Skip to main content
Proceedings of the National Academy of Sciences of the United States of America logoLink to Proceedings of the National Academy of Sciences of the United States of America
. 2023 Dec 22;120(52):e2312480120. doi: 10.1073/pnas.2312480120

Promoting C–F bond activation via proton donor for CF4 decomposition

Yingkang Chen a, Wenqiang Qu b, Tao Luo a, Hang Zhang a, Junwei Fu a, Hongmei Li a, Changxu Liu c, Dengsong Zhang b, Min Liu a,1
PMCID: PMC10756256  PMID: 38134197

Significance

Tetrafluoromethane (CF4), also known as carbon tetrafluoride, is a permanent potent greenhouse gas. It finds extensive use in semiconductor manufacturing and is the main by-product released during the smelting of electrolytic aluminum and rare earth metals. Due to the lack of effective treatment means, the concentration of CF4 in the atmosphere has been increasing yearly. While thermal catalytic technology can decompose CF4, the high reaction temperature and low activity restrict its practical applications. Here, we developed a novel and effective strategy for CF4 decomposition, achieving efficient decomposition at a record low reaction temperature. Our findings hold significant promise in the context of global warming, offering practical and impactful solutions to combat the detrimental effects of greenhouse gases.

Keywords: C–F bond activation, greenhouse effect, catalysis, proton donor, CF4 hydrolysis

Abstract

Tetrafluoromethane (CF4), the simplest perfluorocarbons, is a permanently potent greenhouse gas due to its powerful infrared radiation adsorption capacity. The highly symmetric and robust C–F bond structure makes its activation a great challenge. Herein, we presented an innovated approach that efficiently activates C–F bond utilizing protonated sulfate (–HSO4) modified Al2O3@ZrO2 (S-Al2O3@ZrO2) catalyst, resulting in highly efficient CF4 decomposition. By combining in situ infrared spectroscopy tests and density function theory simulations, we demonstrate that the introduced –HSO4 proton donor has a stronger interaction on the C–F bond than the hydroxyl (–OH) proton donor, which can effectively stretch the C–F bond for its activation. Consequently, the obtained S-Al2O3@ZrO2 catalyst achieved a stable 100% CF4 decomposition at a record low temperature of 580 °C with a turnover frequency value of ~8.3 times higher than the Al2O3@ZrO2 catalyst without –HSO4 modification, outperforming the previously reported results. This work paves a new way for achieving efficient C–F bond activation to decompose CF4 at a low temperature.


Global warming has emerged as one of the most pressing concerns in the 21st century (13). Perfluorocarbons (PFCs) have alarmed widespread attention due to their potent greenhouse effects (47). Notably, the Carbon Border Adjustment Mechanism issued by European Union has listed PFCs among the greenhouse gases to be accounted for (8). Among them, tetrafluoromethane (CF4), being the simplest and maximum concentrated PFCs in the atmosphere, possesses a remarkably high global warming potential of 7,390 and extraordinarily long atmospheric lifetime of 50,000 y (914). Considering the doubling of its atmospheric concentration since the preindustrial era (15), the development of a cost-effective CF4 decomposition method becomes crucially important and highly desirable for achieving a sustainable future.

Among various CF4 decomposition methods, the thermocatalytic hydrolysis process stands out due to its improved decomposition rate, no toxic by-products, and large-scale applications potential (1621). While the highly symmetrical single carbon structure and C–F bond with strong ionic character make CF4 decomposition require a high temperature (22), the introduction of the catalyst can significantly promote the C–F bond activation and decrease the decomposition temperature. Takita et al. (16) first proposed that CF4 could be hydrolyzed over Ce10%-AlPO4 at 700 °C, well below its pyrolysis of 1,200 °C (23). El-Bahy et al. (19) found that the introduction of Ga promoted C–F bond activation by increasing Lewis (L) acid sites on γ-Al2O3 surface, achieving a 84% CF4 decomposition at 630 °C. Takashi et al. (24) reported that the Zn-modified γ-Al2O3 catalyst with high density of strong L acid sites could completely hydrolyze CF4 at 650 °C. Our previous study (21) revealed the relationship between the L acid strength of γ-Al2O3 surface site and its CF4 decomposition ability. We found that the tricoordination Al (AlIII) sites with the strong L acidity were the main active site for CF4 decomposition.

Despite the aforementioned progress in reducing temperature, the CF4 hydrolysis at beyond 600 °C still falls short of the desired efficiency for energy utilization. In addition to metal L acid sites (25, 26), Glusker et al. (27) discovered that the proton donor was also capable of interacting with C–F bond to activate it. For example, the hydroxyl (–OH) groups could interact with C–F bond to promote fluorochemical decomposition (2830). The interaction strength with C–F bond was affected by the type of the proton donor (27, 3133). Thus, it could be predicted that the C–F bond of CF4 could be further activated by modulating the proton donor on the catalyst surface.

Herein, we achieve stable CF4 decomposition with an efficiency of 100% operating at 580 °C, under a significantly reduced temperature. Through density function theory (DFT) simulations, we unveil that the introduced protonated sulfate (–HSO4) strongly stretches the C–F bond of adsorbed CF4 and facilitates its activation, which is experimentally verified by in situ infrared spectroscopy (in situ IR) tests. To capitalize on this insight, we design and fabricate a unique catalyst, composed of sulfated Al2O3 dispersed on ZrO2 nanosheet (S-Al2O3@ZrO2). A combination of X-ray photoelectron spectroscopy (XPS) and pyridine-infrared (py-IR) tests confirm the introduction of –HSO4 proton donors on the S-Al2O3@ZrO2 surface, facilitating a record low temperature for CF4 decomposition. This work provides a new strategy for efficient C–F bond activation and CF4 decomposition at low temperature, opening new avenues for sustainable catalysis with environmental benefits and promising energy efficiency.

Results and Discussion

Synthesis and Characterization.

S-Al2O3@ZrO2 was prepared by modifying Al2O3@ZrO2 with sulfuric acid, and calcined at 650 °C for 24 h to remove the excess sulfate species (SI Appendix, Fig. S1 and see Materials and Methods for details). X-ray diffraction (XRD) patterns of S-Al2O3@ZrO2 showed only the characteristic peaks of tetragonal-ZrO2 (PDF #79-1767), which was the same as that of unmodified Al2O3@ZrO2 (Fig. 1A), revealing that sulfuric acid modification did not change the crystal structure of the catalysts. Brunauer–Emmett–Teller (BET) measurements (SI Appendix, Fig. S3 and Table S1) showed that the surface area slightly decreased from 61.3 m2 g−1 (Al2O3@ZrO2) to 49.8 m2 g−1 (S-Al2O3@ZrO2), further confirming no significant changes in its structure properties after sulfuric acid modification.

Fig. 1.

Fig. 1.

(A) XRD patterns of Al2O3@ZrO2 and S-Al2O3@ZrO2 catalysts. (B) TEM and (C) EDS mapping images of S-Al2O3@ZrO2 catalyst. XPS spectra of (D) S 2p, (E) Al 2p, and (F) Zr 3d for Al2O3@ZrO2 and S-Al2O3@ZrO2 catalysts.

To detect the morphologies of the synthesized catalysts, transmission electron microscopy (TEM) was performed. The pristine ZrO2 had a nanoparticle morphology with a size of ~25 nm (SI Appendix, Fig. S4A). Al2O3@ZrO2 showed an identical nanoparticle morphology (SI Appendix, Fig. S4C) as the pristine ZrO2, and its high-resolution TEM (HRTEM) image revealed the presence of Al2O3 (SI Appendix, Fig. S5D), indicating that Al2O3 was uniformly dispersed on ZrO2 surface. After sulfuric acid modification, the morphology of S-Al2O3@ZrO2 had no obvious change (Fig. 1B) compared with that of Al2O3@ZrO2. In contrast, S-Al2O3 showed a significant agglomeration after sulfuric acid modification (SI Appendix, Fig. S4D), indicating that ZrO2 ensured the uniform dispersion of Al2O3 during sulfuric acid modification process. Energy-dispersive spectroscopy (EDS) mapping images of S-Al2O3@ZrO2 (Fig. 1C) further proved the uniform distributions of S and Al on ZrO2 surface.

To investigate the surface chemical environments of the S, O, Al, and Zr in S-Al2O3@ZrO2, XPS was carried out. The O 1s spectra (SI Appendix, Fig. S7) validated the presence of sulfate species on S-Al2O3@ZrO2 after sulfuric acid modification. The sulfate species were further analyzed by S 2p spectra. The peaks at 169.0 eV and 170.3 eV of S-Al2O3@ZrO2 could be attributed to –SO4 and –HSO4 (Fig. 1D), respectively (34), indicating the partial sulfate protonation on the catalyst surface. Meanwhile, the binding sites of –HSO4 on S-Al2O3@ZrO2 surface were investigated by Al 2p and Zr 3d spectra (Fig. 1 E and F). The Al–O peaks (35) of S-Al2O3@ZrO2 were shifted to higher energy region compared to that of Al2O3@ZrO2, which could be attributed to the electron-attracting effect of sulfate group to the surface Al site. However, no observable change could be found in Zr 3d region after sulfuric acid modification (Fig. 1F and SI Appendix, Fig. S10) (36), revealing the sulfate group was not bonded with Zr site. The coordination environment of the Zr sites in S-Al2O3@ZrO2 was precisely characterized by synchrotron X-ray absorption spectroscopy (SI Appendix, Fig. S11A) to further determine the binding sites of –HSO4. The fitting result of Zr k-edge extended X-ray absorption fine structure spectrum only found three scattering paths of Zr–O, Zr–Zr, and Zr–Al (SI Appendix, Fig. S11B and Table S3), demonstrating that the introduced –HSO4 proton donor was mainly bonded with Al site.

Properties of –HSO4 Proton Donor.

To determine the properties of –HSO4 proton donor, the surface acidity of Al2O3@ZrO2 and S-Al2O3@ZrO2 was investigated. First, the acid content and strength of the catalysts were tested by NH3 temperature-programmed desorption (NH3-TPD) (Fig. 2A and SI Appendix, Table S4). Al2O3@ZrO2 showed three NH3 desorption peaks, in which two desorption peaks below 200 °C were attributed to weak acid site, and one desorption peak between 200 °C and 300 °C was assigned to medium acid site (37). The desorption peak intensity and desorption temperature of S-Al2O3@ZrO2 were obviously increased compared with those of Al2O3@ZrO2, indicating a significant increase in acid content and strength via sulfuric acid modification. In particular, not only the medium acid peak of S-Al2O3@ZrO2 was shifted to 300 °C and 400 °C but also the medium acid amount increased from 3.63 μmol g−1 (Al2O3@ZrO2) to 12.11 μmol g−1 (S-Al2O3@ZrO2), demonstrating the S-Al2O3@ZrO2 surface possessed more and stronger acid sites compared to that on Al2O3@ZrO2. The increased strong acid sites were further analyzed by py-IR tests at different desorption temperature (Fig. 2 BD). The observed bands at 1,444 cm−1 and 1,544 cm−1 can be assigned to pyridine adsorbed at L acid sites and Brønsted (B) acid sites, respectively (38). The results showed that Al2O3@ZrO2 contained only L acid sites, and the B acid sites were introduced after sulfuric acid modification, which could be attributed to –HSO4 (39, 40). The strong L acid sites were detected on both Al2O3@ZrO2 and S-Al2O3@ZrO2 catalysts upon desorption at 100 °C (Fig. 2C). Combined with the γ-Al2O3 (110) facet exposure observed in HRTEM results (SI Appendix, Fig. S5 D and E), it could be determined that the detected strong L acid sites were AlIII sites with strong CF4 adsorption. With the desorption temperature rising to 200 °C (Fig. 2D), the band corresponding to L acid sites on Al2O3@ZrO2 and S-Al2O3@ZrO2 disappeared, and the band corresponding to B acid sites was observed only on S-Al2O3@ZrO2, indicating the B acid sites in S-Al2O3@ZrO2 had strong acidity. This result confirmed that the introduced –HSO4 proton donor has a strong proton donating ability.

Fig. 2.

Fig. 2.

(A) NH3-TPD profiles of the Al2O3@ZrO2 and S-Al2O3@ZrO2 catalysts. Py-IR spectra of the Al2O3@ZrO2 and S-Al2O3@ZrO2 at (B) 50 °C, (C) 100 °C, and (D) 200 °C desorption temperature.

Influence of –HSO4 Proton Donor on C–F Bond Activation.

To investigate the influence of –HSO4 proton donor on C–F bond activation, the in situ IR spectra of S-Al2O3@ZrO2 at 580 °C were tested under CF4 atmosphere (Fig. 3A). The bands of CF4 decomposition product HF (3,741, 3,785, 3,832, 3,877, and 3,920 cm−1) (41) were detected, indicating that the C–F bond could be directly activation on S-Al2O3@ZrO2 surface. Meanwhile, the depletion of –HSO4 (1,026 cm−1) (42) and the production of –SO4 (998, 1,072, and 1,142 cm−1) (36, 43) revealed that the C–F bond was activated by –HSO4 and generated –SO4 and HF.

Fig. 3.

Fig. 3.

In situ IR spectra of S-Al2O3@ZrO2 at 580 °C (A) with CF4, or (B) with CF4 and H2O, and intermediate switch to with H2O. (C) The calculation model of γ-Al2O3, γ-Al2O3–OH, and γ-Al2O3–HSO4. (D) CF4 adsorption energies and C–F bond length of γ-Al2O3, γ-Al2O3–OH, and γ-Al2O3–HSO4 models. (E) CF4-TPD profiles of the Al2O3@ZrO2 and S-Al2O3@ZrO2. (F) Schematic illustration of –HSO4 proton donor promoting the C–F bond activation for CF4 decomposition.

To determine the C–F bond activation with –HSO4 during CF4 hydrolysis, the in situ IR spectra of S-Al2O3@ZrO2 at 580 °C were tested under CF4 and H2O atmosphere and only H2O atmosphere, respectively (Fig. 3B). When feeding with CF4 and H2O simultaneously, the production of HF and –SO4 as well as the depletion of –HSO4 was observed, which was similar to the case with CF4 only, indicating –HSO4 directly involved in C–F bond activation with or without H2O. After switching to H2O only (Fig. 3B), it could be found that the –SO4 bands gradually disappeared, while the regeneration of –HSO4 was observed. The results demonstrated that the stable C–F bond activation could be achieved with –HSO4 during CF4 hydrolysis, and –HSO4 could be regenerated by H2O dissociation over S-Al2O3@ZrO2.

The effect of proton donor on C–F bond activation was further investigated by DFT calculations. On the basis of γ-Al2O3 (110) facet containing AlIII site (SI Appendix, Fig. S17), the CF4 adsorption energy at AlIII site (Eads) and C–F bond length were calculated on three different models of γ-Al2O3, γ-Al2O3 with –OH (γ-Al2O3–OH), and γ-Al2O3 with –HSO4 (γ-Al2O3–HSO4), respectively (Fig. 3C). The calculation results showed that the Eads of γ-Al2O3–HSO4 was −0.64 eV, which was much stronger than those of γ-Al2O3–OH (−0.51 eV) and γ-Al2O3 (−0.44 eV) (Fig. 3D), indicating that the CF4 adsorption was significantly enhanced by introducing –HSO4. The CF4-temperature programed desorption (CF4-TPD) results of Al2O3@ZrO2 and S-Al2O3@ZrO2 (Fig. 3E) also proved this result, as the CF4 desorption temperature of S-Al2O3@ZrO2 was significantly increased after sulfuric acid modification. On the other hand, the C–F bond length on γ-Al2O3–HSO4 increased from 1.36 Å to 1.45 Å, which was larger than that on γ-Al2O3 (1.42 Å) and γ-Al2O3–OH (1.43 Å) (Fig. 3D). The enhanced adsorption and molecular deformation of CF4 proved that –HSO4 can promote the C–F bond activation. Based on these results, a synergistic mechanism for C–F bond activation was proposed (Fig. 3F). The AlIII site stably securely CF4 molecule, while the adjacent –HSO4 proton donor generates the H···C–F interaction with the adsorbed CF4. The synergistic stretching effect of AlIII-HSO4 pair sites on CF4 promotes the C–F bond activation.

CF4 Hydrolysis Performance.

In order to determine the promotion of –HSO4 for C–F bond activation, the CF4 hydrolysis performances were tested under 440 to 660 °C (Fig. 4A and SI Appendix, Fig. S20). The S-Al2O3@ZrO2 catalyst with abundant –HSO4 showed excellent CF4 hydrolysis activity and could stably achieve 100% CF4 decomposition at the temperature (T100) of 580 °C (Fig. 4B), which was much lower than that of Al2O3@ZrO2 (660 °C) and other previously reported catalysts (SI Appendix, Table S5), proving the great promotion of C–F bond activation by the introduction of –HSO4 proton donor. Similarly, the CF4 hydrolysis test results on Al2O3 (640 °C) and S-Al2O3 (620 °C) also demonstrated this promoted effect. The S-Al2O3@ZrO2 catalyst after stability test was then characterized (SI Appendix, Figs. S21−S24 and Tables S7 and S8). It could be found that the surface –HSO4 proton donor content remained almost unchanged. The key effect of –HSO4 in promoting C–F bond activation was further demonstrated.

Fig. 4.

Fig. 4.

(A) Plots of CF4 decomposition versus temperature over ZrO2, Al2O3, S-Al2O3, Al2O3@ZrO2, and S-Al2O3@ZrO2 catalysts. (B) The stability test of S-Al2O3@ZrO2 catalyst at 580 °C. (C) The Arrhenius plots and (D) reaction rate of Al2O3, S-Al2O3, Al2O3@ZrO2, and S-Al2O3@ZrO2 catalysts for CF4 decomposition. Reaction condition: 2,500 ppm of CF4 and 10% of H2O balanced with Ar, total flow rate of 33.3 mL min−1, and weight hourly space velocity (WHSV) of 1,000 mL g−1 h−1.

Combined with the above DFT simulations and previous report (21), the AlIII site is the main active site for CF4 decomposition. The turnover frequency (TOF) values at 500 °C were calculated according to the content of the surface AlIII site determined by the results of the surface acidity test (SI Appendix, Table S4). The results demonstrated the significantly enhanced of the catalyst intrinsic activity after introducing –HSO4 proton donor, and the S-Al2O3@ZrO2 catalyst presented the highest TOF of 3.91 × 10−3 s−1, which was ~8.3 times as that of Al2O3@ZrO2 (0.47 × 10−3 s−1). Further, the activation energies (Ea) of CF4 over different catalysts (Fig. 4C) was evaluated by the CF4 hydrolysis reaction rates (Fig. 4D). The results showed that the Ea of both S-Al2O3 and S-Al2O3@ZrO2 were significantly reduced compared with those before sulfuric acid modification, demonstrating the activation of CF4 molecules by –HSO4. Especially, the S-Al2O3@ZrO2 had the lowest Ea of 86.5 kJ mol−1, corresponding to its efficient CF4 decomposition.

Conclusion

In summary, a strategy to activate the C–F bond for effective CF4 decomposition has been proposed by enhancing the interaction between –HSO4 proton donor and C–F bond. The S-Al2O3@ZrO2 catalyst with abundant –HSO4 exhibited an excellent CF4 hydrolysis activity, in which CF4 was completely decomposed at a record low temperature of 580 °C with a TOF value ~8.3 times that of the Al2O3@ZrO2 catalyst with –HSO4. It was demonstrated that the enhanced interaction between the introduced –HSO4 proton donor and the C–F bond was decisive to enhance CF4 hydrolysis activity, verified by in situ IR test and DFT simulation. –HSO4 had a stronger interaction with C–F bond than that of –OH could effectively activate the C–F bond, thus promoting CF4 decomposition. This proposed strategy paves a way for the development of efficient C–F bond activation and CF4 decomposition catalyst. As global warming continues to be a pressing concern, our findings open new avenues for practical and impactful solutions to combat the detrimental effects of greenhouse gases.

Materials and Methods

Materials.

All chemicals were obtained commercially and used as received. Zirconium oxychloride octahydrate (ZrOCl2⋅8H2O, 99%) and aluminum nitrate nonahydrate (Al(NO3)3·9H2O, 99%) were purchased from Aladdin. Aqueous ammonia (NH3·H2O) and sulfuric acid (H2SO4) were purchased from Sinopharm.

Preparation of Al2O3@ZrO2.

Al2O3@ZrO2 was synthesized using the hydrothermal method. First, 30 mM (9.7 g) ZrOCl2·8H2O and 3.6 mM (1.4 g) Al(NO3)3·9H2O were dissolved in 100 mL of deionized water by stirring at room temperature, and NH3·H2O was added till pH~10. Then, the above solution was transferred into a 150-mL Teflon-lined stainless-steel autoclave and kept at 150 °C for 24 h. After cooling to room temperature, the precipitate was washed with deionized water, dried under 60 °C overnight, and calcined in air at 600 °C (a heating rate of 3 °C min−1) for 5 h. The pristine Al2O3 and ZrO2 samples were obtained by the same method with only adding aluminum source and zirconium source, respectively.

Preparation of S-Al2O3@ZrO2.

First, 2.0 g Al2O3@ZrO2 and 2.0 g Al2O3 samples were added to 10 mL of 1 M H2SO4 and dried overnight at 60 °C, and then calcined in air at 650 °C (a heating rate of 3 °C min−1) for 24 h to obtain S-Al2O3@ZrO2 and S-Al2O3, respectively.

Catalyst Characterization.

XRD patterns were obtained by using a STADIP automated transmission diffractometer, operated at 36 kV and 20 mA by using CuKa1 radiation. The XRD patterns were scanned in the 2θ range of 15 to 90°.

The TEM and EDS images were obtained by JEOL 3010 operated at 200 kV. The finely ground sample was dispersed in ethanol and then dropped onto a copper grid for TEM and EDX testing.

XPS was recorded on a Kratos Axis Ultra DLD X-ray photoelectron spectrometry, using a standard Al Ka X-ray source and an analyzer pass energy of 40 eV. All binding energies were referenced to the adventitious C 1s line at 284.6 eV.

The BET surface area and pore size distribution of the catalysts were determined by N2 adsorption-desorption analysis using AUTOSORB IQ. Prior to measurements, the samples were degassed at 300 °C for 6 h, at a heating rate of 10 °C min−1.

NH3-TPD and CF4-TPD were performed by using a PCA-1200 on a chemisorption analyzer equipped with a thermal conductivity detector (TCD). The chemisorption analyzer was carried out on the PCA-1200 from Beijing Builder electronic technology Co., Ltd. For each experiment, the weighed sample (100 mg) was pretreated at 600 °C (10 °C min−1) for 2 h under Ar (30 mL min−1) and cooled to room temperature. Then, the NH3 gas (30 mL min−1) or 20% CF4/Ar gas (30 mL min−1) was introduced instead of Ar at this temperature for 1 h to ensure the saturation adsorption. The sample was then purged with Ar for 1 h (30 mL min−1) until the signal returned to the baseline as monitored by a TCD. The desorption curve of NH3 or CF4 was acquired by heating the sample from room temperature to 800 °C (10 °C min−1) under Ar with the flow rate of 30 mL min−1.

Py-IR spectra of samples were analyzed by a Thermo IS-50 Fourier Transform infrared (FTIR) spectrometer. The sample was heated at 500 °C for 5 h and cooled to room temperature. Then, vacuumized to 10−3 Torr, samples were exposed to pyridine vapour (3,000 Pa) at 100 °C for 1 h, followed by reevacuation for 1 h, and lower the temperature to take out our samples. After this step, the sample was analyzed by FTIR.

In situ IR spectra of the sample were also analyzed by a Thermo IS-50 FTIR spectrometer. Self-supported wafer was prepared from catalyst powder (ca. 10 mg). The wafer was loaded into an in situ IR thermal catalytic cell with CaF2 windows and pretreated under Ar flow at 600 °C for 2 h. Then regulated to the target temperature to obtain a background spectrum which should be deducted from the sample spectra. As for the transient reactions between 1) CF4 and water vapor and 2) water vapor, after the background spectra at appointed temperatures under Ar flow were obtained, the catalysts were exposed to 1) 1 mL min−1 20% CF4/Ar + water vapor (50 mL Ar passing through water bottle) or 2) water vapor (50 mL Ar passing through water bottle) at 580 °C and meanwhile the reaction process was recorded as a function of time. For the adsorption of CF4 studies, after the same pretreatment, the catalysts were exposed to a flow of 20% CF4/Ar at 100 °C for 2 h. The desorption process then went on under a flow of Ar, and the temperature was gradually raised to 700 °C (a step of 50 °C) and recorded as a function of temperature.

Catalytic Activity Evaluation and Analytical Methods.

CF4 hydrolysis was measured by using a continuous flow reaction system with a quartz fixed-bed reactor (i.d. 20 mm) under atmospheric pressure in a temperature range from 440 to 660 °C. A gas flow of 33.3 mL min−1 (0.25% CF4 in Ar) controlled by a mass flow controller, together with 0.2 mL h−1 of water pumped by an injection pump, were passed over 2.0 g of catalyst. According to the following equations, the CF4 decomposition is calculated:

CF4 decomposition (%)=CF4in-[CF4]outCF4in×100%,

where [CF4]in and [CF4]out indicate the input and output relative gas concentrations, respectively.

Theoretical Calculation Studies.

All our investigations in this study were based on density functional theory, as implemented in the Vienna ab initio simulation package (44, 45). The exchange-correlation potential is treated with the Perdew–Burke–Ernzerhof formula by using the projected augmented wave method within the generalized gradient approximation (46). The cut-off energy for all calculations was set to be 450 eV. All the positions of atoms are fully relaxed until the Hellmann–Feynman forces on each atom are less than 0.01 eV Å−1. Meanwhile, a k-points Γ-centered mesh is generated for Brillouin zone samples. The DFT-D3 method proposed by Grimme was adopted to describe the van der Waals interactions, which has been shown to accurately describe chemisorption and physisorption properties on layered material. In addition, a vacuum region of about 15 Å was used to decouple the periodic replicas.

Supplementary Material

Appendix 01 (PDF)

Acknowledgments

We thank the National Natural Science Foundation of China (Grant Nos. 22376222, 22002189, 52202125, 52372253, and 22125604), Science and Technology Innovation Program of Hunan Province (Grant No. 2023RC1012), Central South University Research Programme of Advanced Interdisciplinary Studies (Grant No. 2023QYJC012), Central South University Innovation-Driven Research Programme (Grant No. 2023CXQD042). We would like to acknowledge the help from Beam Lines BL01C1 in the National Synchrotron Radiation Research Center (Hsinchu, Taiwan) for various synchrotron-based measurements. We are grateful for technical support from the High Performance Computing Center of Central South University.

Author contributions

Y.C., D.Z., and M.L. designed research; Y.C. and H.Z. performed research; Y.C., T.L., H.Z., J.F., and M.L. contributed new reagents/analytic tools; Y.C., W.Q., T.L., J.F., H.L., D.Z., and M.L. analyzed data; and Y.C., W.Q., J.F., C.L., D.Z., and M.L. wrote the paper.

Competing interests

The authors declare no competing interest.

Footnotes

This article is a PNAS Direct Submission.

Data, Materials, and Software Availability

All study data are included in the article and/or SI Appendix.

Supporting Information

References

  • 1.Bera P. P., Francisco J. S., Lee T. J., Design strategies to minimize the radiative efficiency of global warming molecules. Proc. Natl. Acad. Sci. U.S.A. 107, 9049–9054 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Bolin B., Kheshgi H. S., On strategies for reducing greenhouse gas emissions. Proc. Natl. Acad. Sci. U.S.A. 98, 4850–4854 (2001). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Nauels A., et al. , Attributing long-term sea-level rise to Paris Agreement emission pledges. Proc. Natl. Acad. Sci. U.S.A. 116, 23487–23492 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.McMillan C. A., Keoleian G. A., Not all primary aluminum is created equal: Life cycle greenhouse gas emissions from 1990 to 2005. Environ. Sci. Technol. 43, 1571–1577 (2009). [DOI] [PubMed] [Google Scholar]
  • 5.Saito T., Yokouchi Y., Stohl A., Taguchi S., Mukai H., Large emissions of perfluorocarbons in East Asia deduced from continuous atmospheric measurements. Environ. Sci. Technol. 44, 4089–4095 (2010). [DOI] [PubMed] [Google Scholar]
  • 6.Gschrey B., Schwarz W., Elsner C., Engelhardt R., High increase of global F-gas emissions until 2050. GHG Meas. Manage. 1, 85–92 (2011). [Google Scholar]
  • 7.Choi S., Hong S. H., Lee H. S., Watanabe T., A comparative study of air and nitrogen thermal plasmas for PFCs decomposition. Chem. Eng. J. 185, 193–200 (2012). [Google Scholar]
  • 8.Europäische Union, Regulation (EU) 2023/956 of the European Parliament and of the council of 10 May 2023 establishing a carbon border adjustment mechanism. Off. J. Eur. Union 66, 130–152 (2023). [Google Scholar]
  • 9.Yang S., Zhang H., Zou Z., Li J., Zhong X., Reducing PFCs with local anode effect detection and independently controlled feeders in aluminum reduction cells. JOM 72, 229–238 (2020). [Google Scholar]
  • 10.Wofford B. A., Jackson M. W., Hartz C., Bevan J. W., Surface wave plasma abatement of CHF3 and CF4 containing semiconductor process emissions. Environ. Sci. Technol. 33, 1892–1897 (1999). [Google Scholar]
  • 11.Harnisch J., Borchers R., Fabian P., Kourtidis K., Aluminium production as a source of atmospheric carbonyl sulfide (COS). Environ. Sci. Pollut. Res. 2, 161–162 (1995). [DOI] [PubMed] [Google Scholar]
  • 12.Hildebrand J. H., Lamoreaux R. H., Diffusivity of gases in liquids. Proc. Natl. Acad. Sci. U.S.A. 71, 3321–3324 (1974). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Cicerone R. J., Atmospheric carbon tetrafluoride: A nearly inert gas. Science 206, 59–61 (1979). [DOI] [PubMed] [Google Scholar]
  • 14.Worton D. R., et al. , Atmospheric trends and radiative forcings of CF4 and C2F6 inferred from firn air. Environ. Sci. Technol. 41, 2184–2189 (2007). [DOI] [PubMed] [Google Scholar]
  • 15.Hockstad L., Hanel L., “Inventory of US greenhouse gas emissions and sinks” in Environmental System Science Data Infrastructure for a Virtual Ecosystem (ESS-DIVE) (United States) (Environmental Protection Agency, 2018). [Google Scholar]
  • 16.Takita Y., et al. , Catalytic decomposition of perfluorocarbons part II. Decomposition of CF4 over AlPO4-rare earth phosphate catalysts. Phys. Chem. Chem. Phys. 1, 4501–4504 (1999). [Google Scholar]
  • 17.Xu X., et al. , The modification and stability of γ-Al2O3 based catalysts for hydrolytic decomposition of CF4. J. Mol. Catal. A: Chem. 266, 131–138 (2007). [Google Scholar]
  • 18.Song J., et al. , The catalytic decomposition of CF4 over Ce/Al2O3 modified by a cerium sulfate precursor. J. Mol. Catal. A: Chem. 370, 50–55 (2013). [Google Scholar]
  • 19.El-Bahy Z. M., Ohnishi R., Ichikawa M., Hydrolytic decomposition of CF4 over alumina-based binary metal oxide catalysts: High catalytic activity of gallia-alumina catalyst. Catal. Today 90, 283–290 (2004). [Google Scholar]
  • 20.Jong Y. J., Xiu-Feng X., Mi H. C., Hee Y. K., Yong-Ki P., Hydrolytic decomposition of PFCs over AlPO4-Al2O3 catalyst. Chem. Commun., 1244–1245 (2003). [DOI] [PubMed] [Google Scholar]
  • 21.Zhang H., et al. , Identification of the active site during CF4 hydrolytic decomposition over γ-Al2O3. Environ. Sci.: Nano 9, 954–963 (2022). [Google Scholar]
  • 22.Anus A., Sheraz M., Jeong S., Kim E., Kim S., Catalytic thermal decomposition of tetrafluoromethane (CF4): A review. J. Anal. Appl. Pyrol. 156, 105126 (2021). [Google Scholar]
  • 23.Erguang H., Zheng H., Shukun W., Liyong X., Mengna B., Thermal decomposition and interaction mechanism of HFC-227ea/n-hexane as a zeotropic working fluid for organic Rankine cycle. Energy 246, 123435 (2022). [Google Scholar]
  • 24.Sasaki T., Kanno S., Role of acid sites on γ-Al2O3 based catalyst for CF4 hydrolysis. Kag. Kog. Ronbunshu 37, 175–180 (2011). [Google Scholar]
  • 25.Jia W., et al. , Influence of Lewis acidity on catalytic activity of the porous alumina for dehydrofluorination of 1,1,1,2-tetrafluoroethane to trifluoroethylene. Catal. Lett. 145, 654–661 (2015). [Google Scholar]
  • 26.Stahl T., Klare H.-F., Oestreich M., Main-group Lewis acids for C-F bond activation. ACS Catal. 3, 1578–1587 (2013). [Google Scholar]
  • 27.Rust M., Stallings W. C., Preston R. K., Intermolecular interactions of the C-F bond: The crystallographic environment of fluorinated carboxylic acids and related structures. J. Am. Chem. Soc. 105, 3206–3214 (1983). [Google Scholar]
  • 28.Von R. S. P., Kos A. J., The importance of negative (anionic) hyperconjugation. Tetrahedron 39, 1141–1150 (1983). [Google Scholar]
  • 29.Howard J. A. K., Hoy V. J., O’Hagan D., Smith G. T., How good is fluorine as a hydrogen bond acceptor? Tetrahedron 52, 12613–12622 (1996). [Google Scholar]
  • 30.Cormanich R. A., Rittner R., Freitas M. P., Bühl M., The seeming lack of CF⋅HO intramolecular hydrogen bonds in linear aliphatic fluoroalcohols in solution. Phys. Chem. Chem. Phys. 16, 19212–19217 (2014). [DOI] [PubMed] [Google Scholar]
  • 31.Champagne P. A., Desroches J., Paquin J. F., Organic fluorine as a hydrogen-bond acceptor: Recent examples and applications. Synthesis 47, 306–322 (2015). [Google Scholar]
  • 32.Champagne P. A., Benhassine Y., Desroches J., Paquin J. F., Friedel-crafts reaction of benzyl fluorides: Selective activation of C-F bonds as enabled by hydrogen bonding. Angew. Chem. Int. Ed. Engl. 53, 13835–13839 (2014). [DOI] [PubMed] [Google Scholar]
  • 33.Nepal B., Scheiner S., Anionic CH⋅⋅⋅X hydrogen bonds: Origin of their strength, geometry, and other properties. Chem. Eur. J. 21, 1474–1481 (2015). [DOI] [PubMed] [Google Scholar]
  • 34.Pârvulescu V., Coman S., Grange P., Pârvulescu V. I., Preparation and characterization of sulfated zirconia catalysts obtained via various procedures. Appl. Catal. A: Gen. 176, 27–43 (1999). [Google Scholar]
  • 35.Makarowicz A., et al. , Electronic structure of lewis acid sites on high surface area aluminium fluorides: A combined XPS and ab initio investigation. Phys. Chem. Chem. Phys. 11, 5664–5673 (2009). [DOI] [PubMed] [Google Scholar]
  • 36.Wang H., et al. , A stable mesoporous super-acid nanocatalyst for eco-friendly synthesis of biodiesel. Chem. Eng. J. 364, 111–122 (2019). [Google Scholar]
  • 37.Chen Y., et al. , Alumina catalyst with high density of medium acid sites for N-arylpyrrolidines synthesis. Appl. Catal. A: Gen. 640, 118668 (2022). [Google Scholar]
  • 38.Zhang H., Yang W., Roslan I. I., Jaenicke S., Chuah G., A combo Zr-HY and Al-HY zeolite catalysts for the one-pot cascade transformation of biomass-derived furfural to γ-valerolactone. J. Catal. 375, 56–67 (2019). [Google Scholar]
  • 39.Dominguez J. M., Hernandez J. L., Sandoval G., Surface and catalytic properties of Al2O3–ZrO2 solid solutions prepared by sol–gel methods. Appl. Catal. A: Gen. 197, 119–130 (2000). [Google Scholar]
  • 40.Ahmed A. I., El-Hakam S. A., Khder A. S., El-Yazeed W. S. A., Nanostructure sulfated tin oxide as an efficient catalyst for the preparation of 7-hydroxy-4-methyl coumarin by Pechmann condensation reaction. J. Mol. Catal. A: Chem. 366, 99–108 (2013). [Google Scholar]
  • 41.Mosiadz M., Juda K. L., Hopkins S. C., Soloducho J., Glowacki B. A., An in-depth in situ IR study of the thermal decomposition of barium trifluoroacetate hydrate. Thermochim. Acta 513, 33–37 (2011). [Google Scholar]
  • 42.Xiao-Fei L., et al. , Sb4O3(TeO3)2(HSO4)(OH): An antimony tellurite sulfate exhibiting large optical anisotropy activated by lone pair stereoactivity. Inorg. Chem. 62, 7123–7129 (2023). [DOI] [PubMed] [Google Scholar]
  • 43.Lazaroff N., Sigal W., Wasserman A., Iron oxidation and precipitation of ferric hydroxysulfates by resting thiobacillus ferrooxidans cells. Appl. Environ. Microbiol. 43, 924–938 (1982). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Kresse G., Hafner J., Ab initio molecular dynamics for liquid metals. Phys. Rev. B 47, 558 (1993). [DOI] [PubMed] [Google Scholar]
  • 45.Kresse G., Furthmüller J., Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 54, 11169 (1996). [DOI] [PubMed] [Google Scholar]
  • 46.Perdew J. P., Burke K., Ernzerhof M., Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996). [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

Appendix 01 (PDF)

Data Availability Statement

All study data are included in the article and/or SI Appendix.


Articles from Proceedings of the National Academy of Sciences of the United States of America are provided here courtesy of National Academy of Sciences

RESOURCES