Skip to main content
Nature Communications logoLink to Nature Communications
. 2025 Jul 1;16:5499. doi: 10.1038/s41467-025-61333-9

Fluorine-induced gradient electric field in mesoporous covalent organic frameworks for efficient separation of polarized perfluorinated gases

Mengyao Chen 1,2,#, Yu Zhao 1,2,#, Dan Wen 1,2, Wenkang Gong 1,2, Yuqing Chen 1,2, Yijing Gao 1,2, Yue Yu 1,2, Guolong Xing 1,2, Yuanbin Zhang 2, Weidong Zhu 1,2, Teng Ben 1,2,
PMCID: PMC12216616  PMID: 40595565

Abstract

The challenge of effectively capturing and separating perfluorinated gases is critical due to environmental concerns and the potential to recover these gases as high-value products in the silicon semiconductor industry. Various strategies have been developed to enhance the mass transfer capabilities of microporous adsorbents, which are vital for producing high-performance adsorbent materials. However, slow mass transfer within micropores significantly limits their performance in applications. In this study, we introduce two isostructural mesoporous covalent organic frameworks (COFs) characterized by high crystallinity, porosity, and stability. Compared to non-fluorinated COFs, the highly fluorinated COF demonstrates superior storage capacity for octafluoropropane (C3F8) and perfluorocyclobutane (c-C4F8). It also achieves remarkable separation efficiencies for C3F8 (or c-C4F8)/N2 (or Ar, H2, O2) mixtures under ambient conditions, establishing a standard in the field. This study highlights the significance of strategically modifying pore surface chemistry based on the polarizability differences of guest molecules. Such modifications enable the efficient separation of mixed gas molecules in mesoporous materials.

Subject terms: Polymer synthesis, Porous materials, Pollution remediation


Capturing and separating perfluorinated gases is critical due to their environmental concerns. However, microporous adsorbents show slow mass transfer within micropores limiting their performance in applications. Here, the authors report fluorinated mesoporous covalent organic frameworks for the storage of octafluoropropane and perfluorocyclobutane and separation from other gases mixtures.

Introduction

Global warming is a complex phenomenon resulting from the accumulation of greenhouse gases (GHGs) in the Earth’s atmosphere, and it has garnered significant attention as a major environmental challenge1,2. The primary driver of this warming trend is the excessive use of fossil fuels by humans, leading to elevated concentrations of GHGs, particularly carbon dioxide (CO2) and methane (CH4), which are critical heat-trapping gases3. In recent decades, researchers have made considerable efforts to mitigate global warming by capturing and storing these gases or converting them into valuable chemicals4,5. However, greenhouse effects are influenced by various GHG mixtures. Among these, fluorocarbons are less recognized in mainstream media6,7. Due to the unique properties of fluorine, the carbon−fluorine (C−F) bond ranks among the strongest covalent single bonds in organic chemistry8, with a bond dissociation energy of up to 485 kJ mol−1. Consequently, the prolonged atmospheric lifetimes of perfluorinated gases present significant environmental challenges that require increased attention.

Octafluoropropane (C3F8) and perfluorocyclobutane (c-C4F8) are critical perfluorinated gases widely used in the semiconductor industry. In recent years, plasmas of C3F8 and c-C4F8 have increasingly replaced conventional tetrafluoromethane (CF4) plasmas for the selective dry etching of SiO2 films in ultra-large-scale integrated circuit manufacturing9,10. This transition is attributed to the higher radial density of CFx species in C3F8 and c-C4F8 plasmas, which provides better etching selectivity for SiO2 over Si11,12. However, during practical etching processes, the conversion rates of these perfluorinated gases are notably low, reaching only 30−40%. This inefficiency results in the emission of significant quantities of unutilized C3F8 and c-C4F8, along with other protective gases (e.g., N2, Ar, H2, or O2), causing resource wastage and environmental pollution (Table 1)1315. Furthermore, C3F8 and c-C4F8 are classified as highly potent GHGs, with global warming potentials (GWPs) 8830 and 9540 times greater than CO2, respectively16,17. Projections suggest that the emission rates of these perfluorinated gases will further increase within the next one to two years18. Given their long atmospheric lifetimes, these gases exert a persistent impact on Earth’s radiative balance, necessitating immediate reduction measures. Developing effective recycling technologies to separate, purify, and reuse C3F8 and c-C4F8 is therefore crucial. Such advancements would substantially mitigate environmental pollution and support sustainable development in the semiconductor industry1922.

Table 1.

Binary gas mixtures of C3F8 and c-C4F8 with application scenarios in semiconductor etching processes

Gas mixtures Application scenarios Substrate materials
c-C4F8/N2 Low-k material etching with reduced damage SiO2, porous SiCOH
c-C4F8/Ar High-aspect-ratio dielectric etching SiO2, low-k dielectrics
c-C4F8/O2 Selective etching of Si3N4 Si3N4
C3F8/H2 Selective tungsten (W) etching W, Si, SiO2
C3F8/O2 Bulk etching of poly-Si or metal layers Poly-Si, Aluminum
C3F8/Ar Shallow etching of SiO2 SiO2

In recent years, various technological approaches have been implemented to prevent the release of perfluorinated gases into the atmosphere and to enable their economically feasible capture and reuse23,24. Compared to traditional low-temperature distillation and liquefaction methods, adsorptive separation technology using porous materials offers advantages such as energy savings and high operational feasibility. This approach is an effective strategy for capturing and recovering C3F8 and c-C4F82529. Although adsorptive separation technology is up-and-coming, most of the literature in this field focuses on non-perfluorocarbon gases, with few studies addressing the adsorption and separation of perfluorinated gases using porous materials as adsorbents30, especially for macro-sized perfluorinated compounds.

Covalent organic frameworks (COFs)3135, as a well-known class of crystalline porous organic solids, have experienced rapid development over the past two decades. These materials have found extensive applications in gas storage and separation3638, catalysis3942, energy storage4346, and biomedicine4750, among other fields5155. COFs are constructed by linking carefully designed organic building blocks through reversible covalent bonds, which confer high crystallinity and good thermochemical stability, enabling them to withstand harsh environments. Given the presence of toxic and corrosive impurities such as HF and NOx in fluorine-containing gas emissions, COFs present significant potential as high-performance adsorbents for perfluorinated gas separations.

To enhance the adsorption capacity for perfluorinated gases and improve selectivity over N2, Ar, H2, and O2, we employed electrophilic fluorination to incorporate perfluorinated gas-polarizing groups into pre-synthesized COFs. To the best of our knowledge, the use of COF-based adsorbents for the adsorption and separation of C3F8 (or c-C4F8)/N2 (or Ar, H2, O2) mixtures has not been previously reported. Furthermore, fluorination introduces an induced electric field gradient along the pore surface of COFs due to the high electronegativity of fluorine. This gradient enhances interactions with highly polarizable perfluorinated gases while maintaining minimal interactions with N2, Ar, H2, and O2 (Supplementary Table 1). This differential affinity significantly improves separation performance5658. On the other hand, considering that mesoporosity facilitates rapid and high-throughput mass transfer, the development of mesoporous-based advanced separators is of great significance for practical applications59,60.

In this study, we synthesized two isostructural mesoporous two-dimensional (2D) COFs: fluorinated CPOF-6 and hydrogenated CPOF-7. The introduction of abundant hydrophobic fluorine groups in CPOF-6 enhanced crystallinity, specific surface area, and chemical stability. The uneven distribution of strong electron-withdrawing fluorine atoms in CPOF-6 induces electric field gradients on its pore surface, leading to strong interactions with highly polarizable perfluorinated gases. Compared to CPOF-7, CPOF-6 exhibits superior capture capacities for C3F8 and c-C4F8, along with higher selectivity for C3F8 (or c-C4F8)/N2 (or Ar, H2, O2) mixtures. Breakthrough experiments confirm that CPOF-6 facilitates more efficient mass transfer of C3F8 and c-C4F8 than traditional microporous adsorbents. Further theoretical analysis using density functional theory (DFT) calculations revealed strong interactions between c-C4F8 and the fluorinated frameworks. These findings underscore the importance of surface property modulation in enhancing the gas separation performance of mesoporous adsorbents.

Results and discussion

CPOF-6 and CPOF-7 were synthesized through Schiff base condensation reactions involving 1,3,5-tris(2,3,5,6-tetrafluoroaniline)benzene (TTFAB) or 1,3,5-tris(4-aminophenyl)benzene (TAPB) with 3,3′′,4,4′′,5,5′′-hexafluoro[1,1′:4′,1′′-terphenyl]−2′,5′-dicarboxaldehyde (HFTPDA) or [1,1′:4′,1′′-terphenyl]−2′,5′-dicarbaldehyde (TPDA) in a 1:1.5 molar ratio (Fig. 1a). The formation of imine bonds was confirmed using Fourier-transform infrared (FT-IR) spectroscopy and solid-state carbon-13 cross-polarization/magic angle spinning nuclear magnetic resonance (13C CP/MAS NMR). FT-IR spectra exhibited peaks at 1617 cm−1 and 1620 cm−1, corresponding to the characteristic absorption of C=N bonds in CPOF-6 and CPOF-7, respectively (Supplementary Figs. 5 and 6). Additionally, the absence of stretching vibration bands associated with the vN−H bond of the amine group (3420 cm−1 for TAPB, 3496 cm−1 for TTFAB) and the vC=O bonds of the aldehyde groups (1673 cm−1 for TPDA, 1682 cm−1 for HFTPDA) in the FT-IR spectra indicated the complete conversion of amine and aldehyde groups. The 13C CP/MAS NMR spectra further confirmed the imine bonds formation, with peaks observed at 165 ppm and 153 ppm corresponding to the chemical shifts of carbons in the imine bonds of CPOF-6 and CPOF-7, respectively (Supplementary Figs. 7 and 8). Scanning electron microscopy (SEM) analysis revealed a meshed morphology for both COFs, characterized by entangled nanofibers (Supplementary Figs. 9 and 10). Transmission electron microscopy (TEM) images showed distinct crystalline lattice fringes, indicating an ordered structural arrangement (Supplementary Figs. 11 and 12). Hydrophobicity was evaluated through water contact angle (WCA) measurements. CPOF-6 demonstrated a WCA of approximately 105°, higher than that of CPOF-7 (Supplementary Figs. 13 and 14), suggesting greater hydrophobicity. This observation was further supported by water adsorption experiments, which confirmed the hydrophobic nature of CPOF-6 (Supplementary Figs. 15 and 16). The pronounced hydrophobicity of CPOF-6 can be attributed to the high incorporation of fluorine functional groups within its framework. Notably, the theoretical fluorine content of CPOF-6 is 36.11 wt.%, significantly surpassing that of previously reported fluorinated COFs6164.

Fig. 1. Synthesis routes and structural characterization of 2D COFs.

Fig. 1

a Topology-directed synthesis of imine-linked CPOF-6 and CPOF-7. b The electrostatic potential (ESP) distribution of CPOF-6. The color scale for the ESP map is shown on the right (the gradation on the scale bar is in Hartree/e). Red and blue colors represent the positive and negative parts of ESP, respectively. c The ESP distribution C3F8 and c-C4F8 molecules. The color scale for the ESP map is shown on the right (the gradation on the scale bar is in Hartree/e). d Experimental (black circle) and Pawley refined (red) PXRD patterns of CPOF-6. The difference plot between the observed and the refined PXRD patterns is presented in dark blue. Reflection positions are shown in blue-grey. e PXRD patterns of CPOF-6 after treatments under different chemical environments for 24 h. f Nitrogen adsorption isotherms of CPOF-6 measured at 77 K. Adsorption (closed symbols) and desorption curves (open symbols). Inset: pore size distribution for CPOF-6.

The chemical stability of CPOF-6 and CPOF-7 was evaluated by exposing the samples to organic solvents with varying polarities, such as N,N-dimethylformamide, ethanol, and n-hexane, as well as to aqueous solutions at different pH levels, including 6 M HCl, water, 12 M NaOH, for 24 h. Powder X-ray diffraction (PXRD) analysis revealed that CPOF-6 maintained its crystal structure under all tested conditions, significantly outperforming CPOF-7 (Fig. 1e and Supplementary Figs. 1719). Furthermore, nitrogen adsorption and FT-IR experiments confirmed that the structure of CPOF-6 remained unchanged (Supplementary Figs. 20 and 21). To our knowledge, the chemical stability of CPOF-6 exceeds that of most existing Schiff base COFs. Thermogravimetric analysis (TGA) further showed that both COFs are stable up to 400 °C in a nitrogen atmosphere without decomposition. (Supplementary Figs. 22a and 23a). These results highlight the promising potential of CPOF-6 for practical applications.

The crystal structure and unit cell parameters of the synthesized COFs were determined through PXRD analysis, complemented by structural simulation and Pawley refinement (Fig. 1d). Their potential stacking modes were optimized using the Forcite module of the Materials Studio 7.0 software package. The theoretical PXRD patterns generated based on the AA stacking mode closely matched the experimentally observed PXRD profiles. Consequently, the optimized structures of both COFs were assigned to the P6 space group, with unit cell parameters as follows: for CPOF-6, a = b = 37.250 Å, c = 4.574 Å, α = β = 90°, and γ = 120°; for CPOF-7, a = b = 33.572 Å, c = 4.577 Å, α = β = 90°, and γ = 120°. Based on the experimental PXRD results, full-profile pattern matching (Pawley refinement) was employed to fit the proposed models for CPOF-6 and CPOF-7 (Fig. 1d and Supplementary Figs. 2427), yielding favorable agreement factors (Rp = 4.62% and Rwp = 7.94% for CPOF-6; Rp = 4.94% and Rwp = 6.22% for CPOF-7). Peaks at 2θ = 2.75, 4.76, 5.52, 7.31, and 8.29° for CPOF-6 were attributed to the (100), (110), (200), (210), and (300) facets, respectively. These findings suggest that CPOF-6 and CPOF-7 are likely to possess similar theoretical pore diameters of approximately 2.8 nm in a honeycomb-like network. Notably, the inner surface of CPOF-6 pores is almost entirely occupied by fluorine atoms.

Nitrogen adsorption-desorption isotherms were used to evaluate the permanent porosity of activated CPOF-6 and CPOF-7 at 77 K. As shown in Fig. 1f and Supplementary Figs. 28 and 29, both COFs exhibited characteristic reversible type IV isotherms, confirming their mesoporous nature. The Brunauer−Emmett−Teller (BET) method determined specific surface areas of 1680 m2 g−1 for CPOF-6 and 1054 m2 g−1 for CPOF-7. This notable difference in BET surface areas can be attributed to enhanced interlayer stacking and structural ordering induced by fluorine functional groups, consistent with previous observations in fluorinated porous materials65,66. Pore size distributions, calculated using DFT, revealed pore sizes of 2.8 nm for CPOF-6 and 2.2 nm for CPOF-7, in good agreement with simulation results.

Given the permanent porosity and distinct pore surface environments of the synthesized COFs, we conducted a detailed investigation into their gas adsorption and separation properties. Pure component adsorption isotherms for C3F8, c-C4F8, N2, Ar, H2, and O2 were collected up to 1 bar at 273 and 298 K. Both COFs displayed significantly different adsorption profiles for the perfluorinated gases compared to N2, Ar, H2, and O2, with higher uptake capacities for C3F8 and c-C4F8. Specifically, at 1.0 bar, CPOF-6 exhibited adsorption capacities of 3.37 mmol g−1 and 1.79 mmol g−1 for C3F8 at 273 and 298 K, respectively, and 4.62 mmol g−1 and 3.59 mmol g−1 for c-C4F8 at these temperatures (Fig. 2a, b). These results are higher than those of F-Cage (4.41 mmol g−1 at 273 K and 2.02 mmol g−1 at 298 K)30. In stark contrast, the uptake of N2, Ar, H2, and O2 by CPOF-6 is negligible at both 273 K and 298 K (Supplementary Fig. 30). For CPOF-7, the adsorption capacities for both C3F8 and c-C4F8 were significantly lower than those of CPOF-6 (Fig. 2c, d and Supplementary Table 2), likely due to the hydrogen-occupied pore surfaces and the lower BET surface area of CPOF-7. Nonetheless, the adsorption capacity of CPOF-7 for C3F8 and c-C4F8 remained higher than those for N2, Ar, H2, and O2 (Supplementary Fig. 31).

Fig. 2. Gas sorption properties.

Fig. 2

a N2 and C3F8 adsorption isotherms of CPOF-6 at 273 K (red) and 298 K (blue). b N2 and c-C4F8 adsorption isotherms of CPOF-6 at 273 K (red) and 298 K (blue). c N2 and C3F8 adsorption isotherms of CPOF-7 at 273 K (red) and 298 K (blue). d N2 and c-C4F8 adsorption isotherms of CPOF-7 at 273 K (red) and 298 K (blue). e Calculated isosteric heat of adsorption (Qst) as a function of loading for C3F8, c-C4F8, and N2 on two COFs. f Predicted selectivity for C3F8/N2 (10/90, v/v) and c-C4F8/N2 (10/90, v/v) on two COFs at 298 K and 1 bar.

To determine the binding preferences of each COF for perfluorinated gases relative to N2, Ar, H2, and O2, we calculated the isosteric heats of adsorption (Qst) for C3F8, c-C4F8, N2, Ar, H2, and O2 using the virial fitting method based on the adsorption isotherms (Fig. 2e and Supplementary Figs. 3237). Due to the relatively low adsorption amounts of Ar, H2, and O2, which resulted in unreasonable fitting parameters, these gases are excluded from the analysis. At near-zero coverage, CPOF-6 exhibited Qst values of 37.2 kJ mol−1 for C3F8 and 33.6 kJ mol−1 for c-C4F8, while CPOF-7 showed slightly lower Qst values of 28.7 kJ mol−1 and 28.9 kJ mol−1, respectively. Notably, the Qst values for N2 on CPOF-6 and CPOF-7 were significantly lower, at 9.0 kJ mol−1 and 11.9 kJ mol−1, respectively, compared to those for perfluorinated gases. These results suggest that the introduction of fluorine functional groups through pore surface engineering can effectively enhance the adsorption affinity for perfluorinated gases.

To further evaluate the separation potential of the two COFs for C3F8/N2 and c-C4F8/N2 binary gas mixtures, we calculated adsorption selectivities using ideal adsorbed solution theory (IAST). We fitted single-component adsorption isotherms with single- or dual-site Langmuir-Freundlich (DSLF) models to derive the adsorption parameters required for IAST. As depicted in Fig. 2f and Supplementary Tables 3-10, at 1 bar and 298 K, CPOF-6 exhibited selectivities of 148 for C3F8/N2 (10/90, v/v) and 418 for c-C4F8/N2 (10/90, v/v), significantly higher than the selectivities of 91 for C3F8/N2 (10/90, v/v) and 167 for c-C4F8/N2 (10/90, v/v) observed for CPOF-7. Remarkably, the c-C4F8/N2 (10/90, v/v) selectivity of CPOF-6 represents the highest value reported to date30. We also calculated IAST selectivities for C3F8/N2 (10/90, v/v) and c-C4F8/N2 (10/90, v/v) gas mixtures at 273 K for both COFs. The results, presented in Supplementary Fig. 38 and Supplementary Table 11, confirm that CPOF-6 consistently achieves higher selectivity than CPOF-7 across different temperatures. Consequently, CPOF-6 demonstrates an exceptional combination of high adsorption capacity and selectivity for perfluorinated gases through pore surface fluorination (Supplementary Table 12), setting a potential standard in the design of adsorptive materials.

Building on the high selectivity and strong binding affinity of CPOF-6 for C3F8 and c-C4F8, we conducted an in-depth evaluation of its gas separation performance for C3F8/N2 and c-C4F8/N2 binary mixtures through dynamic breakthrough column experiments. A 10:90 C3F8/N2 or c-C4F8/N2 gas mixture was introduced into an activated CPOF-6 packed column at 298 K with a total flow rate of 10.0 mL min1, using helium (50 vol%) as the carrier gas. Figure 3a, c demonstrate that both gas mixtures were effectively separated in the column, with N2 eluting rapidly. Meanwhile, C3F8 and c-C4F8 achieved breakthrough times of 20.3 min g−1 and 45.2 min g−1, respectively, resulting in substantial intervals of 16.8 min g−1 and 40.6 min g−1 relative to N2. This separation aligns with the differences observed in their adsorption isotherms, indicating that CPOF-6 effectively separates these mixtures under dynamic flow conditions. During the breakthrough process, the total dynamic adsorption capacities of CPOF-6 for C3F8 and c-C4F8 were calculated to be 1.06 mmol g−1 and 2.15 mmol g−1, respectively, close to their saturation adsorption capacities at 298 K and 1 bar. Impressively, CPOF-6 sustained high c-C4F8/N2 separation performance even at 60% relative humidity, highlighting its structural stability and practical viability in humid industrial environments (Supplementary Figs. 40). To assess reusability, multiple-cycle breakthrough tests under identical conditions exhibited stable breakthrough curves for C3F8/N2 and c-C4F8/N2 mixtures after five cycles (Fig. 3e, f), confirming the excellent stability and reusability of CPOF-6. Additionally, breakthrough experiments with various binary mixture ratios were conducted, with results shown in Fig. 3b, d, indicating robust dynamic separation performance across a wide range of perfluorinated gas ratios (Supplementary Figs. 41 and 42). PXRD and N2 adsorption characterization of the used CPOF-6 verified its structural integrity, consistent with the original sample (Supplementary Figs. 43 and 44). For comparison, CPOF-7 underwent identical breakthrough experiments but failed to effectively separate C3F8/N2 and c-C4F8/N2 mixtures (Supplementary Figs. 45 and 46). Since mesoporous COFs lack a size-sieving mechanism for these gases, the superior separation performance of CPOF-6 is primarily attributed to the fluorinated pore surfaces. Thus, CPOF-6 can be regarded as a promising benchmark material for perfluorinated gas purification under room-temperature conditions.

Fig. 3. Breakthrough curves of CPOF-6 for C3F8/N2 and c-C4F8/N2 mixtures at 298 K and 1 bar, using He as the carrier gas.

Fig. 3

a 10/90, v/v and b 20/80, v/v for C3F8/N2. c 10/90, v/v and d 20/80, v/v for c-C4F8/N2. Cycling breakthrough experiments of CPOF-6 for e c-C4F8/N2 (10/90, v/v) and f c-C4F8/N2 (20/80, v/v).

Under the same experimental conditions, we further investigated the gas separation performance of C3F8 (or c-C4F8)/Ar (or H2, O2) binary mixtures on CPOF-6. The dynamic breakthrough curves are shown in Fig. 4. In all experiments, Ar, H2, or O2 breakthrough the packed column first, while C3F8 or c-C4F8 gases showed slower penetration. For example, the breakthrough times for c-C4F8 in the c-C4F8/Ar, c-C4F8/H2, and c-C4F8/O2 systems were 53.9, 29.4, and 51.6 min g−1, respectively, with corresponding breakthrough windows of 50.3, 26.5, and 48.7 min g−1. Similar sequences were observed in the C3F8/Ar and C3F8/O2 systems (Supplementary Fig. 47). These findings highlight the effectiveness of CPOF-6 in separating these mixtures under dynamic flow conditions, emphasizing its potential application in recovering fluorine-containing waste gases from etching processes. Additionally, repeated breakthrough tests confirmed the excellent reproducibility of CPOF-6 (Fig. 4d–f). These results suggest that the fluoride-induced gradient electric field strategy facilitates the efficient separation of perfluorinated gases from N2, Ar, H2, or O2, with the gradient electric field enhancing the interaction with high-polarization gases and having minimal effect on low-polarization gases. Based on our findings, we propose a promising adsorption-based technology for the recovery of fluorinated electronic specialty gases, which could provide a more sustainable alternative to incineration methods. (Supplementary Figs. 48 and 49).

Fig. 4. Breakthrough curves of CPOF-6 for c-C4F8/Ar, c-C4F8/O2, and c-C4F8/H2 mixtures at 298 K and 1 bar, using He as the carrier gas.

Fig. 4

a c-C4F8/Ar (10/90, v/v), b c-C4F8/O2 (10/90, v/v) and c c-C4F8/H2 (10/90, v/v). Cycling breakthrough experiments of CPOF-6 for d c-C4F8/Ar (10/90, v/v), e c-C4F8/O2 (10/90, v/v) and f c-C4F8/H2 (10/90, v/v).

To explore the microstructure of the fluorinated COF and its correlation with the selective adsorption of perfluorinated gases over N2, DFT calculations were performed. The c-C4F8 and N2 mixture served as a model system to identify adsorption sites and binding energies on CPOF-6. As shown in Fig. 5, two primary adsorption sites (sites I and II) are located on the inner surface of CPOF-6 pores. At site I, c-C4F8 binds to the framework through cooperative interactions involving C−H···F−C hydrogen bonds and C−F···F−C halogen bonds, resulting in a calculated static binding energy of 24.6 kJ mol−1. At site II, due to the abundance of positively charged hydrogen atoms, c-C4F8 forms multiple, shorter C−H···F−C hydrogen bonds (2.755−3.011 Å) with the framework, leading to a higher static binding energy of 29.6 kJ mol−1. In contrast, the static binding energies of N2 at these sites are significantly lower, calculated at 6.8 kJ mol−1 for site I and 10.6 kJ mol−1 for site II. These findings suggest a greater affinity of the adsorption sites for c-C4F8, aligning with experimental adsorption isotherms and breakthrough curves. Analysis of the calculated electrostatic potential and DFT results indicates that the significant difference in binding energy can be attributed to charge transfer at the adsorption phase interface, induced by the highly electronegative fluorine groups on the framework. This results in the formation of positive and negative charge regions at the adsorption interface, creating a unique electric field channel that favors the adsorption of highly polarizable c-C4F8 over N2. Based on these findings, we propose a separation mechanism leveraging the electric field gradient effect, whereby channels with electric field gradients can achieve the separation of mixed gases with varying polarizabilities. This mechanism exemplifies an advanced thermodynamically driven separation technology, offering the potential for both large adsorption capacity and high separation efficiency.

Fig. 5. Adsorption configurations and static binding energies of c-C4F8 and N2 in CPOF-6.

Fig. 5

a Optimized configuration of c-C4F8 at Site I showing one C−H···F (3.087 Å) and two C−F···F−C (2.548 Å and 2.819 Å) interactions (ΔE = 24.6 kJ mol−1). b Optimized configuration of c-C4F8 at Site II showing three C−H···F (2.755−3.011 Å) and one C−F···F−C (2.919 Å) interactions (ΔE = 29.6 kJ mol−1). c Adsorption of N2 at Site I showing weak interactions with low static binding energy (ΔE = 6.8 kJ mol−1). d Adsorption of N2 at Site II with slightly stronger binding than Site I (ΔE = 10.6 kJ mol−1), still much weaker than c-C4F8. Color code: C, grey (in CPOF-6) and yellow (in c-C4F8); N, blue; F, green; H, white.

It is important to note that previous studies have reported enhancements in adsorption and separation performance by incorporating fluorine atoms into framework materials, enabling applications such as the capture of perfluorinated pollutants from water and the capture of low-concentration CO2. However, these studies primarily rely on mechanisms such as F−F interactions, hydrophobic effects, anion exchange, or electrostatic attraction6771, which differ fundamentally from the mechanism presented in this work. In addition, conventional gas separation strategies based on size-sieving or kinetic effects typically require precise matching between pore dimensions and molecular sizes of target gases. In contrast, the fluorine-induced gradient electric field mechanism transcends microporous confinement limitations, achieving efficient separation even within mesoporous structures while significantly enhancing mass transfer efficiency. We postulate that this separation paradigm may demonstrate significant potential for broader gas mixture separations, particularly those involving components with substantial disparities in molecular polarizability. This breakthrough establishes a versatile and industrially relevant pathway for high-flux gas separation applications.

Besides the separation performance, the environmental sustainability of the material must also be considered. Perfluorooctanoic acid (PFOA), perfluorooctanesulfonic acid (PFOS), and related per- and polyfluoroalkyl substances (PFAS) have prompted intensified regulatory scrutiny worldwide due to their environmental persistence, bioaccumulative toxicity, and global distribution72,73. Although PFAS-based polymers have not yet been explicitly banned, increasing concerns have emerged regarding their potential to cause serious and irreversible effects throughout their lifecycle. Most current knowledge about the adverse impacts of PFAS pertains to their non-polymeric forms, while the degradation behavior of polymeric PFAS over time remains poorly understood74. Notably, CPOF-6 addresses environmental concerns by enabling the efficient capture and recovery of high-GWP gases (C3F8 and c-C4F8), aligning with sustainability goals. In contrast to non-polymeric PFAS, CPOF-6 is a crystalline, robust framework where fluorine atoms are stably integrated, leading to non-volatility, insolubility, and biological inertness under environmental conditions. This structural stability greatly mitigates risks associated with environmental leaching or degradation. Even at the end of its service life, CPOF-6 can be safely disposed of through established industrial waste management procedures75. Specifically, TGA of CPOF-6 under air was conducted to simulate potential incineration conditions and provide direct insights into its thermal decomposition behavior in oxidative environments. The results showed that decomposition begins at around 390°C, indicating the onset of framework breakdown, and complete degradation occurs near 610°C with negligible residue. These findings confirm that CPOF-6 can be fully decomposed during incineration (Supplementary Figs. 22b and 23b), allowing for safer end-of-life management.

In summary, we have designed and synthesized two isostructural mesoporous COFs. Our findings confirm that the electric field gradient generated by the fluorinated pore surfaces significantly enhances the affinity for perfluorinated gases. This results in notable capture capacities of CPOF-6 for C3F8 (1.79 mmol g−1) and c-C4F8 (3.59 mmol g−1) at 298 K and 1 bar. IAST calculations indicate high separation selectivities for C3F8/N2 (10/90, v/v) and c-C4F8/N2 (10/90, v/v) mixtures, reaching up to 148 and 418 at 298 K and 1 bar, respectively. This positions CPOF-6 as one of the benchmark materials reported to date. The selectivity for c-C4F8/N2 (10/90, v/v) is particularly noteworthy, representing the highest value reported to date. Furthermore, dynamic breakthrough experiments demonstrate that CPOF-6 effectively separates C3F8 (or c-C4F8)/N2 (or Ar, H2, O2) mixtures. This study underscores the effectiveness of tailoring pore surface chemistry by considering the polarizability differences of guest molecules to achieve highly selective and efficient separations using mesoporous materials. Our study paves an alternative pathway for the development of efficient adsorption-based separation technologies.

Methods

Materials

All materials were reagent grade and used as received, unless stated otherwise. 1,3,5-tribromobenzene (Bide Pharmatech Ltd., ≥97.0%), 2,3,5,6-tetrafluoroaniline (Bide Pharmatech Ltd., ≥98.0%), 2,5-dibromo-1,4-benzenedicarboxaldehy (J&K Scientific Ltd., ≥98.0%), 3,4,5-trifluorophenylboronic acid (J&K Scientific Ltd., ≥97.0%), 1,3,5-tris(4-aminophenyl)benzene (Jilin Chinese Academy of Sciences-Yanshen Technology Co., Ltd., ≥98.0%), [1,1′:4′,1′′-terphenyl]-2′,5′-dicarbaldehyde (Jilin Chinese Academy of Sciences-Yanshen Technology Co., Ltd., ≥98.0%), dichloro[1,4-bis(diphenylphosphino)butane]palladium (II) (J&K Scientific Ltd., ≥97.0%), potassium acetate (Bide Pharmatech Ltd., ≥98.0%), N,N-dimethylacetamide (Sinopharm Chemical Reagent Co., Ltd., ≥99.0%), 2,5-dibromo-1,4-benzenedicarboxaldehy (Bide Pharmatech Ltd., ≥97.0%), 3,4,5-trifluorophenylboronic acid (Bide Pharmatech Ltd., ≥98.0%), tetrakis(triphenylphosphine)palladium(0) (J&K Scientific Ltd., ≥99.0%), sodium carbonate (Sinopharm Chemical Reagent Co., Ltd., ≥99.8%), toluene (Sinopharm Chemical Reagent Co., Ltd., ≥99.5%), sodium sulfate (Sinopharm Chemical Reagent Co., Ltd., ≥99.0%), dichloromethane (Sinopharm Chemical Reagent Co., Ltd., ≥99.5%), n-hexane (Sinopharm Chemical Reagent Co., Ltd., ≥97.0%), ethyl acetate (Sinopharm Chemical Reagent Co., Ltd., ≥99.5%), tetrahydrofuran (Sinopharm Chemical Reagent Co., Ltd., ≥99.0%), acetone (Sinopharm Chemical Reagent Co., Ltd., ≥99.0%), HCl 32% (Sinopharm Chemical Reagent Co., Ltd.), acetic acid (Sinopharm Chemical Reagent Co., Ltd., ≥99.5%), NaOH (Sinopharm Chemical Reagent Co., Ltd., ≥96.0%). The solvents were purified and dried according to the standard techniques: N,N-dimethylacetamide was distilled from CaH2. All flash chromatography was performed using silica gel, 60 Å, 300 mesh. TLC analysis was carried out on glass plates coated with silica gel, 0.2 mm thickness. The plates were visualized using a 254 nm ultraviolet lamp.

Instrumentation

1H NMR spectra were measured on a Bruker Fourier 400 or 600 MHz spectrometer. Unless otherwise stated, all spectra were measured at ambient temperature. CP/MAS solid-state 13C-NMR spectra were recorded at ambient pressure on a Bruker Fourier 600 MHz spectrometer using a standard CP pulse sequence probe with 3.2 mm (outside diameter) zirconia rotors at a spinning frequency of 12 kHz. The FTIR spectra (KBr) were obtained using a SHIMADZU IRAffinity−1 Fourier transform infrared spectrophotometer. A SHIMADZU UV−2450 spectrophotometer was used for all absorbance measurements. PXRD patterns were carried out in reflection mode on a Bruker D8 advance powder diffractometer with Cu Kα (λ = 1.5418 Å) line focused radiation at 40 kV and 40 mA from 2θ = 1.0° up to 40° with 0.020481 increments by Bragg−Brentano. The powdered sample was added to the glass and compacted for measurement. TGA was recorded on a SHIMADZU DTG-60 thermal analyzer under nitrogen or air. Nitrogen sorption experiments were performed at 77 K up to 1 bar using a nanometric sorption analyzer. The adsorption-desorption isotherms of N2 were obtained at 77 K using a BELSORP MAX gas sorption analyzer. Gas sorption isotherms of C3F8 and c-C4F8 were obtained from a PhysiChem iPore400 instrument, and the adsorption tube was kept at a constant temperature of 273 K and 298 K by an ethanol bath. Scanning electron microscopy (SEM) was performed on a Zeiss Gemini SEM 300 microscope instrument. Samples were prepared by dispersing the material onto conductive adhesive tapes attached to a flat aluminum sample holder and then coated with gold. High-resolution transmission electron microscopy (HR-TEM) analysis was performed using a JEOL JEM-2100 instrument operated at 200 kV. The synthesized sample was dispersed in ethanol to form a well-dispersed suspension, and a droplet of the suspension was subsequently deposited onto a carbon film-coated TEM grid.

Column breakthrough experiments

Breakthrough experiments were conducted using a custom-built breakthrough column apparatus to evaluate the separation performance of mixed gases. The flow rate of the gas mixture was controlled using a mass flow controller and calibrated with a soap film flowmeter. Pre-activated sample powders were packed into a stainless-steel column (4.65 mm inner diameter, 100 mm length) and subjected to helium (He) purging at 20.0 mL min−1 and 100 °C for 4 h, ensuring complete activation. After cooling to room temperature (298 K), C3F8 (or c-C4F8)/N2 (or Ar, H2, O2) mixtures were introduced at a flow rate of 10.0 mL min−1. The outlet gas composition was monitored using a Pfeiffer GSD 320 mass spectrometer (Germany). After the breakthrough experiment, the sample was regenerated by purging with He at a flow rate of 20 mL min−1 at 100 °C for 4 h to ensure complete removal of adsorbed gases.

Density functional theory calculations

To gain deeper insight into the guest−host interactions at the molecular level, density functional theory (DFT) calculations were performed using the Vienna Ab Initio Simulation Package (VASP, version 5.4.4)7679. The Perdew−Burke−Ernzerhof (PBE) functional within the generalized gradient approximation (GGA)80 was employed to describe the exchange-correlation energy. The projector augmented-wave (PAW)81,82 method was used to represent the interaction between electrons and ions, with a plane-wave cutoff energy of 450 eV. Grimme’s D3 dispersion correction was applied to account for long-range van der Waals interactions83,84. During the geometry optimizations, all atomic positions were fully relaxed until the forces on each atom were smaller than 0.01 eV Å−1, and the electronic self-consistent field (SCF) energy convergence threshold was set to 1.0 × 10−5 eV. Due to the large size of the simulation cell, a 1 × 1 × 1 k-point grid was used for Brillouin zone sampling85. The adsorption energies (Ead) of c-C4F8 and N2 on the CPOF-6 surface were calculated to evaluate the interaction strength between the gas molecules and the framework. The adsorption energy was obtained using the equation:

Ead=EtotalEgasEcof

where Etotal is the total energy of the optimized guest−host complex, Egas is the energy of the isolated gas molecule (either c-C4F8 or N2) in the gas phase, and Ecof is the energy of the pristine CPOF-6 structure.

Synthesis of 1,3,5-tris(2,3,5,6-tetrafluoroaniline)benzene (TTFAB)

1,3,5-tribromobenzene (2.20 g, 7.00 mmol, 1.00 eq.) was added to a Schlenk flask with 4 equiv of 2,3,5,6-tetrafluoroaniline (4.62 g, 28 mmol, 4.00 eq.), along with potassium acetate (1.37 g, 14.00 mmol, 2.00 eq.) and dichloro[1,4-bis(diphenylphosphino)butane]palladium (II) (84.40 mg, 0.14 mmol, 0.02 eq.). The flask was purged with nitrogen, and 21 mL of anhydrous DMAc was added. The solution was heated to 150 °C and stirred rapidly for 20 h. After cooling to r.t., EA (180 mL) was added to the solution, which was then washed with H2O (200 mL × 3) and brine (200 mL × 3) and dried over anhydrous Na2SO4. The solvent was removed, and the residue was purified via column chromatography using EA/n-hexane [4:6] as eluent to afford the product as a light pink solid (1.23 g, 31%). 1H-NMR (400 MHz, DMSO-d6): δH [ppm] = 7.54 (s, 3H), 6.23 (s, 6H). 13C NMR (151 MHz, DMSO-d6): δC [ppm] = 143.80 (dm, J = 240.3 Hz), 135.86 (dm, J = 233.3 Hz), 131.60, 128.50, 128.32, 103.13 (t, J = 17.3 Hz).

Synthesis of 3,3′′,4,4′′,5,5′′-hexafluoro[1,1′:4′,1′′-terphenyl]-2′,5′-dicarboxaldehyde (HFTPDA)

2,5-dibromo-1,4-benzenedicarboxaldehy (0.5 g, 2.00 mmol, 1.00 eq.), 3,4,5-trifluorophenylboronic acid (0.88 g, 5.00 mmol, 2.50 eq.), tetrakis(triphenylphosphine)palladium(0) catalyst (80.0 mg, 0.07 mmol, 0.035 eq.) and toluene (25 mL) were added to a 100 mL Schlenk flask equipped with a magnetic stirrer bar. The mixture was degassed by bubbling N2 for 30 min, at which point degassed aqueous Na2CO3 solution (2 M, 4 mL) was added. The reaction was stirred at 80 °C under an N2 atmosphere for 48 h. After cooling to r.t., 100 mL deionized water was added and extracted with CH2Cl2. The resulting organic extract was washed with water (100 mL × 3) and brine (100 mL × 3), dried over anhydrous Na2SO4, and then concentrated under reduced pressure. The crude solid was recrystallized from CHCl3 to afford the product as a white solid (0.51 g, 75%). 1H-NMR (600 MHz, DMSO-d6): δH [ppm] = 10.03 (s, 2H), 8.01 (s, 2H), 7.64 (dd, J = 8.4, 6.5 Hz, 4H). 13C NMR (151 MHz, DMSO-d6): δC [ppm] = 190.98, 150.90, 149.34, 140.86, 139.95, 138.29, 136.06, 133.20, 131.23, 114.98 (dd, J = 16.9, 4.6 Hz).

Synthesis of CPOF-6

A Pyrex tube measuring o.d. × i.d. = 10 × 8 mm2 was charged with TTFAB (28.4 mg, 0.05 mmol), HFTPDA (29.6 mg, 0.075 mmol) in a mixed solution of DMAc (1.0 mL) and trifluoroacetic acid (0.5 mL). The Pyrex tube was flash-frozen in a liquid nitrogen bath sealed under a vacuum. Upon warming to room temperature, the tube was placed in an oven at 120 °C for 7 days. The yellow solid was isolated by filtration and washed with THF (3 × 15 mL), acetone (3 × 15 mL), and n-hexane (3 × 15 mL). The powder was dried at 80 °C under vacuum overnight to afford the CPOF-6 as a yellow crystalline solid (42.3 mg, Yield: 73%). Elemental analysis for the calculated: C, 58.71%; H, 1.37%; N, 3.81%. Found: C, 57.15%; H, 1.56%; N, 4.14%.

Synthesis of CPOF-7

A Pyrex tube measuring o.d. × i.d. = 10 × 8 mm2 was charged with TAPB (17.6 mg, 0.05 mmol), TPDA (21.5 mg, 0.075 mmol) in a mixed solution of DMAc (1.0 mL) and 6 M aqueous acetic acid (0.2 mL). The Pyrex tube was flash-frozen in a liquid nitrogen bath sealed under a vacuum. Upon warming to room temperature, the tube was placed in an oven at 120 °C for 3 days. The yellow solid was isolated by filtration and washed with THF (3 × 15 mL), acetone (3 × 15 mL), and n-hexane (3 × 15 mL). The powder was dried at 80 °C under vacuum overnight to afford the CPOF-7 as a yellow crystalline solid (29.7 mg, Yield: 76%). Elemental analysis for the calculated: C, 89.26%; H, 4.95%; N, 5.78%. Found: C, 85.56%; H, 4.79%; N, 6.07%.

Supplementary information

Source data

source data (551.2KB, xlsx)

Acknowledgements

This study was supported by the National Key R&D program of China [2021YFA1200400 (T.B.)], the National Natural Science Foundation of China [no. 22475194 (T.B.), 91956108 (T.B.), and 22201256 (Y.Z.)], and the Natural Science Foundation of Zhejiang Province [ZCLMS25B0402 (Y.Z.)].

Author contributions

T.B. and Y.Z. initiated and led the research project. M.C. and D.W. performed the synthesis and characterization of the samples. W.G. and Y.C. conducted the dynamic breakthrough experiments. Y.G. carried out the theoretical calculation. Y.Y., G.X., Y.B.Z., and W.Z. were involved in data analysis and participated in revising the manuscript. T.B., Y.Z., and M.C. organized the results and wrote the paper with discussion and revision of all authors.

Peer review

Peer review information

Nature Communications thanks Heping Ma, Feng He,and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.

Data availability

The data that support the findings detailed in this study are available within the article and supplementary information as well as on figshare: 10.6084/m9.figshare.29108999. All data are available from the corresponding author upon request. Source data are provided with the paper. Source data are provided with this paper.

Competing interests

The authors declare no competing interests.

Footnotes

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

These authors contributed equally: Mengyao Chen, Yu Zhao.

Supplementary information

The online version contains supplementary material available at 10.1038/s41467-025-61333-9.

References

  • 1.Xu, Y. Y., Ramanathan, V. & Victor, D. G. Global warming will happen faster than we think. Nature564, 30–32 (2018). [DOI] [PubMed] [Google Scholar]
  • 2.Al-Ghussain, L. Global warming: review on driving forces and mitigation. Environ. Prog. Sustain. Energy38, 13–21 (2019). [Google Scholar]
  • 3.Trenberth, K. E. et al. Global warming and changes in drought. Nat. Clim. Change4, 17–22 (2014). [Google Scholar]
  • 4.Bui, M. et al. Carbon capture and storage (CCS): the way forward. Energy Environ. Sci.11, 1062–1176 (2018). [Google Scholar]
  • 5.Budinis, S., Krevor, S., Mac Dowell, N., Brandon, N. & Hawkes, A. An assessment of CCS costs, barriers and potential. Energy Strat. Rev.22, 61–81 (2018). [Google Scholar]
  • 6.Anderson, J. G., Toohey, D. W. & Brune, W. H. Free radicals within the antarctic vortex: the role of CFCs in Antarctic ozone loss. Science251, 39–46 (1991). [DOI] [PubMed] [Google Scholar]
  • 7.Chen, T. H. et al. Thermally robust and porous noncovalent organic framework with high affinity for fluorocarbons and CFCs. Nat. Commun.5, 5131 (2014). [DOI] [PubMed] [Google Scholar]
  • 8.Zhang, Z. & Miljanić, O. Š Fluorinated organic porous materials. Org. Mater.01, 019–029 (2019). [Google Scholar]
  • 9.Sasaki, K., Kawai, Y., Suzuki, C. & Kadota, K. Absolute density and reaction kinetics of fluorine atoms in high-density c-C4F8 plasmas. J. Appl. Phys.83, 7482–7487 (1998). [Google Scholar]
  • 10.Jelisavcic, M. et al. Electron scattering from perfluorocyclobutane (c-C4F8). J. Chem. Phys.121, 5272–5280 (2004). [DOI] [PubMed] [Google Scholar]
  • 11.Donnelly, V. M. Review article: reactions of fluorine atoms with silicon, revisited, again. J. Vac. Sci. Technol. A35, 05C202 (2017). [Google Scholar]
  • 12.Knizikevicius, R. Statistical insights into the reaction of fluorine atoms with silicon. Sci. Rep.10, 13634 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Nagai, M., Hayashi, T., Hori, M. & Okamoto, H. Low-k SiOCH film etching process and its diagnostics employing Ar/C5F10O/N2 plasma. Jpn. J. Appl. Phys.45, 7100–7104 (2006). [Google Scholar]
  • 14.Furuya, K. & Hatano, Y. Quantitative analysis of CF4 produced in the SiO2 etching process using c-C4F8, C3F8, and C2F6 plasmas by in situ mass spectrometry. Jpn. J. Appl. Phys.43, 342–346 (2004). [Google Scholar]
  • 15.Samukawa, S. & Mukai, T. High-performance silicon dioxide etching for less than 0.1-μm-high-aspect contact holes. J. Vac. Sci. Technol. B18, 166–171 (2000). [Google Scholar]
  • 16.Ravishankara, A. R., Solomon, S., Turnipseed, A. A. & Warren, R. F. Atmospheric lifetimes of long-lived halogenated species. Science259, 194–199 (1993). [DOI] [PubMed] [Google Scholar]
  • 17.Lindley, A. & McCulloch, A. Regulating to reduce emissions of fluorinated greenhouse gases. J. Fluor. Chem.126, 1457–1462 (2005). [Google Scholar]
  • 18.Muhle, J. et al. Perfluorocyclobutane (PFC-318, c-C4F8) in the global atmosphere. Atmos. Chem. Phys.19, 10335–10359 (2019). [Google Scholar]
  • 19.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. Sust. Energ. Rev.141, 110759 (2021). [Google Scholar]
  • 20.Zhang, W. X. et al. Fluorinated porous organic polymers for efficient recovery perfluorinated electronic specialty gas from exhaust gas of plasma etching. Sep. Purif. Technol.287, 120561 (2022). [Google Scholar]
  • 21.Åhlén, M. et al. Pore size effect of 1,3,6,8-tetrakis(4-carboxyphenyl)pyrene-based metal-organic frameworks for enhanced SF6 adsorption with high selectivity. Microporous Mesoporous Mater.343, 112161 (2022). [Google Scholar]
  • 22.Yang, X. Y. et al. Mesoporous polyimide-linked covalent organic framework with multiple redox-active sites for high-performance cathodic Li storage. Angew. Chem. Int. Ed.61, e202207043 (2022). [DOI] [PubMed] [Google Scholar]
  • 23.Vorotyntsev, V. M. et al. The physico-chemical bases of separation and high purification of fluorocarbons and simple gases. Petrol. Chem.51, 492–495 (2011). [Google Scholar]
  • 24.Stanisch, B., Wellsandt, T. & Strube, J. Development of micro separation technology modules part 2: distillation. Chem. Ing. Tech.87, 1207–1214 (2015). [Google Scholar]
  • 25.Zhang, W. X. et al. Fluorine-induced electric field gradient in 3D porous aromatic frameworks for highly efficient capture of Xe and F-gases. ACS Appl. Mater. Interfaces14, 35126–35137 (2022). [DOI] [PubMed] [Google Scholar]
  • 26.Builes, S., Roussel, T. & Vega, L. F. Optimization of the separation of sulfur hexafluoride and nitrogen by selective adsorption using monte carlo simulations. AIChE J.57, 962–974 (2011). [Google Scholar]
  • 27.Skarmoutsos, I., Eddaoudi, M. & Maurin, G. Highly tunable sulfur hexafluoride separation by interpenetration control in metal organic frameworks. Microporous Mesoporous Mater.281, 44–49 (2019). [Google Scholar]
  • 28.Matito-Martos, I. et al. Zeolites for the selective adsorption of sulfur hexafluoride. Phys. Chem. Chem. Phys.17, 18121–18130 (2015). [DOI] [PubMed] [Google Scholar]
  • 29.Wang, T. G. et al. Calcium-based metal-organic framework for efficient capture of sulfur hexafluoride at low concentrations. Ind. Eng. Chem. Res.60, 5976–5983 (2021). [Google Scholar]
  • 30.Tian, K. et al. Highly selective adsorption of perfluorinated greenhouse gases by porous organic cages. Adv. Mater.34, 2202290 (2022). [DOI] [PubMed] [Google Scholar]
  • 31.Côté, A. P. et al. Porous, crystalline, covalent organic frameworks. Science310, 1166–1170 (2005). [DOI] [PubMed] [Google Scholar]
  • 32.Feng, X., Ding, X. & Jiang, D. Covalent organic frameworks. Chem. Soc. Rev.41, 6010–6022 (2012). [DOI] [PubMed] [Google Scholar]
  • 33.Ding, S. Y. & Wang, W. Covalent organic frameworks (COFs): from design to applications. Chem. Soc. Rev.42, 548–568 (2013). [DOI] [PubMed] [Google Scholar]
  • 34.Wang, Z. F., Zhang, S. N., Chen, Y., Zhang, Z. J. & Ma, S. Q. Covalent organic frameworks for separation applications. Chem. Soc. Rev.49, 708–735 (2020). [DOI] [PubMed] [Google Scholar]
  • 35.Colson, J. W. & Dichtel, W. R. Rationally synthesized two-dimensional polymers. Nat. Chem.5, 453–465 (2013). [DOI] [PubMed] [Google Scholar]
  • 36.Yang, H. et al. Covalent organic framework membranes through a mixed-dimensional assembly for molecular separations. Nat. Commun.10, 2101 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Zhao, Y. et al. Highly selective separation of benzene/cyclohexane by three-dimensional covalent organic framework with 8,8-connected bcu net topology. ACS Mater. Lett.6, 3063–3070 (2024). [Google Scholar]
  • 38.Yu, C. Y. et al. Three-dimensional triptycene-functionalized covalent organic frameworks with hea net for hydrogen adsorption. Angew. Chem. Int. Ed.61, e202117101 (2022). [DOI] [PubMed] [Google Scholar]
  • 39.Zhang, J., Han, X., Wu, X. W., Liu, Y. & Cui, Y. Multivariate chiral covalent organic frameworks with controlled crystallinity and stability for asymmetric catalysis. J. Am. Chem. Soc.139, 8277–8285 (2017). [DOI] [PubMed] [Google Scholar]
  • 40.Chen, D. et al. N-rich 2D heptazine covalent organic frameworks as efficient metal-free photocatalysts. ACS Catal.12, 616–623 (2021). [Google Scholar]
  • 41.Parvatkar, P. T. et al. A tailored COF for visible-light photosynthesis of 2,3-dihydrobenzofurans. J. Am. Chem. Soc.145, 5074–5082 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Basak, A., Karak, S. & Banerjee, R. Covalent organic frameworks as porous pigments for photocatalytic metal-free C-H borylation. J. Am. Chem. Soc.145, 7592–7599 (2023). [DOI] [PubMed] [Google Scholar]
  • 43.Li, W. Q. et al. Thiazolo[5,4-d]thiazole-based donor-acceptor covalent organic framework for sunlight-driven hydrogen evolution. Angew. Chem. Int. Ed.60, 1869–1874 (2021). [DOI] [PubMed] [Google Scholar]
  • 44.Pachfule, P. et al. Diacetylene functionalized covalent organic framework (COF) for photocatalytic hydrogen generation. J. Am. Chem. Soc.140, 1423–1427 (2018). [DOI] [PubMed] [Google Scholar]
  • 45.Xing, G. et al. Nonplanar rhombus and kagome 2D covalent organic frameworks from distorted aromatics for electrical conduction. J. Am. Chem. Soc.144, 5042–5050 (2022). [DOI] [PubMed] [Google Scholar]
  • 46.DeBlase, C. R., Silberstein, K. E., Truong, T. T., Abruña, H. D. & Dichtel, W. R. β-ketoenamine-linked covalent organic frameworks capable of pseudocapacitive energy storage. J. Am. Chem. Soc.135, 16821–16824 (2013). [DOI] [PubMed] [Google Scholar]
  • 47.Zhao, Y. et al. Record ultralarge-pores, low density three-dimensional covalent organic framework for controlled drug delivery. Angew. Chem. Int. Ed.62, e202300172 (2023). [DOI] [PubMed] [Google Scholar]
  • 48.Bhunia, S., Deo, K. A. & Gaharwar, A. K. 2D covalent organic frameworks for biomedical applications. Adv. Funct. Mater.30, 2002046 (2020). [Google Scholar]
  • 49.Bai, L. Y. et al. Nanoscale covalent organic frameworks as smart carriers for drug delivery. Chem. Commun.52, 4128–4131 (2016). [DOI] [PubMed] [Google Scholar]
  • 50.Zhang, L. et al. Engineering multienzyme-mimicking covalent organic frameworks as Pyroptosis inducers for boosting antitumor immunity. Adv. Mater.34, 2108174 (2022). [DOI] [PubMed] [Google Scholar]
  • 51.Zhang, W. W. et al. Reconstructed covalent organic frameworks. Nature604, 72–79 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Gao, C. et al. Isostructural three-dimensional covalent organic frameworks. Angew. Chem. Int. Ed.58, 9770–9775 (2019). [DOI] [PubMed] [Google Scholar]
  • 53.Liu, S. et al. Topochemical cross-linking of diacetylene in a highly interpenetrated three-dimensional covalent organic framework. Chem. Commun.60, 8051–8054 (2024). [DOI] [PubMed] [Google Scholar]
  • 54.Kang, C. J. et al. Covalent organic framework atropisomers with multiple gas-triggered structural flexibilities. Nat. Mater.22, 636–643 (2023). [DOI] [PubMed] [Google Scholar]
  • 55.Wen, D. et al. Catalyst-free solid-state cross-linking of covalent organic frameworks in confined space. Chem. Synth.4, 9 (2024). [Google Scholar]
  • 56.Mandal, S., Natarajan, S., Mani, P. & Pankajakshan, A. Post-synthetic modification of metal-organic frameworks toward applications. Adv. Funct. Mater.31, 2006291 (2021). [Google Scholar]
  • 57.Hong, S. M. et al. Enhanced lithium- and sodium-ion storage in an interconnected carbon network comprising electronegative fluorine. ACS Appl. Mater. Interfaces9, 18790–18798 (2017). [DOI] [PubMed] [Google Scholar]
  • 58.Zhao, Y. F., Yao, K. X., Teng, B. Y., Zhang, T. & Han, Y. A perfluorinated covalent triazine-based framework for highly selective and water-tolerant CO2 capture. Energy Environ. Sci.6, 3684–3692 (2013). [Google Scholar]
  • 59.Lysova, A. A. et al. A series of mesoporous metal-organic frameworks with tunable windows sizes and exceptionally high ethane over ethylene adsorption selectivity. Angew. Chem. Int. Ed.59, 20561–20567 (2020). [DOI] [PubMed] [Google Scholar]
  • 60.Lysova, A. A., Kovalenko, K. A., Nizovtsev, A. S., Dybtsev, D. N. & Fedin, V. P. Efficient separation of methane, ethane and propane on mesoporous metal-organic frameworks. Chem. Eng. J.453, 139642 (2023). [Google Scholar]
  • 61.Alahakoon, S. B., McCandless, G. T., Karunathilake, A. A. K., Thompson, C. M. & Smaldone, R. A. Enhanced structural organization in covalent organic frameworks through fluorination. Chemistry23, 4255–4259 (2017). [DOI] [PubMed] [Google Scholar]
  • 62.Liu, Y. et al. Two-dimensional fluorinated covalent organic frameworks with tunable hydrophobicity for ultrafast oil–water separation. Angew. Chem. Int. Ed.61, e202113348 (2022). [DOI] [PubMed] [Google Scholar]
  • 63.Alahakoon, S. B. et al. Experimental and theoretical insight into the effect of fluorine substituents on the properties of azine-linked covalent organic frameworks. CrystEngComm19, 4882–4885 (2017). [Google Scholar]
  • 64.Chen, J. et al. Fluorine-functionalized 2D covalent organic frameworks with kgd topology for efficient C2H2/CO2 separation. N. J. Chem.47, 6759–6764 (2023). [Google Scholar]
  • 65.Braunecker, W. A. et al. Phenyl/perfluorophenyl stacking interactions enhance structural order in two-dimensional covalent organic frameworks. Cryst. Growth Des.18, 4160–4166 (2018). [Google Scholar]
  • 66.Li, W. B., Cheng, Y. Z., Yang, D. H., Liu, Y. W. & Han, B. H. Fluorine-containing covalent organic frameworks: synthesis and application. Macromol. Rapid Commun.44, 2200778 (2022). [DOI] [PubMed] [Google Scholar]
  • 67.Wang, W., Zhou, S. X., Jiang, X. Z., Yu, G. & Deng, S. B. Fluorinated quaternary ammonium covalent organic frameworks for selective and efficient removal of typical per- and polyfluoroalkyl substances. Chem. Eng. J.474, 145629 (2023). [Google Scholar]
  • 68.Qiu, L. Q. et al. Ionic pairs-engineered fluorinated covalent organic frameworks toward direct air capture of CO2. Small20, 2401798 (2024). [DOI] [PubMed] [Google Scholar]
  • 69.Zhang, K. et al. Fluorinated covalent organic framework-based nanofluidic interface for robust lithium-sulfur batteries. ACS Nano17, 2901–2911 (2023). [DOI] [PubMed] [Google Scholar]
  • 70.Alduraiei, F., Kumar, S., Liu, J. T., Nunes, S. P. & Szekely, G. Rapid fabrication of fluorinated covalent organic polymer membranes for organic solvent nanofiltration. J. Membr. Sci.648, 120345 (2022). [Google Scholar]
  • 71.Gayle, J. et al. Fluorinated 2D conjugated porous organic polymer films with modular structural topology for controlled molecular sieving. J. Mater. Chem. A12, 23023–2303 (2024). [Google Scholar]
  • 72.Brennan, N. M., Evans, A. T., Fritz, M. K., Peak, S. A. & von Holst, H. E. Trends in the regulation of per- and polyfluoroalkyl substances (PFAS): a scoping review. Int. J. Environ. Res. Public Health18, 10900 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Sonne, C. et al. EU need to protect its environment from toxic per- and polyfluoroalkyl substances. Sci. Total Environ.876, 162770 (2023). [DOI] [PubMed] [Google Scholar]
  • 74.European Environment Agency. PFAS polymers in focus: supporting Europe’s zero pollution, low-carbon and circular economy ambitions; https://www.eea.europa.eu/en/analysis/publications/pfas-polymers-in-focus (2025).
  • 75.Améduri, B. & Hori, H. Recycling and the end of life assessment of fluoropolymers: recent developments, challenges and future trends. Chem. Soc. Rev.52, 4208–4247 (2023). [DOI] [PubMed] [Google Scholar]
  • 76.Kresse, G. & Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B47, 558–561 (1993). [DOI] [PubMed] [Google Scholar]
  • 77.Kresse, G. & Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium. Phys. Rev. B49, 14251–14269 (1994). [DOI] [PubMed] [Google Scholar]
  • 78.Kresse, G. & Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B54, 11169 (1996). [DOI] [PubMed] [Google Scholar]
  • 79.Kresse, G. & Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci.6, 15–50 (1996). [DOI] [PubMed] [Google Scholar]
  • 80.Perdew, J. P., Burke, K. & Wang, Y. Generalized gradient approximation for the exchange-correlation hole of a many-electron system. Phys. Rev. B54, 16533–16539 (1996). [DOI] [PubMed] [Google Scholar]
  • 81.Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B59, 1758–1775 (1999). [Google Scholar]
  • 82.Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B50, 17953 (1994). [DOI] [PubMed] [Google Scholar]
  • 83.Grimme, S., Ehrlich, S. & Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem.32, 1456–1465 (2011). [DOI] [PubMed] [Google Scholar]
  • 84.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.132, 154104 (2010). [DOI] [PubMed] [Google Scholar]
  • 85.Monkhorst, H. J. & Pack, J. D. Special points for Brillouin-zone integrations. Phys. Rev. B13, 5188–5192 (1976). [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

source data (551.2KB, xlsx)

Data Availability Statement

The data that support the findings detailed in this study are available within the article and supplementary information as well as on figshare: 10.6084/m9.figshare.29108999. All data are available from the corresponding author upon request. Source data are provided with the paper. Source data are provided with this paper.


Articles from Nature Communications are provided here courtesy of Nature Publishing Group

RESOURCES