Next Article in Journal
Intervention Effects of Deer-Tendon Collagen Hydrolysates on Osteoporosis In Vitro and In Vivo
Next Article in Special Issue
Efficient Selective Capture of Carbon Dioxide from Nitrogen and Methane Using a Metal-Organic Framework-Based Nanotrap
Previous Article in Journal
Heterometallic ZnHoMOF as a Dual-Responsive Luminescence Sensor for Efficient Detection of Hippuric Acid Biomarker and Nitrofuran Antibiotics
Previous Article in Special Issue
A Microporous Zn(bdc)(ted)0.5 with Super High Ethane Uptake for Efficient Selective Adsorption and Separation of Light Hydrocarbons
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exquisitely Constructing a Robust MOF with Dual Pore Sizes for Efficient CO2 Capture

1
CNPC Petrochemical Research Institute Company Limited, Beijing 102206, China
2
College of Chemical Engineering and Environment, China University of Petroleum-Beijing, Beijing 102249, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(17), 6276; https://doi.org/10.3390/molecules28176276
Submission received: 26 July 2023 / Revised: 17 August 2023 / Accepted: 18 August 2023 / Published: 28 August 2023
(This article belongs to the Special Issue Zeolites and Porous Materials: Synthesis, Properties and Applications)

Abstract

:
Developing metal–organic framework (MOF) adsorbents with excellent performance and robust stability is of critical importance to reduce CO2 emissions yet challenging. Herein, a robust ultra-microporous MOF, Cu(bpfb)(bdc), with mixed ligands of N, N′-(1,4-phenylene)diisonicotinamide (bpfb), and 1,4-dicarboxybenzene (bdc) was delicately constructed. Structurally, this material possesses double-interpenetrated frameworks formed by two staggered, independent frameworks, resulting in two types of narrow ultra-micropores of 3.4 × 5.0 and 4.2 × 12.8 Å2, respectively. The above structural properties make its highly selective separation at 273~298 K with a CO2 capacity of 71.0~86.2 mg/g. Its adsorption heat over CO2 and IAST selectivity were calculated to be 27 kJ/mol and 52.2, respectively. Remarkably, cyclic breakthrough experiments corroborate its impressive performance in CO2/N2 separation in not only dry but also 75% RH humid conditions. Molecular simulation reveals that C-H···OCO2 in the pores plays a pivotal role in the high selectivity of CO2 adsorption. These results point out the huge potential application of this material for CO2/N2 separation.

Graphical Abstract

1. Introduction

With worldwide rapid industrialization, CO2 has become a major greenhouse gas and causes many serious problems, such as global warming and climate change, threatening the sustainable development of human society [1,2]. One of the most important reasons for incremental CO2 emission into the atmosphere is the burning of fossil fuels [3]. In particular, flue gas, which consists of ~15% CO2, 75% N2, and other impurities, released from power plants, steel, cement, and petrochemical industries, is a main source of CO2 emissions [4]. Hence, CO2 capture from flue gas is of great significance and urgency to achieve net zero emission goals, attracting considerable attention from scientists.
Among all separation technologies, absorption of CO2 by amines is the most popular way [5,6]. Although effective, it endures high energy costs during the regeneration process and causes pipeline corrosion. On the other hand, adsorption has milder performing conditions and lower energy costs, as well as simple equipment, providing an alternative technique [7]. As the highly close physical properties and kinetic diameters among CO2 (3.3 Å), N2 (3.65 Å), and other gas molecules, adsorbents possessing highly selective discrimination for CO2, as well as excellent stability from humidity, serve as the core in such a process.
Metal–organic frameworks (MOFs) are promising adsorbents for gas separation, for their many attractive properties including high specific surface area and facile regulation of pore size and structure [8,9,10]. To date, a huge number of MOF adsorbents have been reported. Generally, introducing open metal sites (OMS, metal sites with solvent molecules coordinated or unsaturated) and grafting Lewis basic sites (-NH2, alkylamine, arylamine, etc.) or other functional polar sites (-F, -CN, -PF6, etc.), which could enhance MOF–CO2 interaction, are effective strategies to achieve high adsorption capacity or separation selectivity, but usually result in higher binding energy or adsorption heat (Qst) [11]. For instance, Mg-MOF-74 [12,13], with abundant OMS, reached a dramatic capacity of 8.0 mmol/g at 1 bar, 296 K. However, this MOF suffers from high Qst value and vulnerability to moisture as a side-effect of the introduced OMS. Besides the above-mentioned strategies, molecular sieving through crystal engineering strategies on MOF, particularly for pore engineering by tuning the pore size, volume, shape, and surface, provides as an alternative way with a low Qst value [14,15,16,17,18,19,20,21,22]. Generally, a MOF with a confined pore size (<4 Å) usually exhibits high selectivity but low adsorption capacity, while an expanded pore size (>4 Å) usually shows high adsorption capacity but low selectivity [19]. Therefore, in contrast to MOFs representing uniform 1D pore/channel, it might achieve an ideal balance between adsorption uptake and selectivity by constructing a MOF featuring two types of pores in size. However, precise control of a MOF’s pore/channel in the high resolution of 0.1 Å remains a great challenge.
Based on these concepts, we synthesized an ultra-microporous pillar-layered MOF, Cu(bpfb)(bdc), termed PRI-1 (PRI stands for Petrochemical Research Institute), with dual pore sizes for efficient CO2/N2 separation. Firstly, the material’s crystal purity and porosity were characterized by PXRD (powder X-ray diffraction) and low-temperature CO2 adsorption isotherm. Secondly, the adsorbent’s thermal and chemical stability was investigated via TG and solvent-immersing tests. Thirdly, its adsorption isotherms for CO2 and N2 were characterized under various temperatures and the IAST model [23,24] was utilized to estimate its selectivity over CO2/N2 binary mixture at 298 K. Fourthly, its dynamic adsorption/separation performance for CO2/N2 mixtures in dry and highly humid conditions was studied. Finally, the molecular simulation was performed to reveal the adsorption mechanism of CO2/N2 on PRI-1.

2. Results and Discussion

2.1. Sample Characterization

PRI-1 was facilely synthesized via solvothermal reaction from Cu(NO3)2·3H2O, bpfb, and H2bdc in DMF (Figure 1A and Figure S1). The PXRD patterns of as-synthesized PRI-1 and simulated (CCDC No: 641709) [25] were nearly identical (Figure 1B). The characteristic peaks of 9.0° and 13.3° of the as-synthesized PRI-1 agree well with the calculated peaks. Togethering with IR spectroscopy [25] (Figure S2), the PXRD patterns indicate the successful synthesis of PRI-1. Meanwhile, the mother liquor circulation synthesis method was successfully utilized to further reduce the synthesis cost of this material (Figure S3). The morphology and size of PRI-1 were revealed by scanning electron microscopy. Figure S4 shows that PRI-1 crystalized in rods with particle sizes in micrometers.
Structurally, this MOF is described by a pcu-type framework with double-interpenetrated nets. In each net, two Cu2+ cations serve as the metal center and are coordinated by four bdc bidentate ligands, forming the layer. In the axial orientation, four bpfb ligands are coordinated, serving as the pillar. In the occurrence to the twofold interpenetration, there are 3675.6 Å3 accessible voids (30.7% of the cell volume), as calculated using the Mercury software [26]. Meanwhile, PRI-1 possesses two types of 1D channels in parallel and their sizes are 3.4 × 5.0 and 4.2 × 12.8 Å2, respectively. These parameters are very close to the dynamic radius of CO2.
The CO2 adsorption-desorption isotherm of PRI-1 at 195 K showed the ultra-microporous structure of the material (Figure 1C) with adsorption reaching 87.5 cm3/g at 1 bar and a calculated BET surface area of 317.3 m2/g. The porosity endows the material with the ability to separate CO2 efficiently.
The thermal stability of PRI-1 was examined via thermogravimetric analysis in an air atmosphere. As shown in Figure 1D, there is a minor weight loss starting from 30 to 290 °C, which is mainly attributed to the removal of guest solvents, and then a major weight loss emerged after 290 °C, indicating the framework collapse of the material. The TGA curve shows the high thermal robustness of the material until 290 °C, which is superior to most recently-reported CO2 adsorbents, including Mg-gallate (140 °C) [27], dptz-CuTiF6 (150 °C) [21], SIFSIX-3-Zn (160 °C) [28], FJUT-4 (200 °C) [29], and dptz-CuGeF6 (200 °C) [30], MUT-4 (250 °C) [31], as well as slight lower than that of TiFSIX-Cu-TPA (308 °C) [19] and CALF-20 (350 °C) [32] (Figure 1D).

2.2. CO2 and N2 Isotherms, Adsorption Heat, and Selectivity of PRI-1

Inspired by its structural features, the CO2 and N2 isotherms of PRI-1 were collected at 273, 283, and 298 K, respectively, as depicted in Figure 2A. The CO2 capacity was up to 71.0~86.2 mg/g at 1 bar, while only 8.1~10.2 mg/g for N2, suggesting an obviously preferential CO2 adsorption. The isosteric heat (Qst) of CO2 was calculated using the Clausius–Clapeyron equation [25] based on adsorption isotherms at 283 and 298 K, demonstrating a value of 27.0 kJ mol−1 at CO2 loading of 0.1 mmol/g (Figure 2B). Its Qst value is much lower than aqueous amine (105 kJ/mol) [21] and many MOF adsorbents (Figure 2C), including Mg2(dobpdc)(3-4-3) (99 kJ/mol) [20], Cu-BTTri (90 kJ/mol) [33], mmen-Mg2(dobpdc) (71 kJ/mol) [34], CD-MOF-2 (67.2 kJ/mol) [35], MOF-808-Gly (46 kJ/mol) [36], Mg-MOF-74 (47 kJ/mol) [37], Zeolite 13X (44~54 kJ/mol), SIFSIX-3-Zn (45 kJ/mol) [28], FJUT-3 (41.7 kJ/mol) [38], TiFSIX-Cu-TPA (39.2 kJ/mol) [19], CALF-20 (38.4 kJ/mol) [32], dptz-CuTiF6 (38.2 kJ/mol) [21], Mg–gallate (37 kJ/mol) [27], FJUT-4 (35.2 kJ/mol) [29], dptz-CuGeF6 (30.3 kJ/mol) [30], etc. This low Qst value would largely reduce the regeneration energy cost.
Based on the isotherm data, the ideal adsorbed solution theory (IAST) [23,24] was employed to qualitatively estimate the CO2/N2 selectivity of PRI-1. The dual-site Langmuir and Freundlich (DSLF) model was applied to fit the single component isotherms of CO2 and N2 at 298 K. Figures S5 and S6 show the fitting curves and parameters of the DSLF model. Figure S7 and Figure 2D show that the CO2/N2 selectivity of PRI-1 reached 52.2 (the detailed calculation procedures of IAST selectivity see Supplementary Excel file), higher than many MOFs in the literature.

2.3. Adsorption Mechanism of CO2 and N2 on PRI-1

DFT calculations were employed to elucidate the adsorption mechanism of carbon dioxide and nitrogen on PRI-1. Figure 3 illustrates the adsorption sites of CO2 and N2 on the PRI-1 framework. The figure reveals that CO2 molecules are captured within the main channels of PRI-1, showcasing the van der Waals interaction between gas molecules and hydrogen on the benzene ring and pyridine ring. The distance between the O···H-C bond is calculated as 2.924~3.102 Å, as depicted in Figure 3A, thereby highlighting the robust gas–ligand interactions during the adsorption process. This interaction arises from multi-directional hydrogen bond donors, which collectively confine the adsorption of CO2 molecules within the pores of PRI-1. Simultaneously, N2 molecules are captured within the main channels of PRI-1 due to the influence of hydrogen originating from the same side pyridine ring and amide group, exhibiting an N...H bond distance of 3.079~3.388 Å, as depicted in Figure 3B. Compared to the multi-directional adsorption of CO2, the unidirectional adsorption of N2 is relatively weaker within the PRI-1 framework. Furthermore, the calculated adsorption energies of carbon dioxide and nitrogen on PRI-1 amount to 28.44 kJ/mol and 15.52 kJ/mol, respectively, aligning with the experimental findings. PRI-1 possesses a distinctive pore structure and multiple ligands that facilitate extensive interactions between gas molecules and the framework, rendering it an excellent MOF material for the efficient separation of CO2/N2.

2.4. Breakthrough Curve of CO2/N2 Binary Mixtures

To evaluate the dynamic separation properties of PRI-1, breakthrough tests were carried out with a simulated binary mixture of CO2/N2 (15:85) at 298 K. Figure 4 presents the breakthrough curve at 298 K of the binary mixture through a fixed bed of PRI-1. It was obvious that N2 was eluted out almost immediately after the gas mixture fed in, while CO2 was retained for 30.3 min/g (Figure 4A) under the flow rate of 2 mL/min, superior to the separation performance to many MOFs including the recently reported molecular sieving adsorbents, FJUT-4 [23]. As recyclability is of great importance for practical use, the breakthrough experiments were repeated for five cycles and the separation performance displayed a negligible loss (Figure 4B).
More importantly, the separation properties in humid conditions also play a vital role in its practical use. Considering its robustness in various harsh solvents as mentioned above, we further conducted the breakthrough experiments under humid conditions. Similar to the robust ZIF-94 [39] and FJUT-3 [38], the presence of humidity (75% RH, 298 K) did not significantly alter the CO2 breakthrough time for PRI-1 (Figure 4C), showing efficient CO2/N2 separation ability in both dry and humid conditions.

3. Conclusions

In this paper, an ultra-microporous MOF, PRI-1, was facilely synthesized via a solvothermal method for efficient separation of CO2/N2. Structurally, PRI-1 possesses double-interpenetrated frameworks formed by two staggered nets, resulting in dual pore sizes. PRI-1 shows its high thermal stability up to 290 °C, as well as its structural robustness from moist air and various polar solvents, including water. Its CO2 and N2 uptake were 71.0~86.2 mg/g and 8.1~10.2 mg/g at 1 bar and 273~298 K, respectively. Qst of CO2 was calculated to be 27.0 kJ/mol at the loading of 0.1 mmol/g. The calculated IAST selectivity for CO2/N2 (15:85, v/v) was 52.2 at 298 K and 1 bar. The breakthrough experiments prove it possesses excellent dynamic separation properties under both dry and 75% RH humid conditions. Molecular simulation revealed that the tight C-H…OCO2 interaction contributes. This work proves that the exquisite construction of an ultra-microporous MOF with dual pore sizes might be an efficient and facile strategy for efficient CO2 separation.

4. Materials and Methods

4.1. Materials

Copper (II) nitrate hexahydrate (Cu(NO3)2 3H2O, 99%), N,N-Dimethylformamide (DMF, 99.5%), and methanol (CH3OH, 99%) were purchased from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Benzene-1,4-dicarboxylic acid (H2bdc, 99%) was purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). N,N′-(1,4-phenylene)diisonicotinamide (bpfb, 95%) was purchased from Jilin Chinese Academy of Sciences—Yanshen Technology Co., Ltd. (Changchun, China). All reagents were used directly without any purification.

4.2. Synthesis of PRI-1

The synthesis procedure of PRI-1 was optimized from previous literature [25]. Firstly, Cu(NO3)2 3H2O (0.5 mmol, 0.121 g), bpfb (0.5 mmol, 0.159 g), and H2bdc (0.5 mmol, 0.083 g) were dissolved in DMF (20 mL) and under ultrasonicated for 30 min at room temperature. Afterward, the resulting brown clear solution was transferred into the autoclave and heated at 393 K for 48 h, and then cooled down at the rate of 3 K/h. The obtained brown powders were washed with DMF and CH3OH thrice, respectively. After being soaked in CH3OH for 1 day and heated at 373 K under vacuum for 12 h, 0.259 g of sample was obtained at a yield of 94.9% (based on Cu).

4.3. Mother Liquor Circulation Synthesized Method

The mother liquor after the first reaction was reused for the next turn of PRI-1 synthesis. Cu(NO3)2 3H2O (0.25 mmol, 0.061 g), bpfb (0.25 mmol, 0.080 g) and H2bdc (0.25 mmol, 0.042 g) were dissolved into the filtered mother liquor and ultrasonicated for 30 min at room temperature. Afterward, the resulting brown clear solution was transferred into the autoclave and heated at 393 K for 48 h, as above. The sample was obtained with a yield of 78.4%. The mother liquor circulation synthesized method would significantly reduce the cost of organic solvents in synthesis.

4.4. Characterization

The crystal structure and crystallinity of the sample were measured using PXRD on an X’Pert PRO powder diffractometer (PANalytical, Almelo, The Netherlands) with a Cu Kα radiation source (λ = 1.5418 Å). Scanning was performed over a 2θ range of 5~50° at a scanning rate of 2°/min. Thermogravimetric analysis (TGA) of the sample was carried out on a thermal analyzer (HCT-4, Beijing Hengjiu Experimental Equipment Co., Ltd. Beijing, China). The sample was heated at a rate of 5 °C/min under an air atmosphere. The morphology of the as-synthesized sample was observed and confirmed using SEM on a JEOL JSM-IT200 instrument. CO2 adsorption–desorption isotherms were recorded at 195 K on BSD-PM instruments. Before measurement, the sample was degassed at 100 °C for 12 h and all gases were of 99.999% purity.

4.5. Adsorption Isotherms Measurement

Single-component adsorption isotherms of CO2 and N2 were collected on a vacuum vapor/gas sorption analyzer (BSD-VVS, 3H-2000PW). Before each measurement, the sample was degassed and activated at 373 K under vacuum for at least 6 h until no weight loss was observed and subsequently was cooled to room temperature. High purity (99.9999%) CO2 and N2 were utilized in the gas adsorption experiments.

4.6. Solvent Stability Test

The stability of the samples was investigated by PXRD after being immersed into acetonitrile, DMF, methanol, and DI water for 24 h, respectively.

4.7. Fixed-Bed Breakthrough Experiments

The breakthrough curves of the gas mixture CO2/N2 (15:85, v/v) were tested on homemade dynamic breakthrough equipment (Figure S8). Similar to the packing technique in recent articles [19,27], 460 mg of dry sample was filled in a stainless-steel HPLC adsorption column (Φ4.6 × 50 mm). Before the experiment, the column was purged and activated under flowing He gas at 373 K overnight. During the experiment, the above gas mixture was injected with a flow rate of 2 mL min−1. The out-gas mixture was monitored in real time by a gas chromatography apparatus (7890A, Agilent Technologies, Inc. Santa Clara, CA, USA), equipped with a HayeSep Q column and a TCD detector. After each test, the column was regenerated with He gas flow at room temperature for 30 min, and the cycling breakthrough tests were performed under identical conditions as above.

4.8. Fitting with Dual-Site Langmuir-Freundlich Model

Single-component adsorption isotherms of CO2 and N2 on PRI-1 obtained at different temperatures were fitted using the dual-site Langmuir-Freundlich model [40] through Origin software:
n = N 1 × a × p b 1 + a × p b + N 2 × c × p d 1 + c × p d
where n is the equilibrium amount adsorbed in mmol/g, p is the equilibrium pressure in kPa, N1, and N2 is the adsorbed amount at site 1 and site 2, respectively; a is the maximal loading in mmol/g, a and c are the affinity constants of site 1 and site 2, respectively; and 1/b and 1/d are the deviations from an ideal homogeneous surface.

4.9. Ideal Adsorbed Solution Theory Calculations

In the context of IAST, it is postulated that the adsorbed mixture behaves as an ideal solution under constant spreading pressure and temperature. According to this theory, all components within the mixture conform to a rule similar to Raoult’s law, and the chemical potential of the adsorbed solution is assumed to be in equilibrium with that of the gas phase.
From IAST, the spreading pressure π is given by
π i 0 p i 0 = R T A 0 p i 0 a d ln p
π * = π A R T = 0 p i 0 q i p d p
where A is the specific surface area of the adsorbent, π and π* are the spreading pressure and reduced spreading pressure, respectively. p i 0 represents the individual gas pressures corresponding to component i at a given spreading pressure π of the gas mixture.
Under constant temperature, the spreading pressure remains constant for each individual component.
π 1 * = π 1 * = = π n * = π
In the case of binary adsorption involving component 1 and component 2, the IAST formulation requires the following expression,
y 1 p t = x 1 p 1 ( 1 y 1 ) p t = 1 x 1 p 2
where y1 and x1 represent the molar fractions of component 1 in the gas phase and adsorbed phase, respectively. The total gas pressure is denoted as pt, while p1 and p2 represent the pressures of component 1 and component 2 at the same spreading pressure as that of the mixture.
The adsorption selectivity in a binary mixture of component 1 and component 2 can be defined as
S a d s = x 1 / x 2 y 1 / y 2
where x1 and x2 represent the component molar loadings within the MOF, and y1 and y2 are the corresponding mole fraction in the bulk phase, respectively.

4.10. Calculation of Isosteric Heat of Adsorption

The isosteric heat of adsorption was calculated by analyzing the fitted adsorption isotherms at three different temperatures (i.e., 273 K, 283 K, 298 K) based on the Clausius–Clapeyron equation. The equation can be presented as follows:
l n p = H S R T + C
where p is the pressure (kPa), H S is the isosteric heat of adsorption at a given loading (kJ/mol), R is the ideal gas constant (8.314 kJ/mol/K), T is the temperature (K) and C is the integral constant.

4.11. Computational Method

All DFT calculations in this work utilize the VASP package [41,42] with the generalized gradient approximation (GGA) based on the Perdew–Burke–Ernzerhof (PBE) functional [43]. Valence electrons were simulated utilizing projector-augmented-wave (PAW) pseudopotentials [42,44], with a cutoff energy of 400 eV. A 3 × 3 × 2 Monkhorst–Pack k-point mesh sampling was utilized during simulations [45]. The required structural optimization accuracy was reached when the forces on the relaxed atoms were less than 0.05 eV/Å. The DFT-D3 approximation method was utilized to correct van der Waals (vdW) interactions [46]. Adsorption energies are calculated as ΔEads = EMOF+gas − (EMOF + Egas), where EMOF+gas is the total energy of the adsorption complex, and gas represents the CO2 or N2.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28176276/s1, Figure S1. The asymmetric unit of PRI-1. Figure S2. The FT-IR spectra of Cu(bpfb)(bdc) and raw materials. Figure S3. PXRD patterns of PRI-1 samples from the fresh synthesis and the mother liquor circulation synthesized method. Figure S4. SEM pictures of PRI. Figure S5. Gas adsorption isotherm of CO2 with the DSLF fit for PRI-1 at 273, 283 and 298 K. Figure S6. Gas adsorption isotherm of N2 with the DSLF fit for PRI-1 at 273, 283 and 298 K. Figure S7. IAST selectivity of PRI-1 for CO2/N2 (15:85) at 298 K. Figure S8. Diagram of the homemade dynamic breakthrough experimental apparatus. Supplementary Excel file. Detailed calculation procedures of IAST selectivity.

Author Contributions

Conceptualization, Y.L.; formal analysis, Y.L. and Q.G.; materials synthesis, Y.L. and Y.B. (Yuhua Bai); materials characterization, Y.L., Y.B. (Yuhua Bai) and H.X.; Data analysis, Y.L., Q.G. and Y.B. (Yuhua Bai). DFT calculations, Z.W.; breakthrough apparatus setup: Y.L., M.L., Y.B. (Yuhua Bai) and Y.B. (Yawen Bo); writing—original draft preparation, Y.L. and Z.W.; writing—review and editing, Q.G., K.C. and G.J.; visualization, Y.L., Z.W. and Q.G; funding, Q.G. and K.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by the Research and Development Department Project of PetroChina (2021DJ6002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Acknowledgments

The authors acknowledge Hao, Li from the Petrochemical Research Institute, PetroChina for developing automatic process software on GC data. We thank Beishide Instrument Technology (Beijing, China) Co., Ltd. and SCI-GO (www.sci-go.com, accessed on 15 August 2023) for the support of adsorption characterization.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Xu, Y.; Ramanathan, V.; Victor, D.G. Global Warming Will Happen Faster than We Think. Nature 2018, 564, 30–32. [Google Scholar] [CrossRef]
  2. Ritchie, H.; Roser, M.; Rosado, P. CO₂ and Greenhouse Gas Emissions. OurWorldInData.org. 2020. Available online: https://ourworldindata.org/co2-and-greenhouse-gas-emissions (accessed on 15 August 2023).
  3. Abas, N.; Kalair, A.; Khan, N. Review of Fossil Fuels and Future Energy Technologies. Futures 2015, 69, 31–49. [Google Scholar] [CrossRef]
  4. 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]
  5. Wang, M.; Lawal, A.; Stephenson, P.; Sidders, J.; Ramshaw, C. Post-Combustion CO2 Capture with Chemical Absorption: A State-of-the-Art Review. Chem. Eng. Res. Des. 2011, 89, 1609–1624. [Google Scholar] [CrossRef]
  6. Zhang, Z.; Borhani, T.N.; Olabi, A.G. Status and Perspective of CO2 Absorption Process. Energy 2020, 205, 118057. [Google Scholar] [CrossRef]
  7. Yu, C.-H.; Huang, C.-H.; Tan, C.-S. A Review of CO2 Capture by Absorption and Adsorption. Aerosol Air Qual. Res. 2012, 12, 745–769. [Google Scholar] [CrossRef]
  8. Al-Rowaili, F.N.; Zahid, U.; Onaizi, S.; Khaled, M.; Jamal, A.; AL-Mutairi, E.M. A Review for Metal-Organic Frameworks (MOFs) Utilization in Capture and Conversion of Carbon Dioxide into Valuable Products. J. CO2 Util. 2021, 53, 101715. [Google Scholar] [CrossRef]
  9. Ding, M.; Flaig, R.W.; Jiang, H.-L.; Yaghi, O.M. Carbon Capture and Conversion Using Metal-Organic Frameworks and MOF-Based Materials. Chem. Soc. Rev. 2019, 48, 2783–2828. [Google Scholar] [CrossRef]
  10. Kancharlapalli, S.; Snurr, R.Q. High-Throughput Screening of the CoRE-MOF-2019 Database for CO2 Capture from Wet Flue Gas: A Multi-Scale Modeling Strategy. ACS Appl. Mater. Interfaces 2023, 15, 28084–28092. [Google Scholar] [CrossRef]
  11. 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]
  12. Caskey, S.R.; Wong-Foy, A.G.; Matzger, A.J. Dramatic Tuning of Carbon Dioxide Uptake via Metal Substitution in a Coordination Polymer with Cylindrical Pores. J. Am. Chem. Soc. 2008, 130, 10870–10871. [Google Scholar] [CrossRef]
  13. Choe, J.H.; Kim, H.; Hong, C.S. MOF-74 Type Variants for CO2 Capture. Mater. Chem. Front. 2021, 5, 5172–5185. [Google Scholar] [CrossRef]
  14. 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]
  15. Guo, J.; Xue, X.; Yu, H.; Duan, Y.; Li, F.; Lian, Y.; Liu, Y.; Zhao, M. Metal–Organic Frameworks Based on Infinite Secondary Building Units: Recent Progress and Future Outlooks. J. Mater. Chem. A 2022, 10, 19320–19347. [Google Scholar] [CrossRef]
  16. Chen, C.-X.; Wei, Z.; Jiang, J.-J.; Fan, Y.-Z.; Zheng, S.-P.; Cao, C.-C.; Li, Y.-H.; Fenske, D.; Su, C.-Y. Precise Modulation of the Breathing Behavior and Pore Surface in Zr-MOFs by Reversible Post-Synthetic Variable-Spacer Installation to Fine-Tune the Expansion Magnitude and Sorption Properties. Angew. Chem. Int. Ed. 2016, 55, 9932–9936. [Google Scholar] [CrossRef] [PubMed]
  17. Yue, L.; Wang, X.; Lv, C.; Zhang, T.; Li, B.; Chen, D.-L.; He, Y. Substituent Engineering-Enabled Structural Rigidification and Performance Improvement for C2/CO2 Separation in Three Isoreticular Coordination Frameworks. Inorg. Chem. 2022, 61, 21076–21086. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, B.; Zhou, H.-F.; Hou, L.; Wang, Y.-Y. Functionalization of MOFs via a Mixed-Ligand Strategy: Enhanced CO2 Uptake by Pore Surface Modification. Dalton Trans. 2018, 47, 5298–5303. [Google Scholar] [CrossRef]
  19. Hu, Y.; Jiang, Y.; Li, J.; Wang, L.; Steiner, M.; Neumann, R.F.; Luan, B.; Zhang, Y. New-Generation Anion-Pillared Metal–Organic Frameworks with Customized Cages for Highly Efficient CO2 Capture. Adv. Funct. Mater. 2023, 33, 2213915. [Google Scholar] [CrossRef]
  20. Kim, E.J.; Siegelman, R.L.; Jiang, H.Z.H.; Forse, A.C.; Lee, J.-H.; Martell, J.D.; Milner, P.J.; Falkowski, J.M.; Neaton, J.B.; Reimer, J.A.; et al. Cooperative Carbon Capture and Steam Regeneration with Tetraamine-Appended Metal–Organic Frameworks. Science 2020, 369, 392–396. [Google Scholar] [CrossRef]
  21. Liang, W.; Bhatt, P.M.; Shkurenko, A.; Adil, K.; Mouchaham, G.; Aggarwal, H.; Mallick, A.; Jamal, A.; Belmabkhout, Y.; Eddaoudi, M. A Tailor-Made Interpenetrated MOF with Exceptional Carbon-Capture Performance from Flue Gas. Chem 2019, 5, 950–963. [Google Scholar] [CrossRef]
  22. Hu, Z.; Faucher, S.; Zhuo, Y.; Sun, Y.; Wang, S.; Zhao, D. Combination of Optimization and Metalated-Ligand Exchange: An Effective Approach to Functionalize UiO-66(Zr) MOFs for CO2 Separation. Chem. Eur. J. 2015, 21, 17246–17255. [Google Scholar] [CrossRef]
  23. 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]
  24. Myers, A.L.; Prausnitz, J.M. Thermodynamics of Mixed-Gas Adsorption. AIChE J. 1965, 11, 121–127. [Google Scholar] [CrossRef]
  25. Luo, F.; Che, Y.; Zheng, J. Layer-Pillared Metal-Organic Framework Showing Two-Fold Interpenetration and Considerable Solvent-Accessible Channels. Microporous Mesoporous Mater. 2009, 117, 486–489. [Google Scholar] [CrossRef]
  26. Macrae, C.F.; Sovago, I.; Cottrell, S.J.; Galek, P.T.A.; McCabe, P.; Pidcock, E.; Platings, M.; Shields, G.P.; Stevens, J.S.; Towler, M.; et al. Mercury 4.0: From Visualization to Analysis, Design and Prediction. J. Appl. Crystallogr. 2020, 53, 226–235. [Google Scholar] [CrossRef]
  27. Chen, F.; Wang, J.; Guo, L.; Huang, X.; Zhang, Z.; Yang, Q.; Yang, Y.; Ren, Q.; Bao, Z. Carbon Dioxide Capture in Gallate-Based Metal-Organic Frameworks. Sep. Purif. Technol. 2022, 292, 121031. [Google Scholar] [CrossRef]
  28. Nugent, P.; Belmabkhout, Y.; Burd, S.D.; Cairns, A.J.; Luebke, R.; Forrest, K.; Pham, T.; Ma, S.; Space, B.; Wojtas, L.; et al. Porous Materials with Optimal Adsorption Thermodynamics and Kinetics for CO2 Separation. Nature 2013, 495, 80–84. [Google Scholar] [CrossRef]
  29. Zhang, L.; Lin, S.; Liu, Y.; Zeng, X.; You, J.; Xiao, T.; Feng, Y.; He, Z.; Chen, S.; Hua, N.; et al. Optimized Pore Nanospace through the Construction of a Cagelike Metal–Organic Framework for CO2/N2 Separation. Inorg. Chem. 2023, 62, 8058–8063. [Google Scholar] [CrossRef]
  30. Shen, S.; Xu, F.; Chen, X.; Miao, G.; Li, Z.; Zhou, X.; Wang, X. Facile Synthesis of Dptz-CuGeF6 at Room Temperature and Its Adsorption Performance for Separation of CO2, CH4 and N2. Sep. Purif. Technol. 2022, 302, 122054. [Google Scholar] [CrossRef]
  31. Parsaei, M.; Akhbari, K.; White, J. Modulating Carbon Dioxide Storage by Facile Synthesis of Nanoporous Pillared-Layered Metal–Organic Framework with Different Synthetic Routes. Inorg. Chem. 2022, 61, 3893–3902. [Google Scholar] [CrossRef]
  32. 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]
  33. Demessence, A.; D’Alessandro, D.M.; Foo, M.L.; Long, J.R. Strong CO2 Binding in a Water-Stable, Triazolate-Bridged Metal-Organic Framework Functionalized with Ethylenediamine. J. Am. Chem. Soc. 2009, 131, 8784–8786. [Google Scholar] [CrossRef] [PubMed]
  34. McDonald, T.M.; Mason, J.A.; Kong, X.; Bloch, E.D.; Gygi, D.; Dani, A.; Crocellà, V.; Giordanino, F.; Odoh, S.O.; Drisdell, W.S.; et al. Cooperative Insertion of CO2 in Diamine-Appended Metal-Organic Frameworks. Nature 2015, 519, 303–308. [Google Scholar] [CrossRef] [PubMed]
  35. Li, L.; Wang, J.; Zhang, Z.; Yang, Q.; Yang, Y.; Su, B.; Bao, Z.; Ren, Q. Inverse Adsorption Separation of CO2/C2H2 Mixture in Cyclodextrin-Based Metal–Organic Frameworks. ACS Appl. Mater. Interfaces 2019, 11, 2543–2550. [Google Scholar] [CrossRef]
  36. Lyu, H.; Chen, O.I.-F.; Hanikel, N.; Hossain, M.I.; Flaig, R.W.; Pei, X.; Amin, A.; Doherty, M.D.; Impastato, R.K.; Glover, T.G.; et al. Carbon Dioxide Capture Chemistry of Amino Acid Functionalized Metal-Organic Frameworks in Humid Flue Gas. J. Am. Chem. Soc. 2022, 144, 2387–2396. [Google Scholar] [CrossRef]
  37. Choe, J.H.; Kim, H.; Kang, M.; Yun, H.; Kim, S.Y.; Lee, S.M.; Hong, C.S. Functionalization of Diamine-Appended MOF-Based Adsorbents by Ring Opening of Epoxide: Long-Term Stability and CO2 Recyclability under Humid Conditions. J. Am. Chem. Soc. 2022, 144, 10309–10319. [Google Scholar] [CrossRef]
  38. Zhang, L.; He, Z.; Liu, Y.; You, J.; Lin, L.; Jia, J.; Chen, S.; Hua, N.; Ma, L.-A.; Ye, X.; et al. A Robust Squarate-Cobalt Metal–Organic Framework for CO2/N2 Separation. ACS Appl. Mater. Interfaces 2023, 15, 30394–30401. [Google Scholar] [CrossRef]
  39. Wang, Q.; Chen, Y.; Liu, P.; Wang, Y.; Yang, J.; Li, J.; Li, L. CO2 Capture from High-Humidity Flue Gas Using a Stable Metal-Organic Framework. Molecules 2022, 27, 5608. [Google Scholar] [CrossRef]
  40. Nuhnen, A.; Janiak, C. A Practical Guide to Calculate the Isosteric Heat/Enthalpy of Adsorption via Adsorption Isotherms in Metal–Organic Frameworks, MOFs. Dalton Trans. 2020, 49, 10295–10307. [Google Scholar] [CrossRef]
  41. Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  42. Kresse, G.; Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  43. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  44. Blöchl, P.E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [PubMed]
  45. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  46. 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]
Figure 1. (A) The synthesis of PRI-1. (B) PXRD patterns of the simulated and as-synthesized PRI-1, as well as those after various treatments. (C) 195 K CO2 adsorption-desorption isotherms. (D) TGA curve. (E) Comparison of thermal stabilities of PRI-1 and other MOF adsorbents. The distances are in Å.
Figure 1. (A) The synthesis of PRI-1. (B) PXRD patterns of the simulated and as-synthesized PRI-1, as well as those after various treatments. (C) 195 K CO2 adsorption-desorption isotherms. (D) TGA curve. (E) Comparison of thermal stabilities of PRI-1 and other MOF adsorbents. The distances are in Å.
Molecules 28 06276 g001
Figure 2. (A) CO2 and N2 adsorption isotherms of PRI-1 at 273, 283, and 298 K. (B) Qst of CO2 in PRI-1. (C) Comparison of CO2 Qst for PRI-1 with those of other top-performing MOF adsorbents. (D) Comparison of CO2/N2 (15:85, v/v) IAST selectivity at 298 K and 1 bar and Qst of CO2 under ambient conditions in PRI-1 with those of reported CO2-selective MOFs.
Figure 2. (A) CO2 and N2 adsorption isotherms of PRI-1 at 273, 283, and 298 K. (B) Qst of CO2 in PRI-1. (C) Comparison of CO2 Qst for PRI-1 with those of other top-performing MOF adsorbents. (D) Comparison of CO2/N2 (15:85, v/v) IAST selectivity at 298 K and 1 bar and Qst of CO2 under ambient conditions in PRI-1 with those of reported CO2-selective MOFs.
Molecules 28 06276 g002
Figure 3. Calculated preferential binding sites for CO2 (A) and N2 (B) on PRI-1. The distances are in Å. Cu, C, N, O, and H atoms are shown in pink, grey, blue, red, and white, respectively.
Figure 3. Calculated preferential binding sites for CO2 (A) and N2 (B) on PRI-1. The distances are in Å. Cu, C, N, O, and H atoms are shown in pink, grey, blue, red, and white, respectively.
Molecules 28 06276 g003
Figure 4. (A) Experimental breakthrough curves of PRI-1 for CO2/N2 (15:85) binary mixture at 298 K and 1 bar. (B) Experimental cycling breakthrough curves of a CO2/N2 (15:85) binary mixture at 298 K and 1 bar. (C) Experimental breakthrough for CO2/N2 (15:85) under both dry and 75% RH humid conditions. C and C0 stand for outlet concentration and inlet concentration of CO2 and N2 in the gas mixture flow, respectively.
Figure 4. (A) Experimental breakthrough curves of PRI-1 for CO2/N2 (15:85) binary mixture at 298 K and 1 bar. (B) Experimental cycling breakthrough curves of a CO2/N2 (15:85) binary mixture at 298 K and 1 bar. (C) Experimental breakthrough for CO2/N2 (15:85) under both dry and 75% RH humid conditions. C and C0 stand for outlet concentration and inlet concentration of CO2 and N2 in the gas mixture flow, respectively.
Molecules 28 06276 g004
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

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. https://doi.org/10.3390/molecules28176276

AMA Style

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(17):6276. https://doi.org/10.3390/molecules28176276

Chicago/Turabian Style

Li, Yanxi, Yuhua Bai, Zhuozheng Wang, Qihan Gong, Mengchen Li, Yawen Bo, Hua Xu, Guiyuan Jiang, and Kebin Chi. 2023. "Exquisitely Constructing a Robust MOF with Dual Pore Sizes for Efficient CO2 Capture" Molecules 28, no. 17: 6276. https://doi.org/10.3390/molecules28176276

Article Metrics

Back to TopTop