electrostatic interactions that facilitate efficient capture of CO2. The structural composition and stability were verified by FTIR, TGA, and PXRD techniques. Particularly, ZnAtzCO3 demonstrated high CO2 capacity in a wide range of pressures, with values of 44.8 cm3/g at the typical CO2 fraction of the flue gas (15 kPa) and 56.0 cm3/g at the CO2 fraction of the biogas (50 kPa). Moreover, ultrahigh selectivities over CO2/N2 (15:85, v:v) and CO2/CH4 (50:50, v:v) of 3538 and 151 were achieved, respectively. Molecular simulations suggest that the carbon atom of CO2 can form strong electrostatic Cδ+···δ−O-C interactions with four oxygen atoms in the carbonate ligands, while the oxygen atom of CO2 can interact with the hydrogen atoms in the triazolate ligands through Oδ−···δ+H-C interactions, which makes ZnAtzCO3 an optimal nanotrap for CO2 fixation. Furthermore, breakthrough experiments confirmed excellent real-world separation toward CO2/N2 and CO2/CH4 mixtures on ZnAtzCO3, demonstrating its great potential for selective CO2 capture.

Graphical Abstract

1. Introduction

Global warming, one of the biggest global issues, is causing various long-term disastrous environmental effects, including abnormal climate patterns, rising sea levels, accelerated extinction of species, and shifts in agricultural patterns, which pose severe threats to the survival and development of humanity [1,2,3,4]. As the main source of greenhouse gases, carbon dioxide (CO2), with a large annual emission (e.g., 36.8 Gt in 2022), contributes to 69.4% of the anthropogenic greenhouse gases [5,6]. Hence, CO2 capture and utilization is of utmost importance to curb global warming [7]. In particular, electricity generation from the combustion of fossil fuels, resulting in exhaust gas composed mainly of CO2 and nitrogen (N2), is the primary source of anthropogenic CO2 emissions [8,9]. Additionally, biogas, acknowledged as “renewable natural gas”, is a green fuel with the efficient component of methane (CH4) [10]. Nevertheless, biogas contains a considerable amount of CO2 that could significantly decrease the calorific value and lead to severe erosion in the equipment. Therefore, efficient selective capture of CO2 from N2 and CH4 is of paramount significance for environmental protection and biogas upgrading.
Benefiting from a plausible gentle operation condition and reduced energy consumption for regeneration, adsorptive separation of CO2 has been recognized as a promising alternative among the current CO2 capture techniques [11,12,13,14]. Fundamentally, an ideal adsorbent with high capacity and selectivity for the target molecule is the key to achieving satisfactory separation performance [15,16]. As a class of innovative porous adsorbents, metal-organic frameworks (MOFs) are showing extraordinary versatility in pore tunability and chemical functionalities [17]. Hence, adsorptive separation based on MOFs has been explored in miscellaneous separation circumstances, including carbon capture [8,9,11,18,19,20]. In general, there are several strategies that can enhance the CO2 capture performance on MOF materials, including (i) incorporation of functionalities (such as open metal sites, Lewis basic sites, and polar functional groups) into the frameworks, (ii) utilization of ultramicropores to maximize the confinement effect, (iii) use of kinetic difference and even the molecular sieving effect, (iv) introduction of structural flexibility, (v) the combination thereof [16,21,22]. Specifically, MOFs constructed by triazolate linkers represent one subclass of MOFs that are comprised of excellent CO2 capture performance, affordable costs, and good stability under humid conditions [23,24]. For example, by introducing amino groups into the triazolate linkers of the prototype MOF ZnF(TZ), the resultant ZnF(daTZ) demonstrates an appreciable volumetric CO2 uptake of 75 cm3/cm3 and high CO2/N2 equilibrium selectivity of 120 based on the ideal adsorbed solution theory (IAST) as well as excellent CO2/H2O kinetic selectivity of 70 [25]. This remarkable separation performance was probably originated from the main adsorptive site at the channel center for CO2 to afford electrostatic interactions with the amino groups, while H2O was more likely to locate at the channel corner, according to GCMC simulations. Built from dual ligands of oxalate and 1,2,4-triazolate, zinc-based Calgary Framework 20 (CALF-20) exhibits a high CO2 capacity of 87.36 cm3/g at atmospheric conditions, excellent selectivities over N2 and H2O, and the facility for scalable production, making it the first practical MOF for industrial carbon capture [26,27]. Molecular simulations suggest the CO2 adsorption location in CALF-20 also lies in the channel center to form interactions with the zinc/nitrogen/carbon atoms in the framework. Likewise, the amine-appended zinc-oxalate-triazolate MOFs demonstrated enhanced CO2 capacity at low CO2 concentrations compared to CALF-20, due to higher-density interaction sites and more contracted pore sizes [28,29]. Recently, we prepared a flexible MOF, namely ZnDatzBdc, that showed step-shaped CO2 isotherm due to breakage/reformation of intra-framework hydrogen bonds and rotation of the phenyl rings, giving rise to an excellent CO2 theoretical working capacity of 94.9 cm3/cm3 if performed in typical pressure vacuum swing adsorption at 273 K [30].
By now, developing nanotraps with multiple host-guest interactions toward the target molecules offers a feasible strategy to accomplish high adsorption capacity and selectivity, which has been successfully applied in separation circumstances, such as C2H2/CO2 separation [31,32], C3H4/C3H6 separation [33,34], CH4/N2 separation [35], and olefin/paraffin separation [36]. For CO2 adsorption, if a contracted pore exhibits opposite electrostatics on the adjacent positions and the same electrostatics on its opposite side, it can form strong electrostatic interactions with both the carbon and oxygen atoms in the CO2 molecule, as shown in Figure 1. Hence, this type of pore can act as a suitable nanotrap for CO2 fixation, from which a remarkable capacity and selectivity for CO2 can be achieved.
Herein, we synthesized a novel MOF, namely ZnAtzCO3, with 3-amino-1,2,4-triazolate (Atz-) and carbonate (CO32−) as dual ligands. Inspiringly, ZnAtzCO3 shows the desired ultramicropores due to the small-sized ligands and suitable crystal structure, which is desirable for the adsorption and separation of small molecules, such as CO2. Moreover, the favorable electrostatic environment of ZnAtzCO3 makes it a feasible nanotrap to form multiple host-guest interactions with CO2, and hence, efficient CO2 capture from N2 and CH4 could be achieved. In particular, at atmospheric temperature, equilibrium isotherms showed high CO2 capacities of ZnAtzCO3 in a wide pressure range, with values of 44.8 cm3/g (STP, standard temperature and pressure) at the typical fraction of the flue gas (15 kPa) and 56.0 cm3/g at the fraction of the biogas (50 kPa). Moreover, adsorptive selectivity based on the IAST model indicated that ultra-high CO2/N2 and CO2/CH4 selectivities of 3538 and 151 were realized at ambient conditions, respectively. The excellence in capacity and selectivity of this MOF-based nanotrap was illustrated by molecular simulations in terms of preferential adsorption sites, binding energy, and adsorption distributions. Furthermore, breakthrough experiments toward the binary mixtures of CO2/N2 and CO2/CH4 were conducted on ZnAtzCO3, which verified its efficient dynamic separation performance.

2. Results and Discussion

2.1. Crystal Structure and Pore Properties

Reactions of ZnSO4 and 3-amino-1H-1,2,4-triazole (HAtz) in a binary solution of DMF/H2O afforded high-quality crystals of ZnAtzCO3. The single-crystal X-ray diffraction (SCXRD) measurement indicates that ZnAtzCO3 belongs to the triclinic crystal system (a = 9.6217 Å, b = 9.6316 Å, c = 16.3408 Å, α = 81.355°, β = 86.938°, γ = 76.093°). Each asymmetric unit contains four zinc atoms, four 3-aminotriazolate ligands, and two triangular carbonate linkers (Figure S1). Because no carbonate was added to the reactants, it is assumed that the carbonate linker originated from the decomposition of the DMF molecule [37]. In addition, the elemental analysis suggests no sulfur element in the framework, which further confirms the existence of the carbonate linker instead of the sulfate or sulfite linker in the framework. As shown in Figure 2a, the zinc atom coordinates with one oxygen atom from the carbonate linker and three nitrogen atoms from three different triazolate rings. Each carbonate linker contains one uncoordinated oxygen atom, which is fixed by the intra-framework hydrogen bonding with the amino group in the triazolate ligand. We tried to obtain isoreticular structures by replacing the HAtz ligand with 1H-1,2,4-triazole (HTz) and 3,5-diamine-1H-1,2,4-triazole (HDatz), but the trial failed. Hence, we assume that these weak intra-framework hydrogen bonds are vital for the structural formation of ZnAtzCO3. ZnAtzCO3 can be regarded as a pillared–layered structure by connecting the wavy and continuous zinc-triazolate layers by the carbonate pillars (Figure 2b). Inspiringly, the small size of the carbonate and triazolate ligands gives rise to an ultramicroporous structure desirable for adsorptive separation. Specifically, ZnAtzCO3 contains two types of zig-zag channels with a cross-section area of 2.9 × 5.1 Å2 and 3.5 × 5.1 Å2 that interconnect with the adjacent channels through small apertures of 3.0 × 3.9 Å2, 2.2 × 2.8 Å2, and 2.6 × 3.3 Å2. We noticed that a similar form of Zn2(atz)2(CO3) was previously reported by the reaction of Zn(NO3)2, NaHCO3, and HAtz. Zn2(atz)2(CO3) displayed the same connectivity but belonged to another different space group of a Pnma unit cell (a = 9.806 Å, b = 9.3353 Å, c = 16.194 Å, α = β = γ = 90°) [38]. The difference in structure is probably because the degrees of buckling for zinc-triazolate layers can vary significantly under different synthetic conditions. Because the specific crystal structure affects the pore systems and, subsequently, the sorption behavior, the following discussion was carried out with our obtained crystal data.

2.2. Characterizations

The physicochemical behavior of ZnAtzCO3 was measured to investigate its textural characteristics. Fourier transform infrared reflection (FTIR) patterns were performed to further confirm the existence of the carbonate linker. Figure 3a suggests an intense broad band at 1408 cm−1 and an additional band at 1310 cm−1, corresponding to the asymmetric stretching modes of carbonate [39]. Besides, the medium band at 850 cm−1 was assigned to the bending mode of the carbonate. Figure 3b depicts the powder X-ray diffraction (PXRD) pattern of ZnAtzCO3, which is identical to that derived from SCXRD measurement, indicative of the high purity of the powder sample. In addition, the activation step did not lead to transformation or decomposition of the structure, as suggested by the well-maintained PXRD patterns. Thermogravimetric (TG) analysis in Figure 3c indicates that ZnAtzCO3 is stable up to 500 K, and hence, it holds enough thermal stability for adsorptive separation, which usually requires moderate heating for regeneration. The porosity feature was derived from CO2 sorption isotherms at 195 K, as shown in Figure 3d. The typical type-I CO2 isotherms indicated the microporous nature of ZnAtzCO3, giving a Brunauer–Emmett–Teller (BET) surface area of 455.6 m2/g and a micropore volume of 0.196 cm3/g.

2.3. Adsorption Equilibrium Behavior of CO2, N2, and CH4

Figure 4a depicts the pure adsorption isotherms of CO2, N2, and CH4 on ZnAtzCO3 at 298 K. It is noticed that CO2 capacity increased sharply at low pressure, while N2 and CH4 uptakes showed a slow increment as the pressure rose. Hence, ZnAtzCO3 showed a significantly higher CO2 capacity than N2 and CH4 between 0–100 kPa. At atmospheric pressure, CO2 capacity reached as high as 62.8 cm3/g, while this value for N2 and CH4 was 4.3 and 14.7 cm3/g, respectively. This distinction in adsorption capacity suggests excellent thermodynamic separation for CO2/N2 and CO2/CH4. Moreover, ZnAtzCO3 demonstrated an exceptional CO2 adsorption capacity of 44.8 cm3/g at 15 kPa, highlighting its promising application prospects in low-concentration CO2 capture, such as CO2 elimination from the exhaust gas. Moreover, ZnAtzCO3 exhibits a high CO2 capacity of 56.0 cm3/g at 50 kPa, indicative of its potential in CO2 separation with higher CO2 concentrations, including CO2 removal from the biogas.
To evaluate the competitive separation of CO2/CH4 and CO2/N2 on ZnAtzCO3, the IAST selectivities were calculated by means of the IAST model, taking into account the composition of CO2/N2 (15:85, v:v) and CO2/CH4 (50:50,v:v) in the exhaust gas and biogas, respectively [40,41]. By incorporating the dual-site Langmuir–Freundlich (DSLF) parameter (Table S1) into the IAST model [42], the IAST selectivities were obtained and shown in Figure 4b. Significantly, CO2/CH4 and CO2/N2 selectivities are quite high in the whole pressure range of 0–100 kPa. At ambient pressure, the IAST selectivity for CO2/CH4 and CO2/N2 reached as high as 3538 and 151, respectively, which are comparable to those benchmark MOFs for selective CO2 capture through thermodynamic separation [12]. Hence, compromised of excellent CO2 capacity and selectivity, ZnAtzCO3 reveals great potential for selective CO2 capture.
Furthermore, the isosteric heat (Qst) of CO2, N2, and CH4 on ZnAtzCO3, which can be derived from their pure adsorption isotherms at various temperatures (273, 288, and 298 K), is a crucial parameter for determining the adsorption interaction strengths. As shown in Figure 5b, the zero-coverage Qst for the three gases followed an order of CO2 (32.6 kJ/mol) > CH4 (22.4 kJ/mol) > N2 (18.1 kJ/mol), consistent with the dipole moment of the guest molecules (CO2: 29.1 × 10−25 cm3, CH4: 25.9 × 10−25 cm3, N2: 17.4 × 10−25 cm3) [43]. Additionally, it is apparent that the Qst for all gases remained constant between 0–100 kPa, suggesting homogeneity on the pore surface. From the Qst result, we speculate that the remarkable CO2 selectivity is primarily attributed to its highest adsorption enthalpy.
In addition, ZnAtzCO3 was compared with other MOFs constructed by triazolate linkers on their CO2 uptake at 15 kPa and 100 kPa, together with the isosteric heat [25,26,29,30,44,45,46,47]. As shown in Table 1, the CO2 uptakes on ZnAtzCO3 exceed those of the ZnF(Tz) series, ZnDatzBdc, Zn(FA)(datrz)2, and Zn2(TRZ)2(BDC), and are comparable to that of ZU-301, and slightly lower than those of ZnAtzOx and CALF-20. Hence, ZnAtzCO3 holds comparatively high CO2 capacity among these triazolate-based MOFs. Hence, ZnAtzCO3 can be regarded as a promising adsorbent in combination with good capacity and selectivity toward CO2.

2.4. Molecular Simulations on the Selective CO2 Adsorption over N2 and CH4

The intrinsic mechanism for the excellent separation performance on ZnAtzCO3 was illustrated by molecular simulations on the preferential adsorption sites, adsorption density distributions, and interaction energy with the aid of Material Studio 7.0 [48]. To visualize the intrinsic host–guest interactions between ZnAtzCO3 and the gas molecules, the preferential interaction site was calculated and depicted in Figure 6. Specifically, for the CO2 molecule, the carbon atom can form four Cδ+···δ−O-C electrostatic interactions with the oxygen atoms in the carbonate linker, and each oxygen atom can form Oδ−···δ+H-C electrostatic interactions with the hydrogen atoms in the aminotriazolate rings. The multiple host–guest interactions verify that the favorable electrostatic environment of ZnAtzCO3 can form an efficient nanotrap for CO2. It is noticed that the amine group did not form electrostatic interactions with CO2, which might originate from insufficient contact with CO2. For CH4, CH4 interacts with three oxygen atoms in the carbonate linker, two nitrogen atoms in the amino groups, and one adjacent triazolate ring through dispersion forces. Considering its significantly smaller polarizability and more inert nature, the preferential site displayed weaker affinity for CH4 than CO2. Likewise, being the weakest adsorbate, N2 forms dispersion interactions with three hydrogen atoms in the aminotriazolate linkers and one oxygen atom in the carbonate linker. From the result above, ZnAtzCO3 shows stronger host–guest interactions with CO2 compared to N2 and CH4.
Figure 7 presents the ambient-temperature adsorption density distributions of the three gases on ZnAtzCO3 at low (15 kPa) and ambient pressure (100 kPa). At both 15 kPa and 100 kPa, the density distribution for CO2 was the highest, followed by CH4 and N2, which is a valid proof of the significantly higher capacity of CO2 than N2 and CH4. As the pressure rose from 15 kPa to 100 kPa, the adsorption density increased for all gases because, generally, the increment in pressure can provide an increasing driving force for gas adsorption.
In addition, the distinction of the simulated interaction energy was calculated to further confirm the difference in the adsorption enthalpy. As shown in Figure 8, the average energies between ZnAtzCO3 and the gas molecules were −6.85, −6.06, and −3.50 kcal/mol for CO2, CH4, and N2, respectively, which shows a consistent trend with the Qst result. Hence, the stronger electrostatic host–guest interactions, apparently higher adsorption density distributions, and larger adsorption energy, comprehensively explain the selective CO2 adsorption over N2 and CH4 on ZnAtzCO3.

2.5. Dynamic Breakthrough Experiments

For evaluation of the capability for selective CO2 capture from the exhaust gas and biogas on ZnAtzCO3, the breakthrough experiments were performed to simulate the dynamic separation performance toward CO2/N2 (15:85, v:v) and CO2/CH4 (50:50, v:v) at ambient conditions (Figure 9). It is noticed that CH4 and N2 were detected shortly after induction of the gas mixtures and reached equilibrium rapidly, confirming their uptake was inappreciable on ZnAtzCO3. In contrast, CO2 broke through at 81 min/g and reached equilibrium at 134 min/g for the CO2/N2 mixture, and the breakthrough time and equilibrium time for CO2/CH4 mixture were 37 and 53 min/g, respectively. In addition, the CO2 breakthrough time remained constant for five cycles for both mixtures, suggesting their excellent cyclicity in real-world CO2 capture conditions.

3. Materials and Methods

3.1. Material Sources

Zinc sulfate heptahydrate (ZnSO4·7H2O, 99.9%), 3-amino-1H-1,2,4-triazole (HAtz, 99.9%), N,N’-dimethylformamide (DMF, 99.5%), and methanol (MeOH, 99.5%) were supplied by Macklin Biochemical Co., Ltd. (Shanghai, China). The gases employed in this work, including single-component gases (CO2, CH4, N2, and He) of high purity (over 99.9%) and binary gas mixtures of CO2/N2 (15:85, v:v) and CO2/CH4 (50:50, v:v), were purchased from Kedi Gas Chemical Industry Co., Ltd. (Foshan, China).

3.2. Synthesis of ZnAtzCO3

HAtz (84.0 mg, 1 mmol) was dispersed in DMF (2 mL) in an autoclave, followed by addition of an aqueous solution (8 mL) of dissolved ZnSO4·7H2O (287.6 mg, 1 mmol) under stirring. After the suspension was mixed by stirring for 15 min, the autoclave was sealed, kept at 423 K for 60 h, and subsequently slowly cooled to ambient temperature. The resultant colorless crystals were recovered by filtration, washed with deionized H2O (2 × 20 mL) and MeOH (3 × 20 mL) to remove the unreacted reactants, and then dried in air. The activated ZnAtzCO3 was prepared by heating to remove the contaminants such as DMF, H2O, and the gases adsorbed from air. CCDC number: 2297804. Crystal Data for Zn4C10H12N16O6 (M = 713.92 g/mol): triclincic, space group P-1 (no. 2), a = 9.6217(2) Å, b = 9.6316(2) Å, c = 16.3408(4) Å, α = 81.355(2) °, β = 86.938(2)°, γ = 76.093(2)°, V = 1453.01(6) Å3, Z = 2, T = 150.00(10) K, μ(CuKα) = 0.790 mm−1, Dcalc = 1.632 g/cm3, 5103 reflections measured (4.735° ≤ 2Θ ≤ 66.988°), 4762 unique (Rint = 0.0381, Rsigma = 0.0300) which were used in all calculations. The final R1 was 0.0623 (I > 2σ(I)) and wR2 was 0.1667 (all data). Elemental analysis (wt%) calculated for Zn4(Atz)4(CO3)2 (Zn4C10H12N16O6): C, 16.83; H, 1.69; N, 31.40; S: 0.00. Found: C, 17.25; H, 2.05; N, 30.5; S, 0.00.

3.3. Characterizations

SCXRD analysis was carried out on a Rigaku Oxford Diffraction (Rigaku, Tokyo, Japan) with a hybrid pixel array detector. Reflections combined with SHELXL corresponding to the crystal class were employed for the calculation of statistics and refinement to solve the non-hydrogen atoms, while the locations and numbers of all hydrogen atoms were calculated theoretically. Elemental analysis was performed on a Vario EL elemental analyzer (Elementar, Langenselbold, Germany) in the CHNS mode. FTIR spectroscopy in the range of 1800–400 cm−1 was recorded on a Thermo Scientific iN10 (Thermo Fisher Scientific, Waltham, MA, USA) microscope with potassium bromide as the matrix. PXRD patterns were collected on a Bruker D8 Advance diffractometer (Bruker, Mannheim, Germany). TG measurements of the as-synthesized ZnAtzCO3 were performed on a TGA 550 thermal gravimetric analyzer (Thermo Fisher Scientific, Waltham, MA, USA), and the sample was heated from 303 K to 973 K at a ramping rate of 10 K/min under flowing nitrogen.

3.4. Single-Component Gas Sorption Isotherm Measurements

Single-component sorption isotherm measurements between 0–100 kPa were carried out on 3Flex (Micromeritics, Norcross, GA, USA) at various temperatures. In the preparation process, approximately 100 mg of ZnAtzCO3 was activated under dynamic vacuum for 6 h at 393 K to afford a guest-free sample. During the test, the sample tube was placed in a thermostatic environment by using the ice-acetone bath (195 K) or circulating water bath (288 K, 298 K, and 313 K) to maintain a constant operational temperature.

3.5. Adsorption Selectivity Based on IAST Model

Before the IAST selectivity calculation, the experimental isotherms of CO2, N2, and CH4 require accurate fitting to a mathematical model. In this work, the DSLF equation, based on the assumption that two types of adsorption sites are present in the structure, was selected and described to describe the adsorption equilibrium of the single-component gases Equation (1) [30,42].
q = q e 1 k 1 p t 1 1 + k 1 p t 1 + q e 2 k 2 p t 2 1 + k 2 p t 2
where p is the specific pressure when the gas phase and adsorbed phase reach a steady state; qei is the saturated uptake of site i; ki is the affinity coefficients of site i; ti represents the divergence from an absolute homogeneous surface on site i.
The IAST selectivity for CO2/N2 and CO2/CH4 on ZnAtzCO3 can be derived according to Equation (2) [40].
S A B = x A / x B y A / y B
where xi and yi refer to the volume fractions of component i in the adsorbed phase and the gas phase, separately.

3.6. Isosteric Heat (Qst) Calculation

The experimental adsorption isotherms of CO2, N2, and CH4 on ZnAtzCO3 at various temperatures were fitted to the Virial equation, Equation (3) [49], and the parameters are shown in Table S2.
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, N is the gas capacity, T is the absolute temperature, ai and bj refer to the corresponding parameter in the Virial equation, while m and n refer to the number required for the accurate fitting of the Virial equation.
Subsequently, the corresponding Qst was figured out by substituting the parameter in the Virial equation into the Clausius–Clapeyron equation, Equation (4) [50].
Q s t = R i = 0 m a i N i
where R is short for the ideal gas constant, 8.314 J/mol/K.

3.7. Breakthrough Experiments

Dynamic separation experiments of CO2/N2 (15:85, v:v) and CO2/CH4 (50:50, v:v) mixtures were performed on self-assembly breakthrough equipment (Figure S3). A small-scale adsorption column was prepared by loading approximately 600 mg of the activated ZnAtzCO3 in a stainless-steel column (Φ 50 × 150 mm). For activation, the column packed with the sample was heated at 393 K for two hours to eliminate the adsorbed contaminants. Subsequently, the column was inserted into the breakthrough equipment and purged by He flow (10 mL/min) at ambient conditions until the baseline was flattened. Finally, the gas was shifted to CO2/N2 or CO2/CH4 at a flow rate of 3 mL/min. The outlet component was monitored on a thermal conductivity detector (TCD) until the outlet composition reached that of the feed gas, which suggested the breakthrough column reached equilibrium. The adsorption column was recovered by purging He flow at 373 K to liberate the adsorbed gas molecules in the cyclability test.

3.8. Simulation Details

The molecular simulations on the adsorption mechanism were calculated by utilizing the Materials Studio 7.0 software [50]. First, the structures of ZnAtzCO3 and the adsorbates were optimized with the aid of the Forcite and Dmol3 modules. The adsorption characteristics, including the optimal adsorption sites, adsorption density distribution, and stabilized adsorption energy, were simulated in the Sorption module with the Metropolis Monte Carlo method. The adsorption behavior of the guest molecules on ZnAtzCO3 was described by several motion types, including exchange, conformation, rotation, translation, and regeneration. The Ewald and atom-based methods were employed to depict the electrostatic interaction and Van der Waals interactions between the structure and the guest molecules, respectively. The cutoff for the Metropolis Monte Carlo simulation was set as 12.5 Å. One gas molecule was randomly inserted into the framework in the Location task in the Sorption module, with 1 × 105 steps for equilibrium and production, separately.

4. Conclusions

In conclusion, we propose an interesting type of MOF-based nanotrap, namely ZnAtzCO3, for efficient selective capture of CO2 from N2 and CH4. The favorable electrostatic environment and narrow pore geometry of ZnAtzCO3 show stronger interaction with CO2 than N2 and CH4. Specifically, ZnAtzCO3 accomplished high CO2 capacities with values of 74.0 cm3/cm3 at the fraction of the flue gas (15 kPa) and 91.4 cm3/cm3 at the fraction of the biogas (50 kPa), together with ultra-high CO2/N2 and CO2/CH4 selectivities of 3538 and 151 at ambient conditions, respectively. This excellent separation performance was comprehensively explained by molecular simulations, which suggests that the carbon atom of CO2 can form strong electrostatic Cδ+···δ−O-C interactions with the oxygen atoms in the carbonate ligand and the oxygen atom of CO2 can interact with the hydrogen atoms in the triazolate ligand through Oδ−···δ+H-C interactions, enabling ZnAtzCO3 as an optimal nanotrap for CO2 fixation. Moreover, breakthrough experiments confirm excellent dynamic separation toward CO2/N2 and CO2/CH4 on ZnAtzCO3, highlighting its potential for selective CO2 capture. Furthermore, constructing suitable nanotraps with optimal electrostatic environment and pore geometry is worthy of further exploration in other separation circumstances.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28237908/s1, Table S1: Fitting parameters of the DSLF model for isotherms of CO2, N2, and CH4 on ZnAtzCO3; Table S2: Fitting parameters and correlation coefficients of the Virial equation for all gases on ZnAtzCO3; Scheme S1: Synthetic pathways for ZnAtzCO3 with starting reactants of ZnSO4/Hatz/DMF/H2O at 523 K; Figure S1: Asymmetric unit of ZnAtzCO3; Figure S2: Pore size distribution of ZnAtzCO3 based on the HK model using CO2 as the probe molecule; Figure S3: CO2 breakthrough times or CO2/N2 (v:v, 15:85) and CO2/CH4 (v:v, 15:85) mixtures in five consecutive cycles of breakthrough experiments on ZnAtzCO3.

Author Contributions

Conceptualization, J.P. and D.L.; methodology, J.P. and J.Z.; software, C.D. and J.X.; validation, C.F. and B.Y.; formal analysis, J.P. and J.Z.; investigation, J.P., C.F. and B.Y.; resources, J.X. and C.D.; data curation, J.Z. and C.F.; writing—original draft preparation, J.P.; writing—review and editing, all authors; visualization, J.P., J.Z. and C.D.; supervision, D.L.; project administration, J.P. and D.L.; funding acquisition, J.P., J.X., C.D. and D.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (No. 21908069 and 22108034), China Postdoctoral Science Foundation (No. 2020M672636), Guangdong Provincial Natural Science Foundation Project (No. 2023A1515012151), Guangdong Basic and Applied Basic Research Foundation (No. 2023A1515011881), Guangdong-Hong Kong Technology Cooperation Funding Scheme (No. 2023A0505010002), and Scientific Research Project of Guangdong Provincial Department of Education (No. 2022KTSCX122).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

Author Jiqin Zhong was employed by the company GAC R&D Center, Guangzhou Automobile Group 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. Bourzac, K. We have the technology. Nature 2017, 550, S66–S69. [Google Scholar] [CrossRef] [PubMed]
  2. Schuur, E.A.; McGuire, A.D.; Schädel, C.; Grosse, G.; Harden, J.; Hayes, D.J.; Hugelius, G.; Koven, C.D.; Kuhry, P.; Lawrence, D.M. Climate change and the permafrost carbon feedback. Nature 2015, 520, 171–179. [Google Scholar] [CrossRef] [PubMed]
  3. Seneviratne, S.I.; Rogelj, J.; Séférian, R.; Wartenburger, R.; Allen, M.R.; Cain, M.; Millar, R.J.; Ebi, K.L.; Ellis, N.; Hoegh-Guldberg, O. The many possible climates from the Paris agreement’s aim of 1.5 C warming. Nature 2018, 558, 41–49. [Google Scholar] [CrossRef] [PubMed]
  4. Majeed, H.; Iftikhar, T.; Ahmad, K.; Qureshi, K.; Tabinda; Altaf, F.; Iqbal, A.; Ahmad, S.; Khalid, A. Bulk industrial production of sustainable cellulosic printing fabric using agricultural waste to reduce the impact of climate change. Int. J. Biol. Macromol. 2023, 253, 126885. [Google Scholar] [CrossRef]
  5. Trends in Atmospheric Carbon Dioxide. Available online: https://gml.noaa.gov/ccgg/trends/mlo.html (accessed on 10 February 2021).
  6. International Energy Agency. CO2 Emissions in 2022; International Energy Agency: Paris, France, 2023.
  7. Sholl, D.S.; Lively, R.P. Seven chemical separations to change the world. Nature 2016, 532, 435–437. [Google Scholar] [CrossRef]
  8. Gao, W.; Liang, S.; Wang, R.; Jiang, Q.; Zhang, Y.; Zheng, Q.; Xie, B.; Toe, C.Y.; Zhu, X.; Wang, J. Industrial carbon dioxide capture and utilization: State of the art and future challenges. Chem. Soc. Rev. 2020, 49, 8584–8686. [Google Scholar] [CrossRef]
  9. Singh, G.; Lee, J.; Karakoti, A.; Bahadur, R.; Yi, J.; Zhao, D.; AlBahily, K.; Vinu, A. Emerging trends in porous materials for CO2 capture and conversion. Chem. Soc. Rev. 2020, 49, 4360–4404. [Google Scholar] [CrossRef]
  10. Assunção, L.R.; Mendes, P.A.; Matos, S.; Borschiver, S. Technology roadmap of renewable natural gas: Identifying trends for research and development to improve biogas upgrading technology management. Appl. Energy 2021, 292, 116849. [Google Scholar] [CrossRef]
  11. Ning, H.; Li, Y.; Zhang, C. Recent progress in the integration of CO2 capture and utilization. Molecules 2023, 28, 4500. [Google Scholar] [CrossRef]
  12. Usman, M.; Iqbal, N.; Noor, T.; Zaman, N.; Asghar, A.; Abdelnaby, M.M.; Galadima, A.; Helal, A. Advanced strategies in metal-organic frameworks for CO2 capture and separation. Chem. Rec. 2022, 22, e202100230. [Google Scholar] [CrossRef]
  13. Modak, A.; Bhaumik, A. Porous carbon derived via KOH activation of a hypercrosslinked porous organic polymer for efficient CO2, CH4, H2 adsorptions and high CO2/N2 selectivity. J. Solid State Chem. 2015, 232, 157–162. [Google Scholar] [CrossRef]
  14. Kundu, S.K.; Bhaumik, A. Novel nitrogen and sulfur rich hyper-cross-linked microporous poly-triazine-thiophene copolymer for superior CO2 capture. ACS Sustain. Chem. Eng. 2016, 4, 3697–3703. [Google Scholar] [CrossRef]
  15. Gao, X.; Yan, W.-H.; Hu, B.-Y.; Huang, Y.-X.; Zheng, S.-M. Porous metal-organic frameworks for light hydrocarbon separation. Molecules 2023, 28, 6337. [Google Scholar] [CrossRef] [PubMed]
  16. Adil, K.; Belmabkhout, Y.; Pillai, R.S.; Cadiau, A.; Bhatt, P.M.; Assen, A.H.; Maurin, G.; Eddaoudi, M. Gas/vapour separation using ultra-microporous metal-organic frameworks: Insights into the structure/separation relationship. Chem. Soc. Rev. 2017, 46, 3402–3430. [Google Scholar] [CrossRef]
  17. Yaghi, O.M.; Kalmutzki, M.J.; Diercks, C.S. Introduction to Reticular Chemistry: Metal-Organic Frameworks and Covalent Organic Frameworks; John Wiley & Sons: Hoboken, NJ, USA, 2019. [Google Scholar]
  18. Rajendran, A.; Subraveti, S.G.; Pai, K.N.; Prasad, V.; Li, Z. How can (or why should) process engineering aid the screening and discovery of solid sorbents for CO2 capture? Acc. Chem. Res. 2023, 56, 2354–2365. [Google Scholar] [CrossRef]
  19. Zhu, M.; Hu, P.; Tong, Z.; Zhao, Z.; Zhao, Z. Enhanced hydrophobic MIL(Cr) metal-organic framework with high capacity and selectivity for benzene VOCs capture from high humid air. Chem. Eng. J. 2017, 313, 1122–1131. [Google Scholar] [CrossRef]
  20. Ahmad, K.; Nazir, M.A.; Qureshi, A.K.; Hussain, E.; Najam, T.; Javed, M.S.; Shah, S.S.A.; Tufail, M.K.; Hussain, S.; Khan, N.A.; et al. Engineering of zirconium based metal-organic frameworks (Zr-MOFs) as efficient adsorbents. Mater. Sci. Eng. B 2020, 262, 114766. [Google Scholar] [CrossRef]
  21. Zhang, Z.; Yao, Z.-Z.; Xiang, S.; Chen, B. Perspective of microporous metal-organic frameworks for CO2 capture and separation. Energy Environ. Sci. 2014, 7, 2868–2899. [Google Scholar] [CrossRef]
  22. Kim, Y.; Huh, S. Pore engineering of metal-organic frameworks: Introduction of chemically accessible lewis basic sites inside mof channels. Cryst. Eng. Comm. 2016, 18, 3524–3550. [Google Scholar] [CrossRef]
  23. He, T.; Kong, X.-J.; Li, J.-R. Chemically stable metal-organic frameworks: Rational construction and application expansion. Acc. Chem. Res. 2021, 54, 3083–3094. [Google Scholar] [CrossRef]
  24. Chen, Z.; Kirlikovali, K.O.; Shi, L.; Farha, O.K. Rational design of stable functional metal-organic frameworks. Mater. Horiz. 2023, 10, 3257–3268. [Google Scholar] [CrossRef] [PubMed]
  25. Shi, Z.; Tao, Y.; Wu, J.; Zhang, C.; He, H.; Long, L.; Lee, Y.; Li, T.; Zhang, Y.-B. Robust metal–triazolate frameworks for CO2 capture from flue gas. J. Am. Chem. Soc. 2020, 142, 2750–2754. [Google Scholar] [CrossRef] [PubMed]
  26. Lin, J.-B.; Nguyen, T.T.T.; Vaidhyanathan, R.; Burner, J.; Taylor, J.M.; Durekova, H.; Akhtar, F.; Mah, R.K.; Ghaffari-Nik, O.; Marx, S.; et al. A scalable metal-organic framework as a durable physisorbent for carbon dioxide capture. Science 2021, 374, 1464–1469. [Google Scholar] [CrossRef] [PubMed]
  27. Nguyen, T.T.T.; Lin, J.-B.; Shimizu, G.K.H.; Rajendran, A. Separation of CO2 and N2 on a hydrophobic metal organic framework CALF-20. Chem. Eng. J. 2022, 442, 136263. [Google Scholar] [CrossRef]
  28. Vaidhyanathan, R.; Iremonger, S.S.; Dawson, K.W.; Shimizu, G.K. An amine-functionalized metal organic framework for preferential CO2 adsorption at low pressures. Chem. Commun. 2009, 35, 5230–5232. [Google Scholar] [CrossRef] [PubMed]
  29. Banerjee, A.; Nandi, S.; Nasa, P.; Vaidhyanathan, R. Enhancing the carbon capture capacities of a rigid ultra-microporous mof through gate-opening at low CO2 pressures assisted by swiveling oxalate pillars. Chem. Commun. 2016, 52, 1851–1854. [Google Scholar] [CrossRef]
  30. Peng, J.; Liu, Z.; Wu, Y.; Xian, S.; Li, Z. High-performance selective CO2 capture on a stable and flexible metal-organic framework via discriminatory gate-opening effect. ACS Appl. Mater. Interfaces 2022, 14, 21089–21097. [Google Scholar] [CrossRef]
  31. 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]
  32. 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. Inter. Ed. 2021, 60, 5283–5288. [Google Scholar] [CrossRef]
  33. Wen, H.-M.; Liu, M.; Ling, Y.; Gu, X.-W.; Liu, D.; Yu, C.; Liang, Y.; Xie, B.; Li, B.; Hu, J. A metal-organic framework based propylene nano-trap with dual functionalities for highly efficient propylene/propane separation. J. Mater. Chem. A 2023, 11, 17821–17827. [Google Scholar] [CrossRef]
  34. Zhu, H.; Wang, Y.; Wang, X.; Fan, Z.-W.; Wang, H.-F.; Niu, Z.; Lang, J.-P. Design of a MOF-based nano-trap for the efficient separation of propane from propylene. Chem. Commum. 2023, 59, 5757–5760. [Google Scholar] [CrossRef]
  35. Chen, Y.; Qiao, Z.; Lv, D.; Duan, C.; Sun, X.; Wu, H.; Shi, R.; Xia, Q.; Li, Z. Efficient adsorptive separation of C3H6 over C3H8 on flexible and thermoresponsive CPL-1. Chem. Eng. J. 2017, 328, 360–367. [Google Scholar] [CrossRef]
  36. Zhang, P.; Yang, L.; Liu, X.; Wang, J.; Suo, X.; Chen, L.; Cui, X.; Xing, H. Ultramicroporous material based parallel and extended paraffin nano-trap for benchmark olefin purification. Nat. Commun. 2022, 13, 4928. [Google Scholar] [CrossRef] [PubMed]
  37. Basnayake, S.A.; Su, J.; Zou, X.; Balkus, K.J., Jr. Carbonate-based zeolitic imidazolate framework for highly selective CO2 capture. Inorg. Chem. 2015, 54, 1816–1821. [Google Scholar] [CrossRef]
  38. Lin, Y.-Y.; Zhang, Y.-B.; Zhang, J.-P.; Chen, X.-M. Pillaring Zn-triazolate layers with flexible aliphatic dicarboxylates into three-dimensional metal-organic frameworks. Cryst. Growth Des. 2008, 8, 3673–3679. [Google Scholar] [CrossRef]
  39. Frost, R.L.; Martens, W.N.; Wain, D.L.; Hales, M.C. Infrared and infrared emission spectroscopy of the zinc carbonate mineral smithsonite. Spectrochim. Acta Part A 2008, 70, 1120–1126. [Google Scholar] [CrossRef] [PubMed]
  40. Myers, A.L.; Prausnitz, J.M. Thermodynamics of mixed-gas adsorption. AIChE J. 1965, 11, 121–127. [Google Scholar] [CrossRef]
  41. Walton, K.S.; Sholl, D.S. Predicting multicomponent adsorption: 50 years of the ideal adsorbed solution theory. AIChE J. 2015, 61, 2757–2762. [Google Scholar] [CrossRef]
  42. Peng, J.; Zhong, J.; Liu, Z.; Xi, H.; Yan, J.; Xu, F.; Chen, X.; Wang, X.; Lv, D.; Li, Z. Multivariate metal-organic frameworks prepared by simultaneous metal/ligand exchange for enhanced c2–c3 selective recovery from natural gas. ACS Appl. Mater. Interfaces 2023, 15, 41466–41475. [Google Scholar] [CrossRef]
  43. Li, J.-R.; Kuppler, R.J.; Zhou, H.-C. Selective gas adsorption and separation in metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1477–1504. [Google Scholar] [CrossRef]
  44. Zhang, J.-P.; Zhu, A.-X.; Lin, R.-B.; Qi, X.-L.; Chen, X.-M. Pore surface tailored sod-type metal-organic zeolites. Adv. Mater. 2011, 23, 1268–1271. [Google Scholar] [CrossRef] [PubMed]
  45. Yu, C.; Ding, Q.; Hu, J.; Wang, Q.; Cui, X.; Xing, H. Selective capture of carbon dioxide from humid gases over a wide temperature range using a robust metal-organic framework. Chem. Eng. J. 2021, 405, 126937. [Google Scholar] [CrossRef]
  46. Xing, G.E.; Liu, Q.; Zhang, Y.; Zhang, S.; Dong, Y. Microporous zinc(ii) metal-organic framework with 6-connected pcu topology: Synthesis, structure, and gas adsorption properties. Z. Anorg. Allg. Chem. 2015, 641, 1556–1559. [Google Scholar] [CrossRef]
  47. Zhai, Q.-G.; Bai, N.; Li, S.N.; Bu, X.; Feng, P. Design of pore size and functionality in pillar-layered zn-triazolate-dicarboxylate frameworks and their high CO2/CH4 and C2 hydrocarbons/CH4 selectivity. Inorg. Chem. 2015, 54, 9862–9868. [Google Scholar] [CrossRef]
  48. Materials Studio, v7.0, Biovia Software Inc.: San Diego, CA, USA, 2013.
  49. Czepirski, L.; JagieŁŁo, J. Virial-type thermal equation of gas-solid adsorption. Chem. Eng. Sci. 1989, 44, 797–801. [Google Scholar] [CrossRef]
  50. Nuhnen, A.; Janiak, C. A practical guide to calculate the isosteric heat/enthalpy of adsorption via adsorption isotherms in metal-organic frameworks, mofs. Dalt. Trans. 2020, 49, 10295–10307. [Google Scholar] [CrossRef]
Figure 1. Selective CO2 capture from N2 and CH4 on a nanotrap with a suitable electrostatic environment via multiple host–guest interactions. The blue dotted lines represent the electrostatic interactions between the framework and the CO2 molecule.
Figure 1. Selective CO2 capture from N2 and CH4 on a nanotrap with a suitable electrostatic environment via multiple host–guest interactions. The blue dotted lines represent the electrostatic interactions between the framework and the CO2 molecule.
Molecules 28 07908 g001
Figure 2. The crystal structure and pore property of ZnAtzCO3: (a) coordination mode, (b) crystal structure shown in the b-axis, and the Connolly surface in the b-axis (c) and a-axis (d) by using a spherical probe exhibiting a radius of 1 Å. The intraframework N-H···O hydrogen bonds are marked by the golden dotted lines. I and II in Figure 2c represent the two types of cavities on the structure.
Figure 2. The crystal structure and pore property of ZnAtzCO3: (a) coordination mode, (b) crystal structure shown in the b-axis, and the Connolly surface in the b-axis (c) and a-axis (d) by using a spherical probe exhibiting a radius of 1 Å. The intraframework N-H···O hydrogen bonds are marked by the golden dotted lines. I and II in Figure 2c represent the two types of cavities on the structure.
Molecules 28 07908 g002
Figure 3. FTIR image (a), PXRD patterns (b), TG curves (c), CO2 adsorption–desorption isotherms at 195 K (d) of ZnAtzCO3.
Figure 3. FTIR image (a), PXRD patterns (b), TG curves (c), CO2 adsorption–desorption isotherms at 195 K (d) of ZnAtzCO3.
Molecules 28 07908 g003
Figure 4. (a) Single-component adsorption isotherms of CO2, N2 and CH4 on ZnAtzCO3 at 298 K. (b) IAST selectivity of CO2/N2 (15:85, v:v) and CO2/CH4 (15:85, v:v) on ZnAtzCO3.
Figure 4. (a) Single-component adsorption isotherms of CO2, N2 and CH4 on ZnAtzCO3 at 298 K. (b) IAST selectivity of CO2/N2 (15:85, v:v) and CO2/CH4 (15:85, v:v) on ZnAtzCO3.
Molecules 28 07908 g004
Figure 5. (a) Single-component adsorption isotherms of CO2, N2, and CH4 on ZnAtzCO3 at different temperatures (273 K, 288 K, and 298 K). (b) Isosteric heat of CO2, N2, and CH4 on ZnAtzCO3.
Figure 5. (a) Single-component adsorption isotherms of CO2, N2, and CH4 on ZnAtzCO3 at different temperatures (273 K, 288 K, and 298 K). (b) Isosteric heat of CO2, N2, and CH4 on ZnAtzCO3.
Molecules 28 07908 g005
Figure 6. Preferential adsorption sites on the ZnAtzCO3 structure for CO2 (a), CH4 (b), and N2 (c) on ZnAtzCO3. The dashed line represents the host–guest interactions between the ZnAtzCO3 and the gas molecules. The unit of the distance between the gas molecules and the adsorption site is Å.
Figure 6. Preferential adsorption sites on the ZnAtzCO3 structure for CO2 (a), CH4 (b), and N2 (c) on ZnAtzCO3. The dashed line represents the host–guest interactions between the ZnAtzCO3 and the gas molecules. The unit of the distance between the gas molecules and the adsorption site is Å.
Molecules 28 07908 g006
Figure 7. (ac) The simulated adsorption density distribution of CO2, CH4, and N2 on ZnAtzCO3 crystal framework at 15 kPa. (df) The simulated adsorption density distribution of CO2, CH4 and N2 on ZnAtzCO3 crystal framework at 100 kPa.
Figure 7. (ac) The simulated adsorption density distribution of CO2, CH4, and N2 on ZnAtzCO3 crystal framework at 15 kPa. (df) The simulated adsorption density distribution of CO2, CH4 and N2 on ZnAtzCO3 crystal framework at 100 kPa.
Molecules 28 07908 g007
Figure 8. The simulated interaction energy for CO2, CH4, and N2 on ZnAtzCO3.
Figure 8. The simulated interaction energy for CO2, CH4, and N2 on ZnAtzCO3.
Molecules 28 07908 g008
Figure 9. Breakthrough curves for CO2/N2 (15:85, v:v) (a) and CO2/CH4 (50:50, v:v) (b) mixture on ZnAtzCO3 at 298 K and 100 kPa.
Figure 9. Breakthrough curves for CO2/N2 (15:85, v:v) (a) and CO2/CH4 (50:50, v:v) (b) mixture on ZnAtzCO3 at 298 K and 100 kPa.
Molecules 28 07908 g009
Table 1. Comparisons of CO2 uptakes at 15 kPa and 50 kPa on typical MOFs constructed by triazolate linkers.
Table 1. Comparisons of CO2 uptakes at 15 kPa and 50 kPa on typical MOFs constructed by triazolate linkers.
MOFsQCO2 at 15 kPa
(cm3/g, STP)
QCO2 at 50 kPa
(cm3/g, STP)
Qst
(kJ/mol)
T
(K)
Ref.
MAF-74.512.525298[44]
ZnF(TZ)6.019.124298[25]
ZnF(daTZ)21.434.133298[25]
ZnDatzBdc1.95.829298[30]
CALF-2053.768.333.5303[26]
ZnAtzOx60.565.055303[29]
ZU-30147.752.639298[45]
Zn(FA)(datrz)212.328.524.0298[46]
Zn2(TRZ)2(BDC)23.541.4-298[47]
Zn2(TRZ)2(FA)34.773.9-298[47]
ZnAtzCO344.856.032.6298This work
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

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. https://doi.org/10.3390/molecules28237908

AMA Style

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(23):7908. https://doi.org/10.3390/molecules28237908

Chicago/Turabian Style

Peng, Junjie, Chengmin Fu, Jiqin Zhong, Bin Ye, Jing Xiao, Chongxiong Duan, and Daofei Lv. 2023. "Efficient Selective Capture of Carbon Dioxide from Nitrogen and Methane Using a Metal-Organic Framework-Based Nanotrap" Molecules 28, no. 23: 7908. https://doi.org/10.3390/molecules28237908

Article Metrics

Back to TopTop