Next Article in Journal
Y-Modified MCM-22 Supported PdOx Nanocrystal Catalysts for Catalytic Oxidation of Toluene
Previous Article in Journal
Effect of Shale Ash-Based Catalyst on the Pyrolysis of Fushun Oil Shale
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Use of a γ-Al2O3 and MgO Mixture in the Catalytic Conversion of 1,1,1,2-Tetrafluoroethane (HFC-134a)

by
Sangjae Jeong
1,
Gamal Luckman Sudibya
1,
Jong-Ki Jeon
2,
Young-Min Kim
3,*,
Caroline Mercy Andrew Swamidoss
4 and
Seungdo Kim
1,5,*
1
Research Center for Climate Change and Energy, Hallym University, Chuncheon 24252, Korea
2
Department of Chemical Engineering, Kongju National University, Cheonan 31080, Korea
3
Department of Environmental Engineering, Daegu University, Gyeongsan 38453, Korea
4
Department of Science and Humanities, Saveetha School of Engineering, Chennai 600124, India
5
Department of Environmental Sciences and Biotechnology, Hallym University, Chuncheon 24252, Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(11), 901; https://doi.org/10.3390/catal9110901
Submission received: 25 September 2019 / Revised: 25 October 2019 / Accepted: 25 October 2019 / Published: 28 October 2019
(This article belongs to the Section Environmental Catalysis)

Abstract

:
This paper reports the improved efficiency of 1,1,1,2-tetrafluoroethane (HFC-134a) decomposition by combined use of MgO with γ-Al2O3. While a high temperature (>900 °C) was required to achieve 90% conversion during non-catalytic pyrolysis of HFC-134a, 100% conversion of HFC-134a was achieved at 600 °C by the use of γ-Al2O3. Among the three catalysts (γ-Al2O3, MgO, and CaO) tested in this study, γ-Al2O3 showed the highest HFC-134a decomposition efficiency, followed by MgO and CaO, due to its large surface area and large amount of weak acid sites. Also with the longest lifetime among the catalysts, durability in maintaining complete decomposition of HFC-134a was shown in γ-Al2O3. The addition of MgO to γ-Al2O3 was effective in extending the lifetime of γ-Al2O3 due to the efficient interaction between HF and MgO, which can delay the deactivation of γ-Al2O3. Compared to the double bed γ-Al2O3-MgO configuration, the use of a mixed γ-Al2O3-MgO bed extended the catalyst lifetime more effectively.

1. Introduction

The anthropogenic impact on climate change has drawn global attention to the reduction of greenhouse gases (GHG). Different from the Kyoto Protocol (1997), which requires the reduction of GHG emission in developed countries, the Paris Agreement (2015) demands all nations of the world to devote decisive action to deeply cutting GHG emissions. Recently, the Paris rulebook was established in the 24st Conference of the Parties (COP24) of United Nations Framework Convention on Climate Change (UNFCCC) held in Katowice, Poland, demanding an increase in greenhouse gas reduction technologies.
Hydrofluorocarbons (HFCs) are one of the six major GHGs in the Kyoto Protocol. Since HFCs show high global warming potential (GWP), ranging from 124 to 14,800 [1], their reduction is considered to be cost-effective. In the EU, HFC emissions account for more than 90% in total fluorinated GHG emissions and have more than tripled since 1990 [2]. Not only in the EU, there will be similar trends in other countries. As HFCs were used since the 1990s, the end of life of the electric devices filled with HFCs is coming and the emissions of HFCs are expected to increase dramatically. The Kigali Amendment promulgated the phase-down on production and consumption of HFCs [3]. It is urgent to develop efficient mitigation technologies for HFCs, along with seeking new generation refrigerants which have no adverse ozone depletion and climate change effects.
Extensive studies have been carried out on the destruction of HFCs, including combustion, plasma technologies and catalytic decomposition [4,5,6,7]. Thermal combustion and plasma technologies have been regarded as reliable destruction and removal processes for ozone-depleting substances like primary chlorofluorocarbons (CFCs) and halons. Despite their excellent destruction efficiency for HFCs, they consume large amounts of energy and can generate secondary air pollutants [4].
Catalytic conversion is considered as an alternative process in decomposing 1,1,1,2-tetrafluoroethane (HFC-134a), which is one of the major HFCs and used extensively as refrigerant. Catalytic conversion has advantages regarding low energy consumption, in that it can react at temperatures lower than 973K with alumina-based catalysts [7]. Also, a precious by-product is produced after catalytic conversion of HFC-134a. Trifluoroethylene (TrFE; CF2=CHF; HFO-1123) is a major by-product during catalytic conversion of HFC-134a with Al2O3. TrFE is a valuable material which can be utilized as a co-monomer or next generation refrigerant revealing a low global warming potential [8,9,10,11].
Up to now, alumina catalyst (Al2O3) in the gamma phase (γ) is known as an efficient catalyst for the decomposition of fluorinated gases. The performance of γ-Al2O3-based catalysts in the decomposition of HFCs [12] and PFCs [13] has been investigated in previous studies. Although the catalytic activity of γ-Al2O3 in the conversion of HFC-134a is high, only a few studies have been conducted on the deactivation mechanism of γ-Al2O3 and its enhancement method. The key points of catalytic conversion process are reaction efficiency, selectivity and the lifetime of catalysts. Previous studies were mainly focused on the reaction efficiencies.
Therefore, in this study, we compared the catalytic efficiencies of three selected catalysts (γ-Al2O3, MgO and CaO) and investigated key factors affecting the reaction efficiency based on measurements of specific surfaces area and the amount of acid sites. The deactivation mechanism of the catalysts was identified with long-term tests and surface characterization of spent catalysts after various reaction times using X-ray diffraction (XRD) and Energy Dispersive X-ray Spectrometry (EDS). Based on the identified deactivation mechanism, we attempted to improve the lifetime of the γ-Al2O3 catalyst by the addition of MgO in the catalyst bed for the decomposition of HFC-134a. This process is much more convenient, simple, and economically viable due to the low cost of MgO compared to that of γ-Al2O3. This application is beneficial in that the HF generated as a result of the dehydrofluorination reaction of HFC-134a can be neutralized by alkali metal oxides, impeding gas-solid reaction between HF and the γ-Al2O3 catalyst. This will prevent rapid deactivation of the main catalyst and, subsequently, extend the catalytic activity.

2. Results and Discussion

2.1. Thermal and Catalytic Decomposition of HFC-134a over Single Bed Catalyst

Figure 1 shows the effect of reaction temperature on the non-catalytic and catalytic conversion of HFC-134a over γ-Al2O3, MgO, and CaO. Owing to the thermal stability of HFC-134a [7], the thermal decomposition of HFC-134a was initiated at temperatures higher than 600 °C. The conversion efficiency of HFC-134a was gradually increased up to 89.9% by elevating the reaction temperature from 600 to 900 °C. Meanwhile, the catalytic conversion of HFC-134a over all catalysts was initiated even at the lowest temperature (300 °C) applied in this study. The conversion efficiencies of HFC-134a over all catalysts were increased by elevating the reaction temperature from 300 to 600 °C. Among the selected catalysts, γ-Al2O3 revealed the highest HFC-134a conversion efficiency at all temperatures, followed by MgO and CaO. At 600 °C, the catalytic conversion of HFC-134a over γ-Al2O3 achieved complete conversion and the conversion efficiencies of MgO and CaO were 92% and 64%, respectively.
Figure 2 shows the time on stream HFC-134a conversion efficiencies over 12 h for γ-Al2O3, MgO, and CaO. Complete conversion of HFC-134a over γ-Al2O3 was maintained until 9 h of reaction, and the conversion efficiency gradually decreased to 88% after 12 h. Meanwhile, the conversion of HFC-134a over MgO was initially high, but rapidly decreased to 37% after 6 h of reaction. CaO revealed the lowest conversion efficiency, even at the initial reaction time, among the selected catalysts. It also showed a dramatic decrease of HFC-134a conversion efficiency compared to γ-Al2O3 and MgO. The conversion efficiency of HFC-134a over CaO rapidly decreased from 64% to 29% within 1 h and maintained a low HFC-134a conversion efficiency between 20 and 40% until the end of the test cycle (12 h). The above results indicate that the γ-Al2O3 displays not only higher conversion efficiency in the initial stage, but also longer lifetime during continuous reaction than those of other catalysts used in this study.
The higher conversion efficiency of HFC-134a on γ-Al2O3 compared to those on CaO and MgO can be explained by their respective specific surface areas and acidic properties. The BET surface area (SBET) of γ-Al2O3, 187 m2/g, was higher than that of MgO (144 m2/g) and much greater than that of CaO (3 m2/g). The higher surface area of catalyst can allow more chances for contact between reactants, thus allowing better mass transfer conditions of the reactants [14,15]. Another important key factor affecting conversion efficiency of HFC-134a was the acidic property of the catalyst. Figure 3 and Table 1 show the temperature-programmed desorption of ammonia (NH3-TPD) curves and the amount of weak, intermediate, and strong acid sites of fresh γ-Al2O3, MgO, and CaO, respectively. The amount of acid sites in the catalyst were classified to weak, intermediate, and strong acid sites based on the NH3-desorption temperature region; weak at below 250 °C, intermediate at between 250 and 400 °C, and strong at above 400 °C [16].
Among the selected catalysts, γ-Al2O3 had the biggest amount of total acid sites, followed by CaO and MgO. This can explain the higher HFC-134a conversion efficiency of γ-Al2O3 than CaO and MgO. Although MgO had a smaller amount of total acid sites than CaO, its HFC-134a decomposition efficiency was higher than that of CaO owing to its much higher BET surface area. The higher performance of MgO than CaO can also be explained with the amount of weak acid sites. Although MgO had a lower amount of intermediate and strong acid sites than CaO, it had a much higher amount of weak acid sites than CaO. This suggests that the amount of weak acid sites on the catalyst is closely related with the decomposition efficiency of HFC-134a. Several researchers have also emphasized that the weak acid sites of the catalyst play a crucial role in the conversion of HFC-134a [12,13]. Jia et al. reported that the weak acid sites act as the reactive ones for HFC-134a decomposition, while strong acid sites cause formation of coke from HFC-134a or TrFE [12]. The descending order of weak acid sites and SBET of the catalysts, γ-Al2O3, MgO, and CaO, was in accordance with that of catalytic conversion efficiencies of HFC-134a, confirming that weak acid sites and SBET play the important role on catalytic conversion of HFC-134a [13,17].

2.2. The Use of Al2O3 -MgO Catalysts on the Catalytic Decomposition of HFC-134a

The major by-products obtained from the catalytic decomposition of HFC-134a over γ-Al2O3 are known to be HF and TrFE [12]. Generated HF can react with γ-Al2O3 by gas-solid reaction forming AlF3. Although AlF3 can be a strong acid catalyst, it shows different catalytic activities depending on different phases. At temperatures over 600 °C, AlF3 becomes an alpha (α) phase which does not exhibit any catalytic activity [18,19]. Since the reaction temperature in this study is 600 °C, γ-Al2O3 is deactivated when it is converted into AlF3 during the HFC-134a conversion reaction. Therefore, the formation of AlF3 during the catalytic conversion of HFC-134a needs to be minimized to extend the lifetime of γ-Al2O3 catalyst. The additional use of MgO as an additive catalyst was introduced to extend the lifetime of γ-Al2O3 on the catalytic conversion of HFC-134a. Between CaO and MgO, MgO was chosen due to its greater amount of weak acid sites and larger BET surface area. Figure 4 shows the different conversion efficiencies of HFC-134a over single bed γ-Al2O3 (single Bed, SB), double bed γ-Al2O3-MgO (Double Bed, DB), and mixed bed γ-Al2O3-MgO (mixed single-bed, MD). Duration time maintaining the conversion efficiency of HFC-134a higher than 99% (DT99) over each catalyst bed were 4.4 h over SB-γ-Al2O3-0.6, 6.9 h over DB-γ-Al2O3-0.6-MgO-0.6, 8.8 h over MB-γ-Al2O3-0.6-MgO-0.6, and 9.1 h over SB-γ-Al2O3-1.2, respectively.
Although it could not achieve higher and longer decomposition efficiency than SB-γ-Al2O3-1.2, DB-γ-Al2O3-0.6-MgO-0.6 revealed the longer DT99 (6.9 h) than that of SB-γ-Al2O3-0.6. This suggests that the additional use of MgO can provide a higher conversion efficiency of HFC-134a, as can be expected with the result in Figure 2. When we compare SB-γ-Al2O3-1.2 and MB-γ-Al2O3-0.6-MgO-0.6, 0.6 g of MgO substituted 0.6 g of γ-Al2O3 with similar DT99. In case of SB-γ-Al2O3-0.6 and MB-γ-Al2O3-0.6-MgO-0.6, DT99 increased 1.6 times with addition of 0.6g MgO. Since the price of γ-Al2O3 is much higher than that of MgO, additional use of MgO with γ-Al2O3 will highly reduce the cost of catalysts. Interestingly, MB-γ-Al2O3-0.6-MgO-0.6 led the longer DT99 (8.8 h) than DB-γ-Al2O3-0.6-MgO-0.6 (6.9 h) on the decomposition of HFC-134a. This suggests that the lifetime of γ-Al2O3, MgO, or both of those catalysts can be extended by the use of mixed γ-Al2O3 and MgO.
The higher and longer catalytic activity of MB-γ-Al2O3-0.6-MgO-0.6 compared to DB-γ-Al2O3-0.6-MgO-0.6 can be explained by the synergistic effect on the conversion of HFC-134a enhanced by the different roles of γ-Al2O3 and MgO. During the thermal treatment of HFC-134a over γ-Al2O3, there are two reactions: (i) HFC-134a decomposition into HF and TrFE and (ii) γ-Al2O3 conversion to AlF3 by gas-solid reaction with HF. When MgO was present together with γ-Al2O3, the HF preferred to react with MgO rather than γ-Al2O3 and MgO was converted to MgF2. The HF consumption of MgO can prolong the lifetime of γ-Al2O3 and the dry mixing of γ-Al2O3, causing a synergistic effect which can increase the reaction efficiency and lifetime of the catalysts. A similar concept was proposed by Han et al. [20]. They achieved stable CF4 conversion efficiency by the additional use of Ca(OH)2 with γ-Al2O3 on the catalytic hydrolysis of CF4 compared to the single γ-Al2O3.
The role of MgO was identified by the comparison XRD of spent catalysts in the presence or absence of MgO. In our result, the diffraction lines of fresh γ-Al2O3 and spent catalysts collected after the specific reaction times (6, 9, 12 h) were different. In the absence of MgO, γ-Al2O3 was converted into AlF3 from the sample collected after the 6 h of reaction. This coincides with the decrease in catalytic conversion efficiency of HFC-134a over SB-γ-Al2O3-0.6 (Figure 5). The diffraction line corresponding to AlF3 also increased with the increase of reaction time. This also confirms that the decrease of HFC-134a conversion efficiency is caused by the conversion of γ-Al2O3 to AlF3. The formation of AlF3 on the catalytic decomposition of HFC-134a is related with the gas-solid reaction of γ-Al2O3 with HF, which can be expressed as following Equation (1) [18]:
Al 2 O 3 + 6 HF 2 AlF 3 + 3 H 2 O ,   Δ R G 0 = 329   kJ   mol 1
Figure 6 shows the XRD results of fresh and spent catalysts used in DB-γ-Al2O3-0.6-MgO-0.6. In the case of DB-γ-Al2O3-0.6-MgO-0.6, γ-Al2O3 can be found until 9 h of reaction (Figure 6a) and MgO was gradually converted into MgF2 (Figure 6b). As shown above, the HFC-134a conversion efficiency over DB-γ-Al2O3-0.6-MgO-0.6 started to decrease after 7 h and a 76% conversion efficiency was measured after 9 h. Combining the long-term test and XRD results, we can conclude that the conversion efficiency of HFC-134a is highly dependent on the presence of γ-Al2O3, and the addition of MgO extended the lifetime γ-Al2O3 by preventing fluorination of γ-Al2O3. Prevention of fluorination of γ-Al2O3 by MgO can also be found in the enthalpy change shown in Equation (2) below:
2 AlF 3 + 3 MgO Al 2 O 3 + 3 MgF 2 ,   Δ R G 0 = 232     kJ   mol 1
EDS was also conducted to identify the quantitative difference between SB-γ-Al2O3-0.6 and DB-γ-Al2O3-0.6-MgO-0.6. EDS result of fresh and spent γ-Al2O3 (Table 2) revealed that the contents of Al and O decreased together with the increase of the contents of F and C during long-term catalytic conversion of HFC-134a. The content of F greatly increased after 6 h and was maximized after 9 h due to the deactivation of γ-Al2O3 in the form of α-AlF3 and the effective defluorination reaction and fixation of HF. After 12 h, the carbon content largely increased, meaning that the large amount of carbon coke is also formed after the deactivation of the catalyst (Figure S1).
Table 3 shows the EDS result for fresh and used γ-Al2O3 and MgO in a serial bed reactor. In contrast to the case of a single bed reactor, oxygen content in γ-Al2O3 was decreased more slowly. The noticeable increase of F and C was initiated after 9 h (Figure S1). This confirms that the deactivation by the accumulation of F and C to γ-Al2O3 was delayed by the additional use of MgO to γ-Al2O3.
According to the long-term conversion test of HFC-134a, XRD and EDS, it was found that the addition of MgO to γ-Al2O3 catalyst can be an efficient method for extending the catalyst lifetime in the catalytic conversion of HFC-134a. Deactivation of γ-Al2O3 was proceeded by fluorination at the first stage and by carbon deposition in the second stage. The lifetime of γ-Al2O3 was extended largely by the additional use of MgO on the catalytic decomposition of HFC-134a because of the slower fluorination of γ-Al2O3 with the help of the added MgO.

3. Experimental

3.1. Catalysts

γ-Alumina (γ-Al2O3, Alfa Aesar, Haverhill, MA, USA), MgO (Daejung Chemicals and Metals Co., Ltd., Siheung, South Korea) and CaO (Junsei Chemicals Co., Ltd., Tokyo, Japan) were used as the catalysts for the catalytic decomposition of HFC-134a in this study. The catalysts were crushed, sieved to make their average particle size of 1.0–1.7 mm, calcined at 650 °C for four hours, and stored in a desiccator before use. In order to analyze the effect of each catalyst on the catalytic conversion of HFC-134a, 1.2 g of each catalyst was prepared as a single bed catalyst, respectively. To investigate the synergistic effect of different catalysts, double catalyst bed and mixed bed were prepared as shown in Figure 7. The double bed consisted of two catalyst layers, an upper bed with γ-Al2O3 (0.6 g) and a lower bed with MgO (0.6 g), and identified as “DB-γ-Al2O3-0.6-MgO-0.6”. The mixed bed was prepared by dry mixing of 0.6 g of MgO and 0.6 g of γ-Al2O3. The catalyst performance of double bed and mixed bed were also evaluated by comparing their HFC-134a conversion efficiency with those of single γ-Al2O3 beds (0.6 or 1.2 g).

3.2. Catalytic Decomposition of HFC-134a Using a Lab Scale Fixed-Bed Reactor

The catalytic decomposition of HFC-134a was carried out in a lab-scale fixed-bed continuous reactor (Figure 8). The reactor system consisted of a reactor, syringe pump, mass flow controller, furnace, and HF trap. Reaction temperature was monitored using a thermocouple. 12 h of continuous reaction was conducted for each experiment. The reactor had 3.6 ml inner volume and was made with SUS-316L tube. Except for the case of 0.6 g γ-Al2O3 loaded single bed configuration, weight hourly space velocity (WHSV) was 5000 ml/g-catalyst while a mixture of input flow was composed of 2 mL/min of HFC-134a and 98 mL/min of nitrogen gas to make a total gas flow of 100 mL/min. Only the amount of γ-Al2O3 was reduced in the case of 0.6 γ-Al2O3 loaded single bed configuration. Flow rates of HFC-134a and nitrogen gas were controlled by mass flow controllers (MFCs) (Kofloc, Kojima Instruments Inc., Kyoto, Japan). The isothermal catalytic decomposition reaction was conducted at the temperature range of 300 °C up to 600 °C in 100 °C intervals. The produced HF was trapped by CaO pellet to reduce the HF amount in the effluent gas before entering the gas chromatograph (GC).
An online gas chromatograph (GC) (Agilent 7890A, Santa Clara, CA, USA) equipped with a GS-GASPRO column and a mass spectroscopy detector (MSD) (Agilent 5975C) was used. The composition of effluent gas was automatically analyzed every 15 minutes. NIST 8th library (National Institute of Standards and Technology, Gaithersburg, MD, USA, 2008) was used to identify each peak. MS peak areas of HFC-134a and main products were calculated in order to estimate relative amounts.
The conversion (%) of HFC-134a was calculated by comparing the MS peak area of HFC-134a in the reactant (before entering the reactor) and in the product gas (after the reaction). This calculation is expressed in Equation (3):
C o n v e r s i o n   o f   H F C 134 a   ( % ) = ( A i n A o u t A i n ) ×   100
where, Ain and Aout are the MS peak areas of HFC-134a before entering the reactor and leaving the reactor, respectively. MS peak area of inflow (Ain) was determined using MS peak of HFC-134a in outflow at room temperature.

3.3. Characterizations

The specific surface area of the catalysts was measured via the BET equation using an Autosorb-iQ 2ST/MP apparatus (Quantachrome, Boynton Beach, FL, US). Before measurement, the samples were degassed at 300 °C for one hour. NH3-TPD measurements to determine the acidic properties of the catalysts were conducted using a Micromeritics Autochem II 2920 analyzer (Norcross, GA, USA) equipped with a thermal conductivity detector (TCD). The sample (0.1 g) was heated at 100 °C for one hour under the He atmosphere at a flow of 50 mL/min and then cooled down to 50 °C. NH3 adsorption was performed under a flow of 50 mL/min for one hour at 100 °C. Then, the physisorbed NH3 was removed by flushing with He flow (50 mL/min) for one hour at 100 °C. Desorption of NH3 was determined from 100 to 700 °C at a rate of 10 °C/min.
The changes in the crystallographic structure of fresh and used catalysts were examined by XRD using an Ultima IV system (Rigaku, Tokyo, Japan) over a 2θ scan range of 10–80° operated at a voltage and current of 40 kV and 40 mA, respectively. Elemental analysis was carried out on a FESEM (S-4800, Hitachi, Tokyo, Japan) at accelerating voltage of 15 kV equipped with EDS to determine the element content of fresh and spent catalyst and absorbent. The content of coke deposition in the used catalysts was also measured by TGA analyzer (Pyris Diamond, Perkin Elmer, Waltham, MA, USA,). For this, the coke deposited catalysts (3 mg) were heated from room temperature to 800 °C under an air atmosphere (150 mL/min) at a heating rate of 10 °C/min.

4. Conclusions

The present study was designed to determine the factors affecting catalytic activity on the conversion of HFC-134a, to identify the deactivation mechanism of the catalysts, and to develop an economic conversion process by adding MgO in order to prevent fluorination of γ-Al2O3 catalyst. The effect of γ-Al2O3, MgO, CaO, and their mixture on the catalytic decomposition of HFC-134a was investigated using a lab scale fixed-bed reactor. Among the selected catalysts, γ-Al2O3 showed the highest conversion efficiency and longest lifetime during the long-term tests due to its larger surface area and greater amount of weak acid sites than MgO and CaO. The additional use of MgO together with γ-Al2O3 further increased the lifetime of γ-Al2O3 because the conversion of γ-Al2O3 to AlF3 is delayed due to preferred reaction between HF and MgO. Compared to the double bed γ-Al2O3-MgO, mixed bed γ-Al2O3-MgO was more efficient in extending the catalyst lifetime and maintaining a high HFC-134a decomposition efficiency.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/11/901/s1, Figure S1. TG and DTG patterns of spent catalyst in (a) SB-γ-Al2O3-0.6 and (b) γ-Al2O3 and (c) MgO in DB-γ-Al2O3-0.6-MgO-0.6.

Author Contributions

Conceptualization, S.J., G.L.S., Y.-M.K. and S.K.; Funding acquisition, S.K.; Investigation, G.L.S.; Supervision, S.J. and S.K.; Writing—original draft, S.J. and G.L.S.; Writing—review & editing, J.-K.J., Y.-M.K., C.M.A.S. and S.K.

Funding

This research was funded by the Korea Ministry of Environment as Waste to Energy –Recycling Human Resource Development Project (YL-WE-17-001) and by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education(NRF-2019R1A6A3A01096378).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Intergovernmental Panels on Climate Change (IPCC). Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change; Solomon, S., Qin, D., Manning, M., Chen, Z., Marquis, M., Averyt, K.B., Tignor, M., Miller, H.L., Eds.; Cambridge University Press: Cambridge, UK, 2007. [Google Scholar]
  2. European Environment Agency (EEA). Emissions and Supply of Fluorinated Greenhouse Gases. 2018. Available online: https://www.eea.europa.eu/data-and-maps/indicators/emissions-and-consumption-of-fluorinated-2/assessment (accessed on 27 October 2019).
  3. Environmental Investigation Agency (EIA). Kigali Amendment to the Montreal Protocol: A Crucial Step in the Fight against Catastrophic Climate Change. 2016. Available online: https://eia-international.org/wp-content/uploads/EIA-Kigali-Amendment-to-the-Montreal-Protocol-FINAL.pdf (accessed on 27 October 2019).
  4. Han, E.; Li, Y.; Tang, H.; Liu, H. Treatment of the potent greenhouse gas, CHF3-an overview. J. Fluorine Chem. 2012, 140, 7–16. [Google Scholar] [CrossRef]
  5. Jasinski, M.; Dors, M.; Mizeraczyk, J. Destruction of Freon HFC-134a Using Nozzless Microwave Plasma Source. Plasma Chem. Plasma P 2009, 29, 363–372. [Google Scholar] [CrossRef]
  6. Ohno, M.; Ozawa, Y.; Ono, T. Decomposition of HFC134a Using Arc Plasma. Int. J. Plasma Environ. Sci. Technol. 2007, 1, 159–165. [Google Scholar]
  7. Han, T.U.; Yoo, B.-S.; Kim, Y.-M.; Hwang, B.; Subidya, 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]
  8. Ryu, J.; No, K.; Kim, Y.; Park, E.; Hong, S. Synthesis and application of ferroelectric poly (vinylidene fluoride-co-trifluoroethylene) films using electrophoretic deposition. Sci. Rep. 2016, 6, 36176. [Google Scholar] [CrossRef] [PubMed]
  9. Choi, Y.-Y.; Kwak, K.; Seo, J.-W.; Park, M.; Paik, H.; Lee, J.-Y.; Hong, J.; No, K. Ferroelectric nanodot formation from spin-coated poly(vinylidene fluoride-co-trifluoroethylene) films and their application to organic solar cells. J. Appl. Polym. Sci. 2015, 132, 41230. [Google Scholar] [CrossRef]
  10. Brown, L.F.; Carlson, R.L.; Sempsrott, J.M. Spin-cast P (VDF-TrFE) films for high performance medical ultrasound transducers. In Proceedings of the 1997 IEEE Ultrasonics Symposium Proceedings. An International Symposium (Cat. No.97CH36118), Toronto, ON, Canada, 5–8 October 1997; pp. 1725–1727. [Google Scholar]
  11. Higashi, Y.; Sakoda, N.; Islam, M.A.; Takada, Y.; Koyama, S.; Akasaka, R. Measurements of Saturation Pressures for Trifluoroethene (R1123) and 3,3,3-Trifluoropropene (R1243zf). J. Chem. Eng. Data 2018, 63, 417–421. [Google Scholar] [CrossRef]
  12. Jia, W.; Liu, M.; Lang, X.; Hu, C.; Li, J.; Zhu, Z. Catalytic dehydrofluorination of 1,1,1,2-tetrafluoroethane to synthesize trifluoroethylene over a modified NiO/Al2O3 catalyst. Catal. Sci. Technol. 2015, 5, 3103–3107. [Google Scholar] [CrossRef]
  13. 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]
  14. Xu, X.; Sun, L.; Wang, Y. NF3 decomposition over Al2O3 reagents without water. J. Nat. Gas Chem. 2015, 20, 418–422. [Google Scholar] [CrossRef]
  15. Gandhi, M.S.; Mok, Y.S. Effect of packing materials on the decomposition of tetrafluoroethane in a packed-bed dielectric barrier discharge plasma reactor. Int. J. Environ. Sci. Technol. 2015, 12, 499–506. [Google Scholar] [CrossRef]
  16. Iengo, P.; Serio, M.D.; Sorrentino, A.; Solinas, V.; Santacesaria, E. Preparation and properties of new acid catalysts obtained by grafting alkoxides and derivatives on the most common supports. Appl. Catal. A Gen. 1998, 167, 85–101. [Google Scholar] [CrossRef]
  17. 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]
  18. Krahl, T.; Keminitz, E. Aluminium fluoride—The strongest solid Lewis acid: Structure and reactivity. Catal. Sci. Technol. 2017, 7, 773–796. [Google Scholar] [CrossRef]
  19. Kemnitz, E.; Menz, D.-H. Fluorinated metal oxides and metal fluorides as heterogeneous catalysts. Prog. Solid State Chem. 1998, 26, 97–153. [Google Scholar] [CrossRef]
  20. Han, J.-Y.; Kim, C.-H.; Lee, B.; Nam, S.-C.; Jung, H.-Y.; Lim, H.; Lee, K.-Y.; Ryi, S.-K. Sorption enhanced catalytic CF4 hydrolysis with a three-stage catalyst-adsorbent reactor. Front. Chem. Sci. Eng. 2017, 11, 537–544. [Google Scholar] [CrossRef]
Figure 1. Short-term conversion efficiencies of HFC-134a obtained from the non-catalytic and catalytic conversion of HFC-134a over γ-Al2O3, MgO, and CaO at different temperatures.
Figure 1. Short-term conversion efficiencies of HFC-134a obtained from the non-catalytic and catalytic conversion of HFC-134a over γ-Al2O3, MgO, and CaO at different temperatures.
Catalysts 09 00901 g001
Figure 2. The changes of HFC-134a conversion efficiencies during continuous reaction of HFC-134a over γ-Al2O3, MgO, and CaO for 12 h at 600 °C.
Figure 2. The changes of HFC-134a conversion efficiencies during continuous reaction of HFC-134a over γ-Al2O3, MgO, and CaO for 12 h at 600 °C.
Catalysts 09 00901 g002
Figure 3. NH3-TPD curves of fresh catalysts used in this study.
Figure 3. NH3-TPD curves of fresh catalysts used in this study.
Catalysts 09 00901 g003
Figure 4. The changes of HFC-134a conversion efficiencies during long-term test over different catalyst beds for 12 h at 600 °C (SB-γ-Al2O3-1.2, single bed-γ-Al2O3 (1.2 g); DB-γ-Al2O3-0.6-MgO-0.6, double bed- γ-Al2O3 (0.6 g) + MgO (0.6 g); MB-γ-Al2O3-0.6-MgO-0.6, mixed bed-γ-Al2O3 (0.6 g) + MgO (0.6 g); SB-γ-Al2O3-0.6, single bed-γ-Al2O3 (0.6 g)).
Figure 4. The changes of HFC-134a conversion efficiencies during long-term test over different catalyst beds for 12 h at 600 °C (SB-γ-Al2O3-1.2, single bed-γ-Al2O3 (1.2 g); DB-γ-Al2O3-0.6-MgO-0.6, double bed- γ-Al2O3 (0.6 g) + MgO (0.6 g); MB-γ-Al2O3-0.6-MgO-0.6, mixed bed-γ-Al2O3 (0.6 g) + MgO (0.6 g); SB-γ-Al2O3-0.6, single bed-γ-Al2O3 (0.6 g)).
Catalysts 09 00901 g004
Figure 5. XRD results of fresh and spent γ-Al2O3 used in SB-γ-Al2O3-0.6.
Figure 5. XRD results of fresh and spent γ-Al2O3 used in SB-γ-Al2O3-0.6.
Catalysts 09 00901 g005
Figure 6. The XRD results of fresh and spent (a) γ-Al2O3 and (b) MgO used in DB-γ-Al2O3-0.6-MgO-0.6.
Figure 6. The XRD results of fresh and spent (a) γ-Al2O3 and (b) MgO used in DB-γ-Al2O3-0.6-MgO-0.6.
Catalysts 09 00901 g006
Figure 7. Catalyst bed configurations.
Figure 7. Catalyst bed configurations.
Catalysts 09 00901 g007
Figure 8. Schematic diagram of experimental setup for the catalytic decomposition of HFC-134a.
Figure 8. Schematic diagram of experimental setup for the catalytic decomposition of HFC-134a.
Catalysts 09 00901 g008
Table 1. The amount of weak, intermediate, strong, and total acid sites of fresh γ-Al2O3, MgO, and CaO.
Table 1. The amount of weak, intermediate, strong, and total acid sites of fresh γ-Al2O3, MgO, and CaO.
CatalystAmount of Acid Sites (mmol-NH3/g-Catalyst)
WeakIntermediateStrongTotal
γ-Al2O31.7762.3452.6236.744
MgO0.6140.7530.3971.764
CaO0.0041.6960.7972.497
Table 2. EDS results for fresh and spent γ-Al2O3 used in the single-bed reactor.
Table 2. EDS results for fresh and spent γ-Al2O3 used in the single-bed reactor.
Sample Conditionsγ-Al2O3 (Element, wt%)
AlOFC
Fresh43.552.004.5
After 6 h21.126.736.116.1
After 9 h18.82.461.117.7
After 12 h10.31.540.847.4
Table 3. EDS results for fresh and spent γ-Al2O3 and MgO used in the serial bed configuration.
Table 3. EDS results for fresh and spent γ-Al2O3 and MgO used in the serial bed configuration.
Sample Conditionsγ-Al2O3 (Element, wt%)MgO (Element, wt%)
AlOFCMgOFC
Fresh43.552.004.546.348.605.1
After 6 h40.530.920.18.544.525.530.00
After 9 h36.830.814.717.732.78.855.33.2
After 12 h12.54.843.639.131.51.659.37.6

Share and Cite

MDPI and ACS Style

Jeong, S.; Sudibya, G.L.; Jeon, J.-K.; Kim, Y.-M.; Swamidoss, C.M.A.; Kim, S. The Use of a γ-Al2O3 and MgO Mixture in the Catalytic Conversion of 1,1,1,2-Tetrafluoroethane (HFC-134a). Catalysts 2019, 9, 901. https://doi.org/10.3390/catal9110901

AMA Style

Jeong S, Sudibya GL, Jeon J-K, Kim Y-M, Swamidoss CMA, Kim S. The Use of a γ-Al2O3 and MgO Mixture in the Catalytic Conversion of 1,1,1,2-Tetrafluoroethane (HFC-134a). Catalysts. 2019; 9(11):901. https://doi.org/10.3390/catal9110901

Chicago/Turabian Style

Jeong, Sangjae, Gamal Luckman Sudibya, Jong-Ki Jeon, Young-Min Kim, Caroline Mercy Andrew Swamidoss, and Seungdo Kim. 2019. "The Use of a γ-Al2O3 and MgO Mixture in the Catalytic Conversion of 1,1,1,2-Tetrafluoroethane (HFC-134a)" Catalysts 9, no. 11: 901. https://doi.org/10.3390/catal9110901

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