Next Article in Journal
Gas-Phase Selective Oxidation of Methane into Methane Oxygenates
Previous Article in Journal
Transition Metal-Catalyzed and MAO-Assisted Olefin Polymerization; Cyclic Isomers of Sinn’s Dimer Are Excellent Ligands in Iron Complexes and Great Methylating Reagents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Zr Modified γ-Al2O3 Catalysts for Stable Hydrolytic Decomposition of CF4 at Low Temperature

1
School of Metallurgy and Environment, Central South University, Changsha 410083, China
2
Chinese National Engineering Research Center for Control & Treatment of Heavy Metal Pollution, Changsha 410083, China
3
School of Chemistry and Chemical Engineering, Central South University, Changsha 410083, China
*
Authors to whom correspondence should be addressed.
Submission received: 8 February 2022 / Revised: 2 March 2022 / Accepted: 7 March 2022 / Published: 9 March 2022
(This article belongs to the Topic Catalysis: Homogeneous and Heterogeneous)

Abstract

:
CF4, one of the Perfluorocompounds (PFCs), also known as a greenhouse gas with high global warming potential. In this study, Zr/γ-Al2O3 catalysts were developed for CF4 decomposition. The addition of Zr onto γ-Al2O3 achieves a high CF4 conversion efficiency of 85% at 650 °C and maintain its activity for more than 60 h, which is obviously higher than that of bare γ-Al2O3 (50%). The mechanism involved in CF4 decomposition over the Zr/γ-Al2O3 are clarified that the surface Lewis acidity sites are the main active center for CF4 directly adsorbing and decomposing. The results of NH3-TPD and FT-IR analyses suggest that the amount of Lewis acidity sites on catalyst surface increases significantly after the introduction of Zr, thereby enhancing the activity of catalyst for CF4 decomposition. The results of XPS analyses confirms the electrons transfer from Zr to Al, which contribute to the increase in Lewis acidity sites. The results of this work will help the development of more effective catalysts for CF4 decomposition.

1. Introduction

Perfluorocompounds (PFCs), such as CF4, have a high global warming potential (GWP), which is about 6500 times higher than that of CO2 over a 100-year time scale. Moreover, the lifetime of PFCs can reach up to 50,000 years [1]. CF4 is an extremely stable molecule because of its symmetry structure. The bond energy of C-F in CF4 molecule is 543 kJ mol−1, as known as the strongest bond in organic chemistry [2,3]. The non-ferrous smelting and semiconductor industry are generally regarded as the main sources of PFCs emissions [4]. Therefore, it is of great importance to remove PFCs presented in the exhaust flue gas from non-ferrous smelting and semiconductor industry.
Several technologies have been developed for CF4 emission control, such as plasma, incineration and catalytic hydrolysis [5,6,7]. Among them the catalytic hydrolysis has been suggested as the most economical choice to decompose CF4. Aluminum-based catalysts have displayed excellent performance in CF4 decomposition due to its high specific surface area and great amount of Lewis acid [8,9]. However, Al2O3 is only active for CF4 decomposition at high temperature, and will fast lose its activity with the prolonging of reaction time. Recently, various methods have been developed to improve the activity of aluminum-based catalyst. El-Bahy et al. [10] researched the CF4 decomposition over Ga-Al oxide catalyst. They revealed that pretreatment with sulfuric acid enhanced the Lewis acid sites, and then improved the activity and lifetime of catalyst. The above results suggest that the amount of Lewis acid sites represent the key property of catalysts for C-F activation.
Some mixed oxides exhibit strong surface acidity due to the generation of excess negative or positive charge in the model structure of the binary oxides [11]. Reddy et al. [12] found that Zr-Al had strong acidic, since the exposed Al3+ and Zr4+ ions could act as Lewis acid sites. Luo et al. [13] reported an efficient Zr-Si oxide catalyst for the synthesis of furanic diesel precursor, and suggested that the addition of Zr species could enhance the formation of Lewis acid sites. Although there are varied Lewis acid type catalysts can be applied in CF4 decomposition, but they often require a high reaction temperature (over 700 °C) to decompose the strong C-F bond. Takita et al. [14] researched the catalytic decomposition of CF4 over metal phosphate-based catalysts including phosphates of Ca, B, Fe, Zn, Bi, and Ni, and found that none of them were active at temperature lower than 700 °C. Xu et al. [15] studied the decomposition of CF4 over γ-Al2O3 modified by several metal elements, and found that the addition of Zn and Ni could enhance the stability of catalysts, but such catalysts were only effective at high temperature (750 °C).
In view of the above reasons, a serious of Zr/Al2O3 catalysts were developed to decompose CF4 at lower temperature, i.e., under 650 °C. The aim of this work is to investigate the precise role of Zr/Al2O3 in CF4 decomposition. The physicochemical properties of catalysts were characterized by different techniques. This study provided in-sights in catalytically decomposing CF4 over Zr/Al2O3.

2. Results and Discussion

2.1. Characterization of Catalysts as Prepared

The elemental composition and distribution of fresh catalysts is measured by SEM-EDS, as Figure 1 and Table 1. The microscopic images of bare γ-Al2O3 (Figure 1a) exhibits the disorder structure consist of irregular blocky particles with rough surface. All the samples containing Zr (Figure 1b–d) present morphology similar to that of alumina. It can be intuitively understood from the EDS element mapping that the existence and uniform distribution of Al and Zr.
To investigate the crystal structure of the modified and bare γ-Al2O3, the X-ray diffraction is performed. The results are shown in Figure 2, that all catalysts contain γ-Al2O3 (JCPDS:10-0425) [16,17]. There are four different polytypes of zirconia (i.e., monoclinic (m-ZrO2), tetragonal (t-ZrO2), cubic (c-ZrO2), and orthorhombic (o-ZrO2)) depend on the temperature and pressure [18]. As for the prepared Zr/γ-Al2O3 catalysts, the peak of t-ZrO2 (JCPDS:80-0965) is observed in XRD patterns [19], indicating the formation of zirconium oxide. The BET surface areas of the samples are listed in Table 2. The specific surface area and pore diameter of catalysts increases after the addition of Zr. The (16%) Zr/Al2O3 shows the largest surface area (132 m2 g−1) and pore diameter (4.1 nm). However, the surface area and pore diameter of Zr/γ-Al2O3 decreases when further increase the amount of Zr to 35%. The BET results are consistent with previous reported, that the low content of Zr highly dispersed on the surface of aluminum oxide, the small particles on the surface provided a large surface area [20,21]. However, only a few Zr ions replaced the Al ion sites during the calcination, because the Zr ion is larger than the Al. As the content of Zr further increase, small particles will agglomerate together to form larger clusters and block the original pore, results in the surface area and pore diameter decrease. It can be anticipated that the catalyst was consist of aluminum oxide and zirconium oxide, they crystallized at the interface during the calcination [22].
The NH3-TPD experiments are conducted to study the acidic properties of catalysts, as shown in Table 2 and Figure 3. The desorbed ammonia is observed at 50–800 °C, according to the different thermal stability of NH3 species, three kinds of acid sites are considered depending on the desorption temperature: weak acidic sites (T < 250 °C), moderate acidic sites (250 °C < T < 400 °C), and strong acidic sites (T > 400 °C) acid sites [23]. These indicate the presence of different acidic sites. The results show that the amounts of three kinds of acidic sites increased with the addition of Zr. The (16%) Zr/γ-Al2O3 has the most amounts of three kinds of acidic sites especially the strong and total acidic sites, which are 2 times larger than that of bare γ-Al2O3. The (9%) Zr/γ-Al2O3 has the second most amounts of weak, strong and total acidic sites. However, the increments of three kinds of acidic sites in (35%) Zr/γ-Al2O3 are very slight. These finding indicates that the addition of Zr mainly increase the amount of strong and total acidic sites, but excess Zr leads to the agglomeration on surface of catalyst, which decreases the surface area and number of acidic sites. It should be noted that, the amounts of acid sites per surface unit have similar trend except the (35%) Zr/γ-Al2O3, but it can be predicted that the catalytic performance of it is limited by its small specific surface area. The above results suggest that the addition of Zr strongly enhances the surface acidity of aluminum oxide. With the addition of Zr to aluminum oxide, it generates excess negative or positive charge in the binary oxides [24]. The exposed Zr4+ and Al3+ can act as Lewis acids, meanwhile, they absorb electrons to generate more Lewis acid centers on the surface of catalysts. Moreover, previous research has reported that the increase of Zr in the Al2O3 catalysts favored the formation of tetra- and, mainly, penta-coordinated Al species, which are considered as potential Lewis acid sites [25], therefore, increase the acidity of catalysts.
In order to gain more information of acidic properties of catalysts, the FT-IR spectra of pyridine adsorption experiments are conducted, as shown in Figure 4. Spectra are recorded over a frequency range of 1400–1700 cm−1, where characteristic vibration modes of adsorbed pyridine will appear. The characteristic peaks at 1450 cm−1, 1491 cm−1, 1577 cm−1 and 1611 cm−1 can be assigned to the Lewis acid sites, whereas no peak of Bronsted acid sites 1545 cm−1 is observed [15,25]. The modified γ-Al2O3 exhibited more Lewis acid sites than the bare γ-Al2O3, especially (16%) Zr/γ-Al2O3, shows the most peak intensity.

2.2. Catalytic Performance of CF4 Decomposition

The conversion reaction of CF4 over catalysts are conducted at 650 °C for 300 min, as shown in Figure 5. The initial conversion over (16%) Zr/γ-Al2O3 reaches 85%, which is much higher than that over bare γ-Al2O3 (50%). Furthermore, the initial conversion over (9%) Zr/γ-Al2O3 and (36%) Zr/γ-Al2O3 are 64% and 48%, respectively. The bare ZrO2 exhibits no catalytic activity. Therefore, (16%) Zr/γ-Al2O3 is chosen as the representative sample, and then the effects of temperature on CF4 conversion over bare γ-Al2O3 and (16%) Zr/γ-Al2O3 are investigated, as shown in Figure 6, the (16%) Zr/γ-Al2O3 catalyst exhibits much better catalytic performance.
The long-term tests are conducted to investigate the stability of catalysts, as shown in Figure 7. The (16%) Zr/γ-Al2O3 catalyst exhibits excellent catalytic activity and stability compares to other works listed in Table 3. With time increasing, the catalytic efficiency of bare γ-Al2O3 decreases rapidly from 30 h (42%) to 60 h (8%) while the efficiency of (16%) Zr/γ-Al2O3 still remains 76%. The large difference between the conversion performances and stabilities in CF4 decomposition are due to the modification by Zr element, which provides more Lewis acidic sites.
The correction between CF4 conversion and amounts of different kind of acidic sites is shown in Figure 8. The CF4 conversion consist with the total and strong acidic sites, since the amount of strong acidic sites dominating over the others, we conclude that the strong acidic sites are the main factor that promotes the activity of the catalysts. Similarly, it has been reported that the Lewis acid sites and total acid sites played a promoting role in the decomposition of hydrofluorocarbons [29,30,31]. Thus, the NH3-TPD and FT-IR results imply that the amount of acid sites on the γ-Al2O3 could be increased by modified with species Zr and then improved the activity and stability of catalysts for CF4 hydrolytic decomposition.

2.3. Mechanism Analysis of Hydrolytic Decomposition of CF4

The close correlation between amount of Lewis acid sites and the conversion of CF4 strongly suggest the direct participation of the acid sites in CF4 decomposition, probably a step for subtracting fluorine atom from CF4 molecule by Lewis acid sites. To understand the mechanisms more detailly, we investigate the physicochemical properties of fresh and used catalysts. The XRD patterns of used catalysts are given in Figure 9. All the used catalysts had the characteristic peaks of α-Al2O3 and AlF3 (JCPDS:84-1672). The peak intensity of α-Al2O3 decreases with the increasing of Zr addition. Moreover, the AlF3 peak intensity of used 16% Zr/γ-Al2O3 is the weakest. SEM-EDS analysis was conducted on the used catalysts as the Table 4. The results showed that the modifier Zr could inhibit the formation of AlF3 on the surface of catalysts, the amount of F on used 16% Zr/Al2O3 catalyst is the lowest in all used catalysts, which showed a good agreement with the stability test. It can be speculated that ZrO2 on the γ-Al2O3 surface prevent the formation of α-Al2O3 and AlF3 to some extent during the CF4 decomposition reaction.
In order to acquire insights into the surface chemical composition of catalysts, X-ray photoelectron spectroscopy (XPS) analyses of catalysts are carried out and shown in Figure 10. The analysis of the Zr 3d core level spectra (Figure 10A) shows that two major binding energy values are located at range 182.9 to 183.1 eV and 185.2 to 185.5 eV, attributing to Zr 3d5/2 and Zr 3d3/2, respectively. These binding energies are higher than that of the bulk ZrO2 due to the higher electronegativity of Al in Zr-O-Al bond, an Al-O- bond around Zr is expected to withdraw more electron density than a Zr-O- bond environment, lead to more positively charges on Zr site [32]. This will increase the amount and strength of Lewis acid sites on catalysts, which shows good agreement with the results of NH3-TPD and FTIR [33,34]. As showed in Figure 10B, for the fresh catalysts, the Al 2p core level spectra exhibit peaks located at 74.3 eV, which can be ascribed to Al-O bond [35,36]. As showed in Figure 10C, for used catalysts, another set of peaks are located in the range of 75.9 to 76.4 eV, which can be attributed to Al-F bond [35,36,37]. Moreover, all the peak intensities of Al-O bond are remarkably decreased with the appearance of peak ascribed to Al-F bond, except 16% Zr/Al2O3 [35,36,37]. It indicates the formation of AlF3 by the reaction between Al2O3 and decomposition product HF, but in the case of 16% Zr/Al2O3 catalyst, species Zr alleviate the formation of AlF3 from Al2O3.

3. Experimental Section

3.1. Catalyst Preparation

The Zr/γ-Al2O3 catalysts were prepared by the homogeneous co-precipitation method. An aqueous solution containing the required amount of zirconium oxychloride (ZrOCl2·8H2O) and Aluminum nitrate (Al(NO3)3·9H2O) were prepared. The solutions were hydrolyzed with dilute ammonium hydroxide under stirring until the pH of the solutions reached to 9, and then subjected to hydrothermal treatment at 90 °C for 24 h. The resulting precipitate was filtered with deionized water several times and dried for 12 h at 100 °C. Finally, the catalysts were calcined at 800 °C for 4 h in N2 atmosphere.

3.2. Catalytic Activity Test

In the aluminum production, the CF4 generated in electrolytic cell during the anode effect, the CF4 content in the flue gas is about 1%. Therefore, we simulate the industrial condition to conduct experiments. The hydrolytic decomposition of CF4 was carried out in a fixed-bed reactor. The temperature was maintained at 650 °C. The gas flow consists of l% CF4, 35% H2O, and balance by Ar. The water was pre-heated at 150 °C and then constantly introduced into the reaction system by using a syringe pump. The effluent gas was washed by aqueous potassium hydroxide to remove the produced HF, and then passed through a cold trap to remove water. Finally, the gas was analyzed by an online gas chromatography (GC-9790 II) equipped with a thermal conductivity detector (TCD) detector.

3.3. Catalyst Characterization

The morphology of the samples was examined by scanning electron microscope (SEM, JSM-IT300LA, JEOL, Japan) with energy dispersive X-ray (EDX) analysis. The X-ray diffraction analysis was performed on a TTR III diffractometer (XRD, Rigaku, Japan). The chemical composition and state of the elements on catalysts surface were investigated by X-ray photoelectron spectroscopy (XPS, EscaLab 250Xi, Thermo Fisher Scientific, USA). The specific surface areas were measured with the N2 adsorption method on an ASAP analyzer (BET, ASAP2020, Micromeritics, USA). Ammonia temperature programmed desorption (NH3-TPD, AutoChem II 2920, Micromeritics, USA) analyses were performed using the following procedures. A 100 mg sample was pretreated at 550 °C, with helium flow of 30 mL/min for 1h, and then cooled to 50 °C. Ammonia (10% NH3/He) was introduced to the catalyst for 1 h at 50 °C. After that, the sample was flushed with helium of 50 mL/min for 1 h to remove absorbed NH3, then the temperature was programmed to increase to 900 °C at a rate of 10 °C/min. The amount of ammonia desorbed from the catalyst was detected using TCD. Fourier transform infrared spectra (FT-IR, Nicolet iS50, Thermo Fisher, USA) of pyridine absorption was conducted using the following procedures. The sample was pressed and put into an IR cell, and then it was degassed at 400 °C under vacuum for 2 h to dehydrate. Next, the cell was cooled to room temperature, and the background signal was recorded. After that, pyridine vapor was introduced to the system until reaching adsorption equilibrium. The sample was evacuated out at 150 °C for 30 min follow by cooling down to 50 °C, and then spectral acquisition was performed.

4. Conclusions

We modified the alumina-based catalysts with Zr to develop an efficient catalyst for CF4 hydrolytic decomposition at 650 °C. The addition of Zr increase the Lewis acid sites on the surface of catalyst. The effects of modification were confirmed by using NH3-TPD and FT-IR of pyridine adsorption methods. We achieved an excellent catalytic CF4 conversion (85%) over modified catalyst. Furthermore, the decomposition efficiency of modified catalyst maintains at 76% after 60 h reaction. These results indicate that the acidity properties of catalysts are strongly improved by the addition of Zr. We also investigated the physicochemical properties of fresh and used catalysts, the results suggested that addition of Zr could inhibit the formation of AlF3.

Author Contributions

All authors contributed to the study conception and design, X.Z. and K.X. developed and designed the methodology of this experiment. X.Z. and F.S. prepared the original draft. H.L. supervised the project and had leadership responsibility for the research. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by National Natural Science Foundation of China (51974378), Innovative Research Group Project of the National Natural Science Foundation of China (52121004) and Special Funding for Innovative Province Construction of Hunan (2019RS3006).

Data Availability Statement

Data is contained within the article.

Conflicts of Interest

The authors declare no competing interests.

References

  1. Mühle, J.; Ganesan, A.L.; Miller, B.R.; Salameh, P.K.; Harth, C.M.; Greally, B.R.; Rigby, M.; Porter, L.W.; Steele, L.P.; Trudinger, C.M.; et al. Perfluorocarbons in the global atmosphere: Tetrafluoromethane, hexafluoroethane, and octafluoropropane. Atmos. Chem. Phys. 2010, 10, 5145–5164. [Google Scholar] [CrossRef] [Green Version]
  2. O’Hagan, D. Understanding organofluorine chemistry. An introduction to the C–F bond. Chem. Soc. Rev. 2008, 37, 308–319. [Google Scholar] [CrossRef] [PubMed]
  3. Hannus, I. Adsorption and transformation of halogenated hydrocarbons over zeolites. Appl. Catal. A Gen. 1999, 189, 263–276. [Google Scholar] [CrossRef]
  4. Pan, K.L.; Chen, Y.S.; Chang, M.B. Effective Removal of CF4 by Combining Nonthermal Plasma with γ-Al2O3. Plasma Chem. Plasma Process. 2019, 39, 877–896. [Google Scholar] [CrossRef]
  5. Chang, M.B.; Lee, H.M. Abatement of perfluorocarbons with combined plasma catalysis in atmospheric-pressure environment. Catal. Today 2004, 89, 109–115. [Google Scholar] [CrossRef]
  6. Setareh, M.; Farnia, M.; Maghari, A.; Bogaerts, A. CF4 decomposition in a low-pressure ICP: Influence of applied power and O2 content. J. Phys. D Appl. Phys. 2014, 47, 355205. [Google Scholar] [CrossRef]
  7. Takita, Y.; Ninomiya, M.; Miyake, H.; Wakamatsu, H.; Yoshinaga, Y.; Ishihara, T. Catalytic decomposition of perfluorocarbons Part II. Decomposition of CF4 over AlPO4-rare earth phosphate catalysts. Phys. Chem. Chem. Phys. 1999, 1, 4501–4504. [Google Scholar] [CrossRef]
  8. Xu, X.-F.; Jeon, J.Y.; Choi, M.H.; Kim, H.Y.; Choi, W.C.; Park, Y.-K. A Strategy to Protect Al2O3-based PFC Decomposition Catalyst from Deactivation. Chem. Lett. 2005, 34, 364–365. [Google Scholar] [CrossRef]
  9. Jeon, J.Y.; Xu, X.-F.; Choi, M.H.; Kim, H.Y.; Park, Y.-K. Hydrolytic decomposition of PFCs over AlPO4–Al2O3catalyst. Chem. Commun. 2003, 2003, 1244–1245. [Google Scholar] [CrossRef]
  10. El-Bahy, Z.M.; Ohnishi, R.; Ichikawa, M. Hydrolysis of CF4 over alumina-based binary metal oxide catalysts. Appl. Catal. B Environ. 2003, 40, 81–91. [Google Scholar] [CrossRef]
  11. Rosenberg, D.J.; Coloma, F.; Anderson, J.A. Modification of the Acid Properties of Silica–Zirconia Aerogels by in situ and ex situ Sulfation. J. Catal. 2002, 210, 218–228. [Google Scholar] [CrossRef]
  12. Reddy, B.M.; Sreekanth, P.M.; Yamada, Y.; Kobayashi, T. Surface characterization and catalytic activity of sulfate-, molybdate- and tungstate-promoted Al2O3–ZrO2 solid acid catalysts. J. Mol. Catal. A Chem. 2005, 227, 81–89. [Google Scholar] [CrossRef]
  13. Luo, Y.-J.; Zhou, Y.-H.; Huang, Y.-B. A New Lewis Acidic Zr Catalyst for the Synthesis of Furanic Diesel Precursor from Biomass Derived Furfural and 2-Methylfuran. Catal. Lett. 2019, 149, 292–302. [Google Scholar] [CrossRef]
  14. Takita, Y.; Morita, C.; Ninomiya, M.; Wakamatsu, H.; Nishiguchi, H.; Ishihara, T. Catalytic Decomposition of CF4 over AlPO4-Based Catalysts. Chem. Lett. 1999, 28, 417–418. [Google Scholar] [CrossRef]
  15. Xu, X.-F.; Jeon, J.Y.; Choi, M.H.; Kim, H.Y.; Choi, W.C.; Park, Y.-K. The modification and stability of γ-Al2O3 based catalysts for hydrolytic decomposition of CF4. J. Mol. Catal. A Chem. 2007, 266, 131–138. [Google Scholar] [CrossRef]
  16. Medeiros, R.L.; Figueredo, G.P.; Macedo, H.P.; Oliveira, Â.A.; Rabelo-Neto, R.C.; Melo, D.M.; Braga, R.M.; Melo, M.A. One-pot microwave-assisted combustion synthesis of Ni-Al2O3 nanocatalysts for hydrogen production via dry reforming of methane. Fuel 2021, 287, 119511. [Google Scholar] [CrossRef]
  17. Romano, P.N.; Filho, J.F.S.D.C.; de Almeida, J.M.A.R.; Sousa-Aguiar, E.F. Screening of mono and bimetallic catalysts for the dry reforming of methane. Catal. Today 2021. [Google Scholar] [CrossRef]
  18. Jomard, G.; Petit, T.; Pasturel, A.; Magaud, L. First-principles calculations to describe zirconia pseudopolymorphs. MRS Online Proc. Libr. 1997, 492, 79–84. [Google Scholar] [CrossRef]
  19. Janampelli, S.; Darbha, S. Highly efficient Pt-MoOx/ZrO2 catalyst for green diesel production. Catal. Commun. 2019, 125, 70–76. [Google Scholar] [CrossRef]
  20. Li, L.; Huo, M.; Zhang, Y.; Li, J. Synthesis of nickel catalysts supported on Zr-doped ordered mesoporous Al2O3 and their catalytic performance for low-temperature CO2 reforming of CH4. J. Porous Mater. 2017, 24, 1613–1625. [Google Scholar] [CrossRef]
  21. Patel, S.; Pant, K.K. Influence of preparation method on performance of Cu(Zn)(Zr)-alumina catalysts for the hydrogen production via steam reforming of methanol. J. Porous Mater. 2006, 13, 373–378. [Google Scholar] [CrossRef]
  22. Song, J.-Y.; Chung, S.-H.; Kim, M.-S.; Seo, M.-G.; Lee, Y.-H.; Lee, K.-Y.; Kim, J.-S. The catalytic decomposition of CF4 over Ce/Al2O3 modified by a cerium sulfate precursor. J. Mol. Catal. A Chem. 2013, 370, 50–55. [Google Scholar] [CrossRef]
  23. Ma, L.; Cheng, Y.; Cavataio, G.; McCabe, R.W.; Fu, L.; Li, J. In situ DRIFTS and temperature-programmed technology study on NH3-SCR of NOx over Cu-SSZ-13 and Cu-SAPO-34 catalysts. Appl. Catal. B Environ. 2014, 156–157, 428–437. [Google Scholar] [CrossRef]
  24. Xia, Y.; Hua, W.; Gao, Z. A new catalyst for n-butane isomerization: Persulfate-modified Al2O3–ZrO2. Appl. Catal. A Gen. 1999, 185, 293–300. [Google Scholar] [CrossRef]
  25. García-Sancho, C.; Jimenez-Gomez, C.P.; Viar-Antunano, N.; Cecilia, J.A.; Moreno-Tost, R.; Merida-Robles, J.M.; Requires, J.; Maireles-Torres, P. Evaluation of the ZrO2/Al2O3 system as catalysts in the catalytic transfer hydrogenation of furfural to obtain furfuryl alcohol. Appl. Catal. A Gen. 2021, 609, 117905. [Google Scholar] [CrossRef]
  26. Jing-Long, H.; Shiue, A.; Yu-Yun, S.; Den-Wei, H.; Chang-Tang, C. Catalytic decomposition of CF4 over copper promoted mesoporous catalysts. Sustain. Environ. Res. 2013, 23, 307–314. [Google Scholar]
  27. Chen, C.-K.; Shiue, A.; Huang, D.-W.; Chang, C.-T. Catalytic Decomposition of CF4 Over Iron Promoted Mesoporous Catalysts. J. Nanosci. Nanotechnol. 2014, 14, 3202–3208. [Google Scholar] [CrossRef]
  28. Xu, X.; Niu, X.; Fan, J.; Wang, Y.; Feng, M. CF4 decomposition without water over a solid ternary mixture consisting of NaF, silicon and one metal oxide. J. Nat. Gas Chem. 2011, 20, 543–546. [Google Scholar] [CrossRef]
  29. Takita, Y.; Tanabe, T.; Ito, M.; Ogura, M.; Muraya, T.; Yasuda, S.; Nishiguchi, H.; Ishihara, T. Decomposition of CH2FCF3 (134a) over Metal Phosphate Catalysts. Ind. Eng. Chem. Res. 2002, 41, 2585–2590. [Google Scholar] [CrossRef]
  30. Swamidoss, C.M.A.; Sheraz, M.; Anus, A.; Jeong, S.; Park, Y.-K.; Kim, Y.-M.; Kim, S. Effect of Mg/Al2O3 and Calcination Temperature on the Catalytic Decomposition of HFC-134a. Catalysts 2019, 9, 270. [Google Scholar] [CrossRef] [Green Version]
  31. Takita, Y.; Moriyama, J.-I.; Nishiguchi, H.; Ishihara, T.; Hayano, F.; Nakajo, T. Decomposition of CCl2F2 over metal sulfate catalysts. Catal. Today 2004, 88, 103–109. [Google Scholar] [CrossRef]
  32. Armelao, L.; Eisenmenger-Sittner, C.; Groenewolt, M.; Gross, S.; Sada, C.; Schubert, U.; Tondello, E.; Zattin, A. Zirconium and hafnium oxoclusters as molecular building blocks for highly dispersed ZrO2 or HfO2 nanoparticles in silica thin films. J. Mater. Chem. 2005, 15, 1838–1848. [Google Scholar] [CrossRef]
  33. García-Sancho, C.; Moreno-Tost, R.; Mérida-Robles, J.; Santamaría-González, J.; Jiménez-López, A.; Maireles-Torres, P. Zirconium doped mesoporous silica catalysts for dehydration of glycerol to high added-value products. Appl. Catal. A Gen. 2012, 433–434, 179–187. [Google Scholar] [CrossRef]
  34. Salas, P.; Wang, J.; Armendariz, H.; Angeles-Chavez, C.; Chen, L. Effect of the Si/Zr molar ratio on the synthesis of Zr-based mesoporous molecular sieves. Mater. Chem. Phys. 2009, 114, 139–144. [Google Scholar] [CrossRef]
  35. Limcharoen, A.; Pakpum, C.; Limsuwan, P. An X-ray Photoelectron Spectroscopy Investigation of Redeposition from Fluorine-based Plasma Etch on Magnetic Recording Slider Head Substrate. Proc. Eng. 2012, 32, 1043–1049. [Google Scholar] [CrossRef] [Green Version]
  36. Kim, S.W.; Park, B.J.; Kang, S.K.; Kong, B.H.; Cho, H.K.; Yeom, G.Y.; Heo, S.; Hwang, H. Characteristics of Al2O3 gate dielectrics partially fluorinated by a low energy fluorine beam. Appl. Phys. Lett. 2008, 93, 191506. [Google Scholar] [CrossRef] [Green Version]
  37. Zhang, T.; Park, J.Y.; Huang, W.; Somorjai, G.A. Influence of reaction with XeF2 on surface adhesion of Al and Al2O3 surfaces. Appl. Phys. Lett. 2008, 93, 141905. [Google Scholar] [CrossRef]
Figure 1. SEM-EDS mapping of (a) bare γ-Al2O3, (b) 9% Zr/γ-Al2O3, (c) 16% Zr/γ-Al2O3, (d) 35% Zr/γ-Al2O3 catalysts.
Figure 1. SEM-EDS mapping of (a) bare γ-Al2O3, (b) 9% Zr/γ-Al2O3, (c) 16% Zr/γ-Al2O3, (d) 35% Zr/γ-Al2O3 catalysts.
Catalysts 12 00313 g001
Figure 2. XRD patterns of catalysts as prepared, (1) γ-Al2O3, (2) 9% Zr/γ-Al2O3, (3) 16% Zr/γ-Al2O3 and (4) 35% Zr/γ-Al2O3.
Figure 2. XRD patterns of catalysts as prepared, (1) γ-Al2O3, (2) 9% Zr/γ-Al2O3, (3) 16% Zr/γ-Al2O3 and (4) 35% Zr/γ-Al2O3.
Catalysts 12 00313 g002
Figure 3. NH3-TPD profiles of catalysts.
Figure 3. NH3-TPD profiles of catalysts.
Catalysts 12 00313 g003
Figure 4. FT-IR spectra of pyridine adsorption on γ-Al2O3 and Zr/γ-Al2O3.
Figure 4. FT-IR spectra of pyridine adsorption on γ-Al2O3 and Zr/γ-Al2O3.
Catalysts 12 00313 g004
Figure 5. CF4 conversion reaction over different loading Zr/γ-Al2O3 catalyst.
Figure 5. CF4 conversion reaction over different loading Zr/γ-Al2O3 catalyst.
Catalysts 12 00313 g005
Figure 6. CF4 conversion reaction over γ-Al2O3 and Zr/γ-Al2O3 catalyst.
Figure 6. CF4 conversion reaction over γ-Al2O3 and Zr/γ-Al2O3 catalyst.
Catalysts 12 00313 g006
Figure 7. CF4 conversion over γ-Al2O3 and Zr/γ-Al2O3 catalyst with time on stream.
Figure 7. CF4 conversion over γ-Al2O3 and Zr/γ-Al2O3 catalyst with time on stream.
Catalysts 12 00313 g007
Figure 8. Correction between CF4 conversion and amounts of different kind of acidic sites.
Figure 8. Correction between CF4 conversion and amounts of different kind of acidic sites.
Catalysts 12 00313 g008
Figure 9. X-ray diffraction (XRD) patterns of the used catalysts, (1) γ-Al2O3, (2) 9% Zr/γ-Al2O3, (3) 16% Zr/γ-Al2O3 and (4) 35% Zr/γ-Al2O3.
Figure 9. X-ray diffraction (XRD) patterns of the used catalysts, (1) γ-Al2O3, (2) 9% Zr/γ-Al2O3, (3) 16% Zr/γ-Al2O3 and (4) 35% Zr/γ-Al2O3.
Catalysts 12 00313 g009
Figure 10. X-ray photo electron spectroscopy (XPS) spectra of catalyst, (A): Zr 3d of fresh, (B): Al 2p of fresh and (C): Al 2p of used. (1) γ-Al2O3, (2) 9% Zr/γ-Al2O3, (3) 16% Zr/γ-Al2O3 and (4) 35% Zr/γ-Al2O3.
Figure 10. X-ray photo electron spectroscopy (XPS) spectra of catalyst, (A): Zr 3d of fresh, (B): Al 2p of fresh and (C): Al 2p of used. (1) γ-Al2O3, (2) 9% Zr/γ-Al2O3, (3) 16% Zr/γ-Al2O3 and (4) 35% Zr/γ-Al2O3.
Catalysts 12 00313 g010
Table 1. Elemental composition of γ-Al2O3 and Zr/γ-Al2O3 catalyst.
Table 1. Elemental composition of γ-Al2O3 and Zr/γ-Al2O3 catalyst.
CatalystAl (wt.%)O (wt.%)Zr (wt.%)
γ-Al2O326.674.4/
(9%) Zr/γ-Al2O339.450.49.0
(16%) Zr/γ-Al2O330.153.016.1
(35%) Zr/γ-Al2O313.550.235.3
Table 2. Specific surface area, pore diameter, NH3-TPD acidic sites and amount of acidic site per surface unit results for the catalyst.
Table 2. Specific surface area, pore diameter, NH3-TPD acidic sites and amount of acidic site per surface unit results for the catalyst.
SamplesBET Surface
Area (m2 g−1)
Pore
Diameter (nm)
Amount of Acid Site (μmol g−1)/
Amount of Acid Site per Surface Unit (μmol g−1 m−2)
WeakModerateStrongTotal
γ-Al2O31193.5235/1.97176/1.48416/3.50827/6.95
(9%) Zr/γ-Al2O31263.7261/2.07191/1.52716/5.621168/9.27
(16%) Zr/γ-Al2O31324.1432/3.27246/1.861168/8.851846/13.98
(35%) Zr/γ-Al2O3892.8254/2.85206/2.31467/5.25927/10.42
Table 3. Comparison of catalytic decomposition of CF4 with other works.
Table 3. Comparison of catalytic decomposition of CF4 with other works.
CatalystsTemperature(°C)Initial CF4 ConversionStabilityRef
16%Zr/Al2O365085%78% after 60 hThis work
γ-Al2O3750100%65% after 20 h[15]
2%Mg/Al2O3750100%60% after 30 h[15]
10%Ce/AlPO470073%55% after 40 h[7]
20%Ce/Al2O365055%45% after 50 h[22]
5%Cu-MCM-4185070%/[26]
5%Fe-MCM-4180081%/[27]
NaF-Si-Al2O385080%20% after 1 h[28]
NaF-Si-MgO850100%40% after 2 h[28]
Table 4. EDS analysis of used catalysts.
Table 4. EDS analysis of used catalysts.
CatalystFluorine (Weight %)
γ-Al2O327.1
(9%) Zr/γ-Al2O323.6
(16%) Zr/γ-Al2O311.2
(35%) Zr/γ-Al2O324.3
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zheng, X.; Xiang, K.; Shen, F.; Liu, H. The Zr Modified γ-Al2O3 Catalysts for Stable Hydrolytic Decomposition of CF4 at Low Temperature. Catalysts 2022, 12, 313. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12030313

AMA Style

Zheng X, Xiang K, Shen F, Liu H. The Zr Modified γ-Al2O3 Catalysts for Stable Hydrolytic Decomposition of CF4 at Low Temperature. Catalysts. 2022; 12(3):313. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12030313

Chicago/Turabian Style

Zheng, Xie, Kaisong Xiang, Fenghua Shen, and Hui Liu. 2022. "The Zr Modified γ-Al2O3 Catalysts for Stable Hydrolytic Decomposition of CF4 at Low Temperature" Catalysts 12, no. 3: 313. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12030313

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop