Next Article in Journal
Polymer-Assisted Biocatalysis: Polyamide 4 Microparticles as Promising Carriers of Enzymatic Function
Previous Article in Journal
Ru-Catalyzed Repetitive Batch Borylative Coupling of Olefins in Ionic Liquids or Ionic Liquids/scCO2 Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Sulfuric Acid Treatment on the Performance of Ga-Al2O3 for the Hydrolytic Decomposition of 1,1,1,2-Tetrafluoroethane (HFC-134a)

1
Department of Chemical and Biological Engineering, Korea University, Anam-ro 145, Seongbuk-gu, Seoul 02841, Korea
2
Korea Institute of Energy Research, 152 Gajeong-ro, Yuseong-gu, Daejeon 34129, Korea
3
School of Environmental Engineering, University of Seoul, Seoul 02504, Korea
4
Graduate School of Energy Science and Technology Chungnam National University, 99 Daehak-ro, Yuseong-gu, Daejeon 305-764, Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(7), 766; https://doi.org/10.3390/catal10070766
Submission received: 8 June 2020 / Revised: 30 June 2020 / Accepted: 7 July 2020 / Published: 9 July 2020
(This article belongs to the Section Environmental Catalysis)

Abstract

:
HFC-134a, one of the representative hydrofluorocarbons (HFCs) used as a coolant gas, is a known greenhouse gas with high global warming potential. Catalytic decomposition is considered a promising technology for the removal of fluorinated hydrocarbons. However, systematic studies on the catalytic decomposition of HFC-134a are rare compared to those for other fluorinated hydrocarbon gases. In this study, Ga-Al2O3 and S/Ga-Al2O3 catalysts were prepared and the change in their properties post-acid treatment was investigated by X-ray diffraction (XRD), Brunauer-Emmett-Teller (BET), temperature-programmed desorption of ammonia (NH3-TPD), in situ Fourier-transform infrared spectroscopy (FT-IR), scanning electron microscopy combined with energy-dispersive X-ray spectroscopy (SEM-EDS), and X-ray photoelectron spectroscopy (XPS). The S/Ga-Al2O3 catalyst achieved a much higher HFC-134a conversion than Ga-Al2O3, which was ascribed to the promotional effect of the sulfuric acid treatment on the Lewis acidity of the catalyst surface, as confirmed by NH3-TPD. Furthermore, the effect of hydrogen fluoride (HF) gas produced by HFC-134a decomposition on the catalyst was investigated. The S/Ga-Al2O3 maintained a more stable and higher HFC-134a conversion than Ga-Al2O3. Combining the results of the stability test and characterization, it was established that the sulfuric acid treatment not only increased the acidity of the catalyst but also preserved the partially reduced Ga species.

Graphical Abstract

1. Introduction

Chlorofluorocarbons (CFCs) and hydrochlorofluorocarbons (HCFCs) are two classes of coolants, which have been found to directly contribute to the destruction of the stratospheric ozone layer [1,2]. The Montreal protocol in 1987 banned the use of these coolants, and hydrofluorocarbons (HFCs) were developed to replace them [1,2]. With the increase in use of air conditioning, the concentration of HFCs in the atmosphere has risen significantly [1,2,3]. HFCs do not deplete the ozone layer, but as greenhouse gases, their global warming potential is ~12,000 times higher than that of CO2 [1,3]. HFC-134a is one of the most commonly used HFC refrigerants today, and measures to remove it from the atmosphere are urgently required to prevent global warming [1].
Several technologies, such as thermal decomposition, plasma, and catalytic decomposition, have been investigated for HFC-134a removal [4,5,6,7,8,9]. Among these methods, catalytic decomposition is considered promising because it can be carried out at lower temperature than other methods [7,8,9]. Izuka et al. experimentally compared thermal and catalytic decompositions [9]. For thermal decomposition, the initiating temperature for decomposition of HFC-134a was higher than 750 °C, and complete conversion was obtained at 900 °C [5,9]. On the other hand, using waste cement as a catalyst, 100% conversion of HFC-134a was achieved at about 600 °C [9].
Various catalysts, such as waste concrete, supported catalysts, and metal phosphate catalysts, have been investigated [7,8,9,10]. Alumina (Al2O3)-based catalysts have been commonly applied for the decomposition of HFC-134a because Al2O3 is inexpensive and a representative acid catalyst [8,10]. Han et al. reported that an Al2O3-based catalyst exhibits a very high activity and showed a higher stability when using water as a hydrogen donor. Swamidoss et al. tested the catalytic decomposition of HFC-134a over Mg-supported Al2O3 catalysts [8]. They found that the Mg/Al2O3 catalyst calcined at 650 °C has a higher amount of weak acid sites, an important factor for HFC-134a decomposition [8]. Song et al. tested CF4 decomposition over metal-supported Al2O3 and elucidated that modification of the catalyst by metal impregnation preserves its active sites [10]. They found that using a metal-sulfate precursor could further enhance the catalytic performance by increasing the acid sites [10]. Takita et al. investigated metal sulfate catalysts for CCl2F2 decomposition [11]. The authors insisted that metal oxides were not stable for CCl2F2 decomposition, due to weak resistance to HF, while metal sulfate catalysts, especially Zr(SO4)2, achieved complete conversion over 350 °C in the presence of water vapor [11]. Previous research on the use of acid-treated catalysts for the decomposition of other fluorinated hydrocarbons suggest that the catalytic efficiency and stability of Al2O3-based catalysts can be increased by acid treatment, but there has been little systematic investigation on using alumina-based catalysts for HFC-134a decomposition [8,10,11,12,13].
In this study, Ga-Al2O3 and S/Ga-Al2O3 catalysts were prepared to investigate the change in the properties of the catalyst on acid treatment. Furthermore, the effect of the HF gas produced by HFC-134a decomposition on the catalyst was investigated.

2. Results and Discussion

2.1. Improvement in Catalytic Performance in HFC-134a Decomposition

Pristine Ga-Al2O3 and sulfuric acid-treated Ga-Al2O3 catalysts were synthesized and tested for HFC-134a decomposition reaction. To ensure the elemental composition of as-prepared catalysts, the amounts of Ga and Al were estimated by inductively coupled plasma, and that of S was measured by an elemental analyzer, given in Table 1. Figure 1 shows the temperature dependence of HFC-134a conversion over Ga-Al2O3 and S/Ga-Al2O3 catalysts. The catalysts exhibited markedly different performances. S/Ga-Al2O3, having a small amount of H2SO4 loading (1 wt.% of S), exhibited a higher HFC-134a conversion (90.5% at 450 °C) than Ga-Al2O3 (62% at 450 °C).
It has been reported that large amounts of HF molecules are inevitably produced during HFC-134a decomposition, which negatively affects the catalyst performance because of halogenide formation on the catalyst surface (Reaction (1)) [7]. In particular, the activity of the alumina-based catalyst is remarkably decreased by formation of AlF3 (Reaction (2)) [7,14]. Thus, it is necessary to observe the catalyst stability during the HFC-134a decomposition reaction.
C H 2 F C F 3 + H 2 O + 3 2 O 2     2 C O 2 + 4 H F
A l 2 O 3 + 6 H F   3 H 2 O + 2 A l F 3
The inset of Figure 1 presents the results of the catalyst stability test. It reveals that with time on stream, HFC-134a conversion over Ga-Al2O3 decreased much faster than that over S/Ga-Al2O3, retaining ~40% and 83% after 30 h, respectively. As both catalysts used the same amount of Ga (15 wt.%), it could be said that the large difference and good stability in HFC-134a decomposition performance are likely due to the pretreatment with sulfuric acid [14].
It has been reported that the catalytic properties such as crystallinity, surface area, and acidity are drastically influenced by pretreatment with sulfuric, hydrofluoric, nitric, and phosphoric acids [10,15,16]. XRD analysis was performed to confirm the crystal structure of our catalysts. Figure 2 presents the XRD patterns of Ga-Al2O3 and S/Ga-Al2O3 catalysts, revealing that both catalysts contained γ-Al2O3 (JCPDS #29-63) [17,18]. No peaks of Ga2O3 were observed in any case, which was ascribed to the high dispersion of Ga or the formation of Ga nanoparticles [17]. Therefore, only the γ-Al2O3 phase was detected by XRD [17,18]. The absence of sulfate-related peaks was attributed to the good dispersion of these species on the catalyst surface [19]. As shown in Table 2, Ga-Al2O3 and S/Ga-Al2O3 had BET surface areas of 227.5 and 187.4 m2 g−1 and total pore volumes of 0.35 and 0.29 m3 g−1, respectively.
When H2SO4 is doped in the mixed oxide, it generates acid sites on the catalyst [14,19]. Moreover, as sulfate ions are Lewis acids, they attract electrons to create new Lewis acid sites that could further improve the catalytic performance for HFC-134a decomposition [14,19]. Temperature-programmed desorption of ammonia (NH3-TPD) and in situ FT-IR analysis were conducted to observe the acidic strength and type of surface acidity on Ga-Al2O3 and S/Ga-Al2O3 catalysts. Figure 3 presents the NH3-TPD profiles of the two catalysts recorded at 55–700 °C. According to desorption temperature T, the sites could be grouped into those with weak (T < 250 °C), medium (250 °C < T < 400 °C), and strong (400 °C < T) acid sites, which implied the presence of sites with different acidic strengths [10,20]. Sulfuric acid treatment increased the amount of weak and medium acid sites, whereas that of strong acid sites was not significantly affected [21,22]. This finding indicates that the addition of sulfate strongly influences the acid properties of alumina-based catalysts [10,14,19]. The total amounts of acid sites of both catalysts are also listed in Table 2. The total acid sites were higher for S/Ga-Al2O3, indicating that sulfate addition increased the surface acidity.
Figure 4 shows in situ FT-IR spectra of Ga-Al2O3 and S/Ga-Al2O3 catalysts exposed to a flow of NH3 at 25 °C for 1 h and then purged with He for 30 min to remove physically adsorbed species. In the case of Ga-Al2O3, peaks at 1262, 1462, 1612, and 1689 cm−1 were detected, losing intensity with increasing temperature. The bands at 1262 and 1612 cm−1 corresponded to the bending vibrations of N–H bonds in coordinated NH3+ on Lewis acid sites, and the peaks at 1462 and 1689 cm−1 were attributable to NH4+ species on Lewis acid sites [14,19,23]. The spectra of S/Ga-Al2O3 were different from those of the Ga-Al2O3 catalyst, featuring adsorption bands at 1386, 1486, 1620, and 1693 cm−1. The band at 1620 cm−1 on S/Ga-Al2O3 was assigned to coordinated ammonia species, the same as 1612 cm−1 on the Ga-Al2O3 catalyst [14,19]. The bands at 1486 and 1693 cm−1 were due to NH4+ species on Lewis acid sites. These IR bands of NH4+ species (1486 and 1693 cm−1) were blue-shifted by ~20 cm−1 compared to those of the non-sulfated catalyst because of the higher NH4+-catalyst bonding strength. Furthermore, a new peak at 1386 cm−1 in S/Ga-Al2O3 was observed at 250 °C, which was not detected below 200 °C, because of the nearby overlapping band. This could be assigned to the presence of medium Lewis acid sites, which are stable up to 500 °C. Thus, the NH3-TPD and FT-IR results imply that the amount of acid sites on the Ga-Al2O3 catalyst could be increased by sulfuric acid treatment.
Like the use of acid-treated catalysts for the decomposition of other fluorinated hydrocarbons, the catalytic activity for HFC-134a decomposition could be enhanced by acid treatment of the catalyst. Although sulfate treatment decreases the surface area and pore volume, it apparently increases the amount of Lewis acid sites that positively influence the HFC-134a decomposition.

2.2. Observation of Change in Surface Properties by HF Poisoning

As mentioned above, in the catalytic decomposition of HFC-134a, poisoning by HF is the main reason for catalytic deactivation. However, most of the studies so far have aimed only at improving the catalytic activity by increasing the acidity of the catalyst, and no detailed study of the physicochemical change on the used catalysts was investigated. We analyzed the change in the surface of the fresh and used catalysts via characterization by XRD, SEM-EDS, and XPS. For these analyses, the catalyst tested for 30 h in the HFC-134a decomposition reaction was referred to as a used catalyst.
The XRD pattern of used Ga-Al2O3 and S/Ga-Al2O3 catalysts is given in Figure 5. Used catalysts had γ-Al2O3 (JCPDS #29-63)-related peaks, and similar to the fresh catalysts, no peaks corresponding to Ga2O3 and sulfate species were detected [10,24]. However, the XRD patterns of used catalysts showed higher crystallinity than that of fresh catalysts, and characteristic peaks of AlF3 were clearly detected for both used catalysts. Thus, it was qualitatively confirmed that regardless of the acid treatment, AlF3 was formed on the catalyst surface during HFC-134a decomposition.
To investigate the formation of AlF3 located on the catalyst surface, SEM-EDS analysis was conducted on the used catalysts. Figure 6 shows the SEM images of the used catalysts, revealing the presence of AlF3 on both catalyst surfaces (in agreement with the XRD analysis in Figure 5). More AlF3 was observed on the surface of Ga-Al2O3 than on the surface of S/Ga-Al2O3. Table 3 shows the elemental compositions as determined by EDS. Although both used catalysts had similar Ga content, Ga-Al2O3 contained almost twice as much F as S/Ga-Al2O3. Therefore, in good agreement with the stability test, it might be concluded that sulfuric acid treatment not only improves the catalytic performance but also inhibits the formation of AlF3 on the catalyst surface.
The surface electronic state and atomic concentration of Ga and Al in the fresh and used catalysts were investigated by XPS analysis. A curve-fitting for this analysis was carried out after Shirley-type background subtraction using a combination of Gaussian and Lorentzian functions. Figure 7A depicts the Ga 2p3/2 spectra for the fresh Ga-Al2O3 and S/Ga-Al2O3 catalysts. The XPS peaks of Ga 2p3/2 at 1117.4 and 1118.7 eV can be ascribed to Ga0 and Ga3+ [25,26,27]. The Ga0 peak increased with sulfuric acid treatment of the Ga-Al2O3 catalyst, indicating that the acid sites on the Ga-Al2O3 catalyst could partially reduce the Ga3+ to Ga0 because they attract electrons to create more Lewis acid sites. Figure 7B presents the Ga 2p3/2 spectra for the used Ga-Al2O3 and S/Ga-Al2O3 catalysts. There was little change in peak position compared to fresh catalysts. The Ga0/(Ga0 + Ga3+) values given in Table 4 are different for the used catalysts, because the HFC-134a decomposition occurs in a highly oxidative atmosphere and at high temperature. The Ga0/(Ga0 + Ga3+) value of the Ga-Al2O3 catalyst decreased from 0.30 to 0.11, while the S/Ga-Al2O3 catalyst retained Ga0 species after the HFC-134a decomposition reaction. This result clearly indicates that the sulfuric acid treatment not only increases the acidity of the catalyst but also increases and preserves partially reduced Ga0 species. The Al 2p spectra of the fresh catalysts are shown in Figure 7C. Both catalysts have a well-developed Al 2p peak located at 74.2 eV, indicating the formation of an Al-O bond [28,29]. The peak shift with Al 2p on sulfuric acid treatment was not observed. However, in Figure 7D, another set of peaks, attributed to the Al-F bond, appeared in the range of 76.6–75.9 eV for the used Ga-Al2O3 and S/Ga-Al2O3 catalysts [28]. According to the literature, the binding energy range of the Al-F bond was found at 75.6–76.6 eV [28]. The XPS results of Al in Figure 7D are very similar to that, which can be thought of as peaks due to the formation of Al-F bonding. In the case of the Ga-Al2O3 catalyst, moreover, the peak intensity of the Al-O bond is significantly decreased by the formation of the Al-F bond [28,29,30]. It indicates that the Al-F bond of AlF3 was formed by the replacement of the Al-O bond of Al2O3 during the HFC-134a decomposition reaction. The appearance of the Al-F peak after the reaction indicates that F incorporation occurs only on the Al2O3 surface, and not Ga2O3. Furthermore, this result suggests that the sulfuric acid treatment on the Ga-Al2O3 catalyst could alleviate the elemental composition change from Al2O3 to AlF3.

3. Materials and Methods

3.1. Catalyst Preparation

Ga-Al2O3 was synthesized by co-precipitation, with the Ga loading fixed at 15 wt.%. Stoichiometric quantities of gallium nitrate (99.9%, Aldrich, St. Louis, MO, USA) and aluminum nitrate (98%, Aldrich) were dissolved in distilled water, and the resulting solution was slowly treated with 15 wt.% aqueous NH4OH with vigorous agitation until the pH reached 9.1. The resulting slurry was aged for 24 h at room temperature, and the precipitate was thoroughly washed to remove impurities, dried at 110 °C for 24 h, and calcined at 600 °C for 5 h to finally obtain the Ga-Al2O3 catalyst. S/Ga-Al2O3 was prepared by impregnating the Ga-Al2O3 catalyst with appropriate amounts of H2SO4 (1 wt.% S) followed by drying at 110 °C for 24 h and calcination at 600 °C for 5 h.

3.2. Catalytic Reaction

The catalytic reaction was performed in a fixed-bed Incornel reactor (10.5 mm i.d.) under atmospheric pressure. The reaction temperature was determined by using a thermocouple directly inserted into the catalyst bed. Prior to the reaction, the catalyst powders were pressed into pellets, crushed, and sieved to 40–60 mesh. The reactant gas mixture (1 vol.% HFC-134a, 25 vol.% H2O, and balance air) was introduced into the reactor at a gas hourly space velocity (GHSV) of 2362 h−1. Water, quantitatively introduced using a syringe pump, was passed through a pre-heater at 200 °C before being injected into the reactor. To remove HF, the product gas was passed through aqueous KOH and then analyzed by an online gas chromatograph equipped with a thermal conductivity detector (iGC 7200, DS Science, Gwangju, Gyeonggi, R. Korea).

3.3. Characterization

The crystal structure of the catalyst was probed by X-ray diffraction (XRD, Rigaku D/MAX-2500, Cu Kα radiation). Brunauer–Emmett–Teller (BET) surface areas were determined from N2 adsorption–desorption isotherms recorded at −196 °C (BELSORP-max, BEL Japan, Inc., Osaka, Japan). The structure of the catalyst samples was observed by scanning electron microscopy (SEM, Hitachi S-4300, Tokyo, Japan) coupled with energy-dispersive X-ray spectroscopy (EDS, Horiba EX-200, Horiba, Tokyo, Japan). Prior to the temperature-programmed desorption of ammonia (NH3-TPD) experiments (BELCAT II, BEL Japan, Inc.), samples were pretreated in helium flow at 400 °C for 1 h to remove impurities, cooled to 50 °C, exposed to excess 5% NH3/He for 1 h, and purged with He. NH3-TPD was performed at temperatures of up to 700 °C in helium flow. In situ Fourier-transform infrared spectroscopy (FT-IR) was carried out in a ceramic IR cell equipped with ZnSe windows using a diffuse-reflectance infrared (IR) accessory (PIKE Technologies, Madison, WI, USA) connected to a Nicolet iS10 (Thermo Scientific, Waltham, MA, USA) IR spectrometer with an MCT-A detector. Spectra were recorded by the averaging of 64 scans with a resolution of 8 cm−1. Before IR spectral observation, samples were pretreated in a flow of He at 400 °C for 1 h to remove impurities, and then cooled down to 25 °C to probe NH3 adsorption behavior in the temperature range of 25–600 °C. X-ray photoelectron spectroscopy (XPS) was conducted on an ESCALAB Mark II spectrometer (Vacuum Generators, Su ssex, UK) using Al Kα radiation ( = 1486.6 eV) at a constant energy of 50 eV. The binding energy was aligned based on the C 1s transition at 285 eV.

4. Conclusions

To investigate the effect of sulfuric acid treatment on catalysts for HFC-134a decomposition, Ga-Al2O3 and S/Ga-Al2O3 catalysts were prepared by co-precipitation and impregnation methods. The S/Ga-Al2O3 catalyst achieved a much higher HFC-134a conversion than Ga-Al2O3, which was ascribed to the promotional effects of sulfuric acid treatment on catalytic activity, as reported in many earlier studies for the catalytic decomposition of other fluorinated hydrocarbons. The effects of sulfuric acid treatment were probed by NH3-TPD and in-situ FT-IR analysis. Treatment with sulfuric acid was shown to influence the amount of Lewis acidity and improve the catalytic activity for HFC-134a decomposition. Furthermore, the S/Ga-Al2O3 catalyst retained its efficiency with minor fluctuation for the 30 h test, with its HFC-134a conversion maintained at ~80%.
The changes in surface structure of the used catalysts were characterized by XRD, SEM-EDS, and XPS analyses. Both catalysts contained AlF3 after 30 h of HFC-134a decomposition reaction, confirmed by XRD. In particular, almost twice as many F sources were detected in the Ga-Al2O3 catalyst compared to the S/Ga-Al2O3 catalyst. Based on the XPS analysis results, the sulfuric acid treatment not only increased the acidity of the catalyst but also preserved the partially reduced Ga species. Moreover, this treatment could alleviate the elemental composition change from Al2O3 to AlF3.

Author Contributions

For this research article, the individual contribution of authors was as follows: Conceptualization, M.-J.K. and S.G.J.; methodology, Y.K. and I.-H.C.; investigation, Y.K., J.-R.Y. and I.-H.C.; formal analysis, S.G.K.; writing—original draft preparation, M.-J.K.; writing—review and editing, M.-J.K. and S.G.J.; visualization, K.-R.H. and Y.-K.P.; funding acquisition, S.-H.M.; supervision, K.B.L., S.G.J. and S.-H.M. All authors have read and agreed to the published version of manuscript.

Funding

This research was funded by the Korea Ministry of Environment (MOE) as the Global Top Environment R&D program grant No. 2017001700002.

Acknowledgments

This project was supported by the R&D Center for reduction of Non-CO2 Greenhouse Gases (2017001700002) funded by the Korea Ministry of Environment (MOE) as the Global Top Environment R&D Program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ravishankara, A.; Solomon, S.; Turnipseed, A.A.; Warren, R. Atmospheric lifetimes of long-lived halogenated species. Science 1993, 259, 194–199. [Google Scholar] [CrossRef] [Green Version]
  2. Franklin, J. The atmospheric degradation and impact of 1, 1, 1, 2-tetrafluoroethane (hydrofluorocarbon 134a). Chemosphere 1993, 27, 1565–1601. [Google Scholar] [CrossRef]
  3. Karmakar, S.; Greene, H.L. Oxidative destruction of chlorofluorocarbons (CFC11 and CFC12) by zeolite catalysts. J. Catal. 1992, 138, 364–376. [Google Scholar] [CrossRef]
  4. Choi, S.-S.; Park, D.-W.; Watanabe, T. Thermal plasma decomposition of fluorinated greenhouse gases. Nucl. Eng. Technol. 2012, 44, 21–32. [Google Scholar] [CrossRef] [Green Version]
  5. Mi, T.; Han, J.; He, X.; Qin, L. Investigation of HFC-134a decomposition by combustion and its kinetic characteristics in a laboratory scale reactor. Environ. Prot. Eng. 2015, 41, 143–150. [Google Scholar]
  6. Watanabe, T.; Tsuru, T. Water plasma generation under atmospheric pressure for HFC destruction. Thin Solid Films 2008, 516, 4391–4396. [Google Scholar] [CrossRef]
  7. 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]
  8. 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]
  9. Iizuka, A.; Ishizaki, H.; Mizukoshi, A.; Noguchi, M.; Yamasaki, A.; Yanagisawa, Y. Simultaneous decomposition and fixation of F-gases using waste concrete. Ind. Eng. Chem. Res. 2011, 50, 11808–11814. [Google Scholar] [CrossRef]
  10. 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]
  11. 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]
  12. Takita, Y.; Moriyama, J.-I.; Yoshinaga, Y.; Nishiguchi, H.; Ishihara, T.; Yasuda, S.; Ueda, Y.; Kubo, M.; Miyamoto, A. Adsorption of water vapor on the AlPO4-based catalysts and reaction mechanism for CFCs decomposition. Appl. Catal. A Gen. 2004, 271, 55–60. [Google Scholar] [CrossRef]
  13. Takita, Y.; Ishihara, T. Catalytic decomposition of CFCs. Catal. Surv. Asia 1998, 2, 165–173. [Google Scholar] [CrossRef]
  14. Jia, W.; Wu, Q.; Lang, X.; Hu, C.; Zhao, G.; Li, J.; Zhu, Z. Influence of lewis acidity on catalytic activity of the porous alumina for dehydrofluorination of 1,1,1,2-tetrafluoroethane to trifluoroethylene. Catal. Lett. 2014, 145, 654–661. [Google Scholar] [CrossRef]
  15. Kashiwagi, D.; Takai, A.; Takubo, T.; Nagaoka, K.; Inoue, T.; Takita, Y. Metal phosphate catalysts effective for degradation of sulfur hexafluoride. Ind. Eng. Chem. Res. 2008, 48, 632–640. [Google Scholar] [CrossRef]
  16. Kashiwagi, D.; Takai, A.; Takubo, T.; Yamada, H.; Inoue, T.; Nagaoka, K.; Takita, Y. Catalytic activity of rare earth phosphates for SF6 decomposition and promotion effects of rare earths added into AlPO4. J. Colloid Interf. Sci. 2009, 332, 136–144. [Google Scholar] [CrossRef]
  17. Afonasenko, T.N.; Leont’eva, N.N.; Talzi, V.P.; Smirnova, N.S.; Savel’eva, G.G.; Shilova, A.V.; Tsyrul’nikov, P.G. Synthesis and properties of γ-Ga2O3–Al2O3 solid solutions. Russ. J. Phys. Chem. A 2017, 91, 1939–1945. [Google Scholar] [CrossRef]
  18. 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 2004, 90, 283–290. [Google Scholar] [CrossRef]
  19. Mekhemer, G.A.H.; Khalaf, H.A.; Mansour, S.A.A.; Nohman, A.K.H. Sulfated alumina catalysts: Consequences of sulfate content and source. Monatsh. Chem. 2005, 136, 2007–2016. [Google Scholar] [CrossRef]
  20. 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 NO over Cu-SSZ-13 and Cu-SAPO-34 catalysts. Appl. Catal. B Environ. 2014, 156–157, 428–437. [Google Scholar] [CrossRef]
  21. 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]
  22. Han, T.U.; Yoo, B.-S.; Kim, Y.-M.; Hwang, B.; Sudibya, G.L.; Park, Y.-K.; Kim, S. Catalytic conversion of 1,1,1,2-tetrafluoroethane (HFC-134a). Korean J. Chem. Eng. 2018, 35, 1611–1619. [Google Scholar] [CrossRef]
  23. Jin, R.; Liu, Y.; Wang, Y.; Cen, W.; Wu, Z.; Wang, H.; Weng, X. The role of cerium in the improved SO2 tolerance for NO reduction with NH3 over Mn-Ce/TiO2 catalyst at low temperature. Appl Catal. B Environ. 2014, 148–149, 582–588. [Google Scholar] [CrossRef]
  24. Takita, Y.; Wakamatsu, H.; Li, G.-L.; Moro-Oka, Y.; Nishiguchi, H.; Ishihara, T. Decomposition of chlorofluorocarbons over metal phosphate catalysis: II. Origin of the stability of AlPO4 and the location of Ce as a promoter. J. Mol. Catal. A Chem. 2000, 155, 111–119. [Google Scholar] [CrossRef]
  25. Yang, K.; Zhu, L.; Zhang, J.; Huo, X.; Lai, W.; Lian, Y.; Fang, W. Co-aromatization of n-butane and methanol over PtSnK-Mo/ZSM-5 zeolite catalysts: The promotion effect of ball-milling. Catalysts 2018, 8, 307. [Google Scholar] [CrossRef] [Green Version]
  26. Chen, H.; Hu, J.; Li, G.D.; Gao, Q.; Wei, C.; Zou, X. Porous Ga-In bimetallic oxide nanofibers with controllable structures for ultrasensitive and selective detection of formaldehyde. ACS Appl. Mater. Interfaces 2017, 9, 4692–4700. [Google Scholar] [CrossRef]
  27. Xiao, H.; Zhang, J.; Wang, X.; Zhang, Q.; Xie, H.; Han, Y.; Tan, Y. A highly efficient Ga/ZSM-5 catalyst prepared by formic acid impregnation and in situ treatment for propane aromatization. Catal. Sci. Technol. 2015, 5, 4081–4090. [Google Scholar] [CrossRef]
  28. Limcharoen, A.; Pakpum, C.; Limsuwan, P. An X-ray photoelectroscopy investigation of re-deposition from fluorine-based plasma etch on magnetic recording slider head substrate. Procedia Eng. 2012, 32, 1043–1049. [Google Scholar] [CrossRef] [Green Version]
  29. 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]
  30. 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. HFC-134a conversion over Ga-Al2O3 and S/Ga-Al2O3 catalysts. Reaction conditions: 1 vol.% HFC-134a, 25 vol.% H2O in air balance; GHSV = 2362 h−1. Inset: Catalytic stability test for the HFC-134a decomposition at 450 °C. Reaction condition: 1 vol.% HFC-134a, 25 vol.% H2O in air balance; GHSV = 2362 h−1.
Figure 1. HFC-134a conversion over Ga-Al2O3 and S/Ga-Al2O3 catalysts. Reaction conditions: 1 vol.% HFC-134a, 25 vol.% H2O in air balance; GHSV = 2362 h−1. Inset: Catalytic stability test for the HFC-134a decomposition at 450 °C. Reaction condition: 1 vol.% HFC-134a, 25 vol.% H2O in air balance; GHSV = 2362 h−1.
Catalysts 10 00766 g001
Figure 2. X-ray diffraction (XRD) patterns of fresh catalysts.
Figure 2. X-ray diffraction (XRD) patterns of fresh catalysts.
Catalysts 10 00766 g002
Figure 3. NH3-TPD following NH3 adsorption in a flow of NH3/He at 50 °C for 1 h.
Figure 3. NH3-TPD following NH3 adsorption in a flow of NH3/He at 50 °C for 1 h.
Catalysts 10 00766 g003
Figure 4. In situ FT-IR spectra as a function of temperature for fresh Ga-Al2O3 (A) and S/Ga-Al2O3 (B) catalysts in He flow after NH3 adsorption.
Figure 4. In situ FT-IR spectra as a function of temperature for fresh Ga-Al2O3 (A) and S/Ga-Al2O3 (B) catalysts in He flow after NH3 adsorption.
Catalysts 10 00766 g004
Figure 5. X-ray diffraction (XRD) patterns of used catalysts.
Figure 5. X-ray diffraction (XRD) patterns of used catalysts.
Catalysts 10 00766 g005
Figure 6. SEM images of used catalyst: Ga-Al2O3 (A), S/Ga-Al2O3 (B).
Figure 6. SEM images of used catalyst: Ga-Al2O3 (A), S/Ga-Al2O3 (B).
Catalysts 10 00766 g006
Figure 7. XPS spectra of Ga-Al2O3 and S/Ga-Al2O3 catalysts: (A) Ga 2p3/2 of fresh, (B) Ga 2p3/2 of used, (C) Al 2p of fresh, and (D) Al 2p of used.
Figure 7. XPS spectra of Ga-Al2O3 and S/Ga-Al2O3 catalysts: (A) Ga 2p3/2 of fresh, (B) Ga 2p3/2 of used, (C) Al 2p of fresh, and (D) Al 2p of used.
Catalysts 10 00766 g007
Table 1. Elemental composition of Ga-Al2O3 and S/Ga-Al2O3 catalysts.
Table 1. Elemental composition of Ga-Al2O3 and S/Ga-Al2O3 catalysts.
SampleGa (wt.%) *Al (wt.%) *S (wt.%) **
Ga-Al2O314.752.8
S/Ga-Al2O315.353.51.51
* Estimated by inductively coupled plasma—optical emission spectrometry. ** Estimated by elemental analyzer.
Table 2. Characterization results of catalysts: BET surface area, pore volume, and temperature-programmed desorption of ammonia (NH3-TPD).
Table 2. Characterization results of catalysts: BET surface area, pore volume, and temperature-programmed desorption of ammonia (NH3-TPD).
SamplesBET Surface Area [m2 g−1]Pore Volume [m3 g−1]Amount of Acid Site [mmol g−1]
WeakMediumStrongTotal
Ga-Al2O3227.50.350.0720.1540.3420.568
S/Ga-Al2O3187.40.290.1170.2070.3120.646
Table 3. EDS analysis of used catalysts.
Table 3. EDS analysis of used catalysts.
Elements (Weight %)Ga-Al2O3S/Ga-Al2O3
Ga13.913.8
Al30.634.4
O30.337.2
F25.213.8
S 0.8
Table 4. Surface atomic concentration of Ga0/(Ga0 + Ga3+) and binding energies for the Ga0 and Ga3+ value in Ga 2p3/2.
Table 4. Surface atomic concentration of Ga0/(Ga0 + Ga3+) and binding energies for the Ga0 and Ga3+ value in Ga 2p3/2.
SamplesGa0/(Ga0 + Ga3+)Binding Energy (eV)
Ga0Ga3+
Ga-Al2O30.301117.41118.7
S/Ga-Al2O30.441117.51118.8
Ga-Al2O3 used0.111117.51118.7
S/Ga-Al2O3 used0.411117.41118.8

Share and Cite

MDPI and ACS Style

Kim, M.-J.; Kim, Y.; Youn, J.-R.; Choi, I.-H.; Hwang, K.-R.; Kim, S.G.; Park, Y.-K.; Moon, S.-H.; Lee, K.B.; Jeon, S.G. Effects of Sulfuric Acid Treatment on the Performance of Ga-Al2O3 for the Hydrolytic Decomposition of 1,1,1,2-Tetrafluoroethane (HFC-134a). Catalysts 2020, 10, 766. https://doi.org/10.3390/catal10070766

AMA Style

Kim M-J, Kim Y, Youn J-R, Choi I-H, Hwang K-R, Kim SG, Park Y-K, Moon S-H, Lee KB, Jeon SG. Effects of Sulfuric Acid Treatment on the Performance of Ga-Al2O3 for the Hydrolytic Decomposition of 1,1,1,2-Tetrafluoroethane (HFC-134a). Catalysts. 2020; 10(7):766. https://doi.org/10.3390/catal10070766

Chicago/Turabian Style

Kim, Min-Jae, Yeonjin Kim, Jae-Rang Youn, Il-Ho Choi, Kyung-Ran Hwang, Seung Gon Kim, Young-Kwon Park, Seung-Hyun Moon, Ki Bong Lee, and Sang Goo Jeon. 2020. "Effects of Sulfuric Acid Treatment on the Performance of Ga-Al2O3 for the Hydrolytic Decomposition of 1,1,1,2-Tetrafluoroethane (HFC-134a)" Catalysts 10, no. 7: 766. https://doi.org/10.3390/catal10070766

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