Next Article in Journal
Helium: Sources, Applications, Supply, and Demand
Previous Article in Journal
Alternative Gaseous Fuels for Marine Vessels towards Zero-Carbon Emissions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Greenhouse Gas Conversion into Hydrocarbons and Oxygenates Using Low Temperature Barrier Discharge Plasma Combined with Zeolite Catalysts

A.V. Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences (TIPS RAS), Moscow 119991, Russia
*
Author to whom correspondence should be addressed.
Gases 2023, 3(4), 165-180; https://doi.org/10.3390/gases3040012
Submission received: 18 August 2023 / Revised: 23 November 2023 / Accepted: 4 December 2023 / Published: 5 December 2023
(This article belongs to the Section Gas Emissions)

Abstract

:
Global warming occurs as a result of the build-up of greenhouse gases in the atmosphere, causing an increase in Earth’s average temperature. Two major greenhouse gases (CH4 and CO2) can be simultaneously converted into value-added chemicals and fuels thereby decreasing their negative impact on the climate. In the present work, we used a plasma-catalytic approach for the conversion of methane and carbon dioxide into syngas, hydrocarbons, and oxygenates. For this purpose, CuCe zeolite-containing catalysts were prepared and characterized (low-temperature N2 adsorption, XRF, XRD, CO2-TPD, NH3-TPD, TPR). The process of carbon dioxide methane reforming was conducted in a dielectric barrier discharge under atmospheric pressure and at low temperature (under 120 °C). It was found that under the studied conditions, the major byproducts of CH4 reforming are CO, H2, and C2H6 with the additional formation of methanol and acetone. The application of a ZSM-12 based catalyst was beneficial as the CH4 conversion increased and the total concentration of liquid products was the highest, which is related to the acidic properties of the catalyst.

1. Introduction

The accumulation of greenhouse gases in the Earth’s atmosphere has become a prominent environmental concern in recent decades due to its profound impact on our planet’s atmosphere and climate. Greenhouse gases, such as carbon dioxide, methane, nitrous oxide, and fluorinated gases, play a crucial role in regulating the Earth’s temperature through the greenhouse effect. However, human activities (burning of fossil fuels, deforestation, and industrial processes) have significantly increased the concentration of these gases, leading to a disbalance in the natural climate system. Consequently, the concentration levels of CO2 and other greenhouse gases have reached unprecedented levels in the last century [1].
The joint processing of greenhouse gases, specifically methane and carbon dioxide, can be advantageous in terms of greenhouse gas reduction, energy, fuels and chemicals production [2,3]. One of the main methods to convert both methane and carbon dioxide in a single process is the dry methane reforming (DRM) process [4,5]. Dry methane reforming offers numerous opportunities for the production of valuable intermediates and chemicals. The product of this reaction is a syngas, which consists of CO and H2. The resulting syngas can be further processed to obtain various fuels and chemicals. For example, Fischer–Tropsch synthesis can convert syngas into synthetic hydrocarbons such as gasoline, diesel, and waxes. Alternatively, the syngas can serve as a precursor for the production of methanol, ammonia, or other chemical compounds [6,7]. However, there are some limitations regarding DRM. The reaction is highly endothermic and requires elevated temperatures for efficient conversion. Carbon deposition is also a common issue, where carbonaceous species can accumulate on the surface of the catalyst, leading to catalyst deactivation [4]. Therefore, effective catalyst design and reactor engineering are crucial to prevent carbon deposition and maintain catalytic activity over extended periods. Traditionally, transition metals like Ni or Co are widely studied in DRM due to their availability and low cost [8,9]. However, its coking resistance ability is relatively low, so numerous catalyst modification attempts are being made. For example, Ni-Co bimetallic mullite catalysts [10], Zr-La-Ti oxide catalysts [11], as well as Ga-doped Ni/CeO2 [12] are reported to show remarkable stability against carbon deposition. As an alternative to thermal-induced dry methane reforming, photocatalytic [13,14] and photo-thermal [15] DRM show great potential for reducing process temperature and enhancing gas conversion.
Another way to overcome the limitations of thermal-catalytic DRM is a plasma-catalytic approach [16]. This process combines the advantages of plasma technology and catalysis to enhance the efficiency and selectivity of the reforming reaction [17]. Plasma can activate the reactant molecules, enhancing their reactivity, and reducing the energy barrier for the reactions. The main advantage of non-thermal plasma is that it makes it possible to conduct thermodynamically unfavorable reactions at low temperatures due to the non-equilibrium character of plasma [18]. In comparison to thermal-activated processes, in non-thermal plasma, the vibrational excitation and the energy transfer from one active species (i.e., electron, ion, radical) to the neutral molecule are the main pathways of molecule activation [19]. The presence of the catalyst helps to improve overall reaction efficiency by providing active sites for the adsorption and activation of reactant molecules. Additionally, plasma can regenerate the catalyst by removing deposited carbon species, thus extending its lifetime.
One of the attractive ways to tune plasma-catalytic dry methane reforming is to use the process not only for syngas production but for oxygenate production via a single-stage process [20,21,22,23,24]. In order to achieve that, the selection of the catalyst should be carefully performed. In the dry methane reforming reaction, methane decomposition occurs on the acidic part of the support [25], so one of the requirements for the catalyst support is the presence of acid centers on the surface. The most commonly used materials for supports are Al2O3, SiO2, and zeolites, which may be modified by TiO2, ZrO2, and CeO2 [20,26,27,28]. Zeolites are structured aluminosilicates, which possess acid properties, and the acidity may be widely varied. With the variation of the acidity in the catalyst, different reaction products can be obtained. For example, a higher concentration of Brønsted acid centers in an HZSM-5-based catalyst for plasma-catalytic DRM leads to the formation of acetic acid, while a catalyst with less acidic properties exhibits higher activity towards methanol formation [29]. As previously stated [30], the course of the reaction is also affected by the presence of oxygen vacancies on the surface; in particular, the addition of magnesium-aluminum oxide to the catalyst increases CO2 conversion due to the formation of the oxygen vacancies. As reported in the literature [31,32], the textural properties of the supports are important as microporosity inhibits the formation of micro discharges inside the pores and thus decreases the conversion. Regarding the active phase, such metals as Ni, Co, Mo, V, Na, Cu, Zn, Pt, Ag, or their combination [33] are commonly used [34,35,36]. It is reported that the adsorption and dissociation of methyl radicals and their subsequent recombination with other particles is particularly effective on copper and zinc centers [29,37,38,39]. The advantage of Cu-based catalysts in plasma-catalytic dry methane reforming is attributed to high activity in C–H bond splitting in CH4 and the subsequent C–C coupling resulting in a higher oxygenate yield [40].
In the present study, we aimed to compare the activity of zeolite-based Cu catalysts for plasma-catalytic dry methane reforming. A zeolite of the MTW (ZSM-12) type was chosen for comparison with the widely used MFI (ZSM-5) zeolite due to its unique porous structure and tuning acidity availability [41]. To the best of our knowledge, the MTW-type zeolite was not studied in plasma-catalytic dry methane reforming so far by other researchers. The zeolites were combined with Al2O3 in order to increase the porous characteristics and to create a shaped body catalyst. A preliminary study was made by our group [42], where Ni- and Co-ZSM-12 based catalysts showed better plasma-catalytic performance compared to ZSM-5-based catalysts. In the current research, we focus on the Cu-containing catalysts and the evaluation of their activity and stability in the plasma-catalytic dry methane reforming process.

2. Materials and Methods

The reagents, which were used for the preparation of the catalysts, are summarized in Table 1. All the reagents were used without further purification.

2.1. Preparation of the Catalysts

The synthesis of the ZSM-12 zeolite was conducted according to the procedure reported before by our group [43]. In a typical synthesis, a gel was prepared by mixing the solutions of Al2(SO4)3·18H2O, NaOH, N-Methyldiethanolamine, and a colloid silica solution. The obtained gel was further hydrothermally treated at 155 °C for 120 h, then washed, dried, and calcined. After several cycles of an ion-exchange procedure, an H-form of the ZSM-12 zeolite was obtained.
For the preparation of the catalyst support, the obtained ZSM-12 zeolite and a commercial ZSM-5 zeolite were used. Boehmite (AlOOH) was used as a binder to obtain a shaped body support. The ratio of zeolite-to-binder was 70/30 by mass. A portion of boehmite (10.7 g) was mixed with 25 g of zeolite, then ground in a mortar to obtain a homogeneous mixture. After that, HNO3 (21 cm3, 1 M) was added until a viscous mass was formed. The mass was further extruded, dried (60 °C for 2 h, 80 °C for 2 h, 110 °C for 2 h), and calcined (550 °C for 4 h). The obtained granules of support were additionally chopped to a size of 0.8–1.5 cm. Based on the zeolite used, two supports were obtained which were designated as ZSM-5/Al2O3 and ZSM-12/Al2O3.
The catalyst support was impregnated with a solution of Cu(NO3)2·3H2O and Ce(NO3)2·6H2O salts with further drying and calcination. The catalyst supported on the ZSM-5/Al2O3 support was designated as CuCe-5 and the catalyst supported on the ZSM-12/Al2O3 support was designated as CuCe-12. Prior to the plasma-catalytic experiments, the catalysts were reduced in H2 flow at 400 °C for 2 h.

2.2. Catalysts Characterization

The specific surface area, pore volume, and pore diameter of the prepared supports and the catalysts were determined by low-temperature N2 adsorption analysis using a Belsorp miniX (Microtrac MRB, Osaka, Japan) surface area and pore size distribution analyzer. Before analysis, the catalysts (supports) were thermally degassed for 8 h (T = 300 °C, p = 10 Pa). The Brunauer–Emmett–Teller method was used for specific surface area calculation at p/p0 = 0.05–0.2.
The elemental composition of the catalysts was determined by X-ray fluorescent spectroscopy using an ARL Perform’s Sequental (Thermo Fisher Scientific, Ecublens, Switzerland) analyzer (X-ray tube power is 2500 W). The sample, before analysis, was ground and pressed into a tablet with H3BO3.
The X-ray diffraction patterns were registered with a range of 2θ = 10°–90° using a Rigaku Rotaflex RU-200 (Tokyo, Japan) diffractometer (CuKα radiation) equipped with a Rigaku D/Max-RC goniometer (a rotation speed of 1°/min; a step 0.04°). The diffractograms were identified using the PDF-2 ICDD database.
The average size (D) of crystallites was calculated using the Scherrer equation:
D = K λ β c o s θ
where D is the crystallite size, λ is the wavelength of the Cu-Kα radiation, K is a constant and its value is taken as 0.9, θ is the diffraction angle (rad), and β is the full-width at half maximum (FWHM) (rad).
The acidity of the samples was measured by NH3 temperature-programmed desorption (NH3-TPD) using a USGA-101 (LLC “Unisit”, Moscow, Russia) device. The sample (m = 0.15–0.2 g) was treated in He flow in order to purify the surface from the adsorbed water and oxygen (T = 512 °C for 40 min). Then, the sample was saturated with NH3 (5 vol% NH3—95 vol% He) at 60 °C for 24 min. The analysis was conducted in helium flow at 100–800 °C (heating rate 7 °C/min). Desorbed NH3 registration was carried out using a thermal conductivity detector. The TPD profiles were deconvoluted using PeakFit (v4.12) software.
The temperature-programmed reduction of the catalysts was carried out using a USGA-101 device. Prior to analysis, the sample was calcined at 500 °C for 30 min in Ar flow, then cooled down to 60 °C. The reduction of the sample was conducted at the 10 °C/min rate up to 950 °C.

2.3. Plasma-Catalytic Gases Utilization

The process of carbon dioxide methane reforming was studied using a plasma-catalytic dielectric barrier discharge reactor unit (Figure 1). The mixture of CH4 and CO2 was fed into the reactor using RRG-20 precise mass-flow controllers (LLC “Eltochpribor”, Zelenograd, Russia). The flow rate of the gases was set to 23.5 mL/min each and the total gas flowrate was 47 mL/min. A dielectric barrier discharge was formed in a quartz tube (16 mm outer diameter, 2 mm wall thickness) between the two electrodes. A steel rod (8 mm diameter) with a tread was placed inside the reactor (inner electrode) and a steel mesh (0.5 mm mesh size, 80 mm length) was placed on the outer wall of the quartz tube (ground electrode). The catalyst (1 g) was placed between the inner electrode and the inner wall of the tube and was fixed with mineral wool at both ends of the packed bed. A high voltage generator had a sine wave signal with a frequency of 23 kHz and was connected to the inner electrode. The electric signals were registered using a Tektronix TDS 2012B oscilloscope (Tektronix, Beaverton, OR, USA). Each experiment was conducted for 2 h with a gradual increase of voltage from 3.5 to 4.8 kV.
Based on the Volt–Columb characteristics of the discharge (Lissajous figure), the plasma absorbed power was calculated according to the equation:
P = f W = f C n A ,
where Cn is the value of the capacitor included in series with the discharge tube, f is a frequency of the applied voltage, and A is the area of a Lissajous figure.
The energy efficiency of the process (η) was calculated as a ratio of converted gas (CH4 or CO2) to the absorbed power using the following equation:
η m m o l × k J 1 = ν c o n v P × 1000 60   ,
where ν c o n v is the quantity of the gas converted (mol/min) and P is the absorbed power (W).
The gaseous outlet mixture was analyzed using a PIA gas chromatograph (LLC “NPF MEMS”, Samara, Russia) with a thermal conductivity detector and equipped with a Hayesep N adsorbent column (l = 2 m) for the determination of CO2 and C2H6 and a molecular sieve 13 Å column (l = 2 m) for the determination of H2, CO, CH4. The content of the gases was calculated based on the calibration curves.
Liquid products were trapped in an ice-chilled vessel which contained 20 mL of distilled water. The product flow from the reactor outlet bubbled through the vessel, so the liquid products were dissolved in water. The obtained solution was analyzed using a Crystallux-4000M (RPC “META-CHROM”, Yoshkar-Ola, Russia) gas chromatograph with a Poraplot Q capillary column (25 m × 0.53 mm).
Based on the data obtained during the chromatographic analysis, the conversion (X), product selectivity (S), and product yield (Y) were calculated according to the equations:
X C O 2 % = ν C O 2 i n   ν C O 2 o u t ν C O 2 ( i n ) × 100 %
X C H 4 % = ν C H 4 i n   ν C H 4 o u t ν C H 4 i n × 100 %
S C O % =   ν C O p r o d ν C O 2 c o n v + ν C H 4 c o n v   × 100 %
S H 2 % =   ν H 2 p r o d 2 × ν C H 4 c o n v   × 100 %
S C 2 H 6 % =   ν C 2 H 6 p r o d ν C O 2 c o n v + ν C H 4 c o n v   × 100 %
Y C O % =   ν C O p r o d ν C O 2 i n + ν C H 4 i n   × 100 %
Y H 2 % =   ν H 2 p r o d 2 × ν C H 4 i n   × 100 %
where ν i n   is the quantity of gas (CO2/CH4) which was injected into the reactor (mol), ν o u t is the quantity of the gas in the outlet stream (mol),   ν p r o d is the quantity of the gas produced (mol), ν c o n v , and is the quantity of the gas (CO2/CH4) which was converted to the products during the reaction.

3. Results and Discussion

3.1. Zeolite Catalyst Synthesis and Characterization

The prepared catalyst supports and the impregnated catalysts were analyzed with various methods to investigate their physicochemical characteristics. According to the low-temperature N2 adsorption, the prepared zeolite supports had a relatively high specific surface area, which decreased slightly after the addition of the metals (Table 2). The pore volume was the same as in the support in case of the CuCe-5 and the CuCe-12 samples, which can indicate that the metal oxides were mostly located on the surface of the support rather than inside the pores.
The studied adsorption isotherms of the ZSM-5 and ZSM-12 zeolites (Figure 2) according to the IUPAC classification belong to type I isotherms, which indicates the microporosity of the zeolites. The adsorption isotherms of the ZSM-5/Al2O3 and ZSM-12/Al2O3 supports as well as the CuCe-5 and CuCe-12 catalysts also belong to type I in the range of low relative pressures, and to isotherm IV in the range of high relative pressures. Thus, the supports and the catalysts contain both micro- and mesopores. The presence of the mesopores is explained by the addition of Al2O3 in order to form a shaped body catalyst.
The pore size distribution of the supports and the catalyst is shown on Figure 3. For each sample, two peaks are present on the graph. The peak at 3.9 nm does not reflect the actual porous characteristics of the samples, and corresponds to the tensile strength effect, which is determined by the properties of the adsorptive [44]. Thus, only the peaks at 5.7 nm (for ZSM-5-based samples) and 6.7 nm (for ZSM-12-based samples) are taken into consideration. It can be seen from the plots that the ZSM-5-based samples have a narrower pore size distribution: the average pore diameter (6.1 nm for ZSM-5/Al2O3 and 6.4 nm for CuCe-5) is close to the data from the BJH plot. For the ZSM-12 based samples, the pore size distribution is wider (peak at 6.7–7.0 nm), which means a less uniform distribution (average pore size 9.1–10.2 nm).
The acidity of the samples was determined by temperature-programmed NH3 desorption. It was stated that the ZSM-5-based samples had more acid sites, which is attributed to Si/Al ratio in the parent zeolites. As the Si/Al ratio tends to diminish (as in the ZSM-5 zeolite case), the acidity increases (Table 3). The TPD-patterns were deconvoluted to obtain three desorption peaks: the low-temperature (200–210 °C), the moderate-temperature (~250 °C) and the high-temperature peak at ~400 °C (Figure 4). According to the literature [45], the low-temperature peak can be ascribed to physically adsorbed ammonia (weak acid sites), so the true acid sites (medium and strong acid sites) can be described by the moderate- and high-temperature peak. It can be seen from the figure that with the addition of Cu and Ce to the support, the medium-temperature peak is shifted towards higher temperature and the area peak is increased compared to that in the TPD pattern of the corresponding supports. This can be related to the effect of the sequential desorption of the NH3 which is attached to Cu ions [46].
On the X-ray diffractograms, it can be seen that the zeolite structure was not destroyed during metal loading (Figure 5). The diffraction patterns in the range of 2θ = 5°–30° are attributed to the zeolite phase and the intensity of these patterns does not change significantly.
The average size of the crystallites was determined for the CeO2 and CuO phases (Table 4). It should be noted that the CuO phase in CuCe-5 sample is not distinguished, which may be the result of the high dispersion of that phase on the ZSM-5/Al2O3 support surface. This support has a higher specific surface area, which leads to a better particle distribution. However, CeO2 crystallites are of a similar size in both catalysts, which indicates no effect of the support properties on the CeO2 crystallite size.
The temperature-programmed H2 reduction of the obtained samples was conducted in order to investigate the reducibility of the catalysts prior to the plasma-catalytic experiments. The reduction temperature of the pure CuO is in the range of 300–400 °C depending on the analysis conditions, as reported in the literature [47]. The reduction of pure CeO2 is observed in a temperature range of 350–580 °C and above 900 °C [48,49]. It is seen from the TPR profiles (Figure 6) that the reduction temperature is lower than that of the pure CuO and CeO2 substances. In the literature, it is stated that CuO–CeO2 systems have a lower reduction temperature because of the interaction of the two oxides [50]. In the TPR pattern of the CuCe-5 sample, only one high-intensity peak is present, which may indicate higher CuO particle dispersion [51], which confirms the XRD results. The TPR pattern of the CuCe-12 differs from the one of CuCe-5 and three peaks are distinguished in the TPR profile. According to the literature [52], the low-temperature peak (~150 °C) is attributed to the reduction of CuOx, which interacts strongly with CeO2, the peak at ~210 °C is attributed to the reduction of highly dispersed CuOx particles on the surface, and the high-temperature peak (~260 °C) is attributed to bulk CuOx particles. The latter is also consistent with the CuO crystallite size data obtained by the XRD analysis.

3.2. Gas Conversion Using Plasma-Catalytic Process

In order to test the activity of the prepared catalysts, a dielectric barrier discharge plasma reactor was used. As a blank test, an experiment in an empty reactor without a catalyst was conducted. On Figure 7, the Lissajous figures, which were registered during the blank test, are presented.
It can be seen that with the increase in the input power, the shape of the figure (parallelogram) remains unchanged, but the figure itself is broadened, which is due to enhanced voltage and current signals. The observed shape of the Lissajous figure is typical for dielectric barrier discharge plasma [53]. When the catalyst is introduced into the discharge zone, the input power is decreased Figure 8d. The discharge behavior in the absence of the catalyst differs from the one in the empty reactor due to the change in the dielectric properties of the matter between the inner and the outer electrodes.
The supports (ZSM-5/Al2O3 and ZSM-12/Al2O3) along with the catalysts were also tested in plasma-catalytic CH4 reforming. From the results, it can be seen that the conversion of CO2 is decreased when the packing material is introduced. As carbon dioxide is a gas with acidic properties, a less acidic catalyst would be preferable. However, when material with less acidity was present (ZSM-12 based catalyst), no significant change in CO2 conversion was observed. Apparently, better adsorption would be achieved on strong basic sites which were not determined in the prepared catalysts. A minor increase in the CO2 conversion was achieved when the CeO2 was impregnated compared to the pure support. Cerium oxide possesses oxygen vacancies in the structure, which promote the splitting of CO2 into CO [54]. The conversion of CH4 increased and was the highest in presence of the CuCe-12 sample. When relating the conversion to the input power, it was seen that in presence of CuCe-5 and CuCe-12, the input power was similar but CH4 conversion in the presence of CuCe-12 was higher, which is attributed to the synergy between the plasma and the catalyst. The activated methane molecule bridges to the Cu atom with subsequent recombination to gaseous and liquid products. Apparently, the higher surface area and the higher distribution of the CuO species on the CuCe-5 surface did not provide much of an effect on CH4 conversion compared to the bigger CuO crystallites in the case of CuCe-12. The products of CH4-CO2 reforming included H2, CO, and C2H6. The yields of the products and their selectivities are presented in Figure 8b,c. It can be seen that the product yields are the highest in the case of an empty reactor, which may be attributed to the higher input power. As was mentioned before, in the presence of packing materials, the consumed power slightly decreases, which may affect product formation. It should be noted that the CuCe-12 sample exhibits the lowest selectivity for gaseous products, which may be related to the formation of liquid products.
The liquid products, which were produced during plasma-catalytic processing of the gases, were collected into a beaker containing distilled water to prevent evaporation. The resulting water solution was subjected to a gas chromatography analysis and the main products, which were methanol and acetone, were determined. From Figure 9 it may be observed that the total quantity of the oxygenated products was the highest in presence of the CeCe-12 catalyst. This is consistent with the suggestion that the decreased selectivity of the gaseous products (H2, CO, C2H6) is attributed to the formation of liquid products. When comparing the two types of catalysts, it can be seen that more liquid products were produced in the presence of the ZSM-12-supported catalyst. A possible explanation can be the difference in acid properties. The catalyst with less acid centers promotes the formation of methanol rather than carboxylic acids [38]. During the plasma activation of the CH4–CO2 gas mixture, CO2*, CH4*, and CO* excited species are formed as well as CH3, CH2, CH, H, and O radicals [55]. As a result of species recombination in the plasma and on the surface of the catalyst, OH, HCOads, H2COads, and H3COads are produced, which further transform into molecules like CH3OH. According to the literature [56], the presence of acetone in the products is attributed to HCOads species which couple with C2 species to form a C3-intermediate resulting in acetone formation.
The results obtained were compared to previously published investigations on plasma-catalytic DRM and CO2 decomposition (Table 5). It can be seen from the table that in the present work, the conversion of CO2 and CH4 was comparable to that in presence of the Cu/Al2O3 catalyst from the literature [57], but the energy efficiency in our system was considerably higher (0.8 kJ/mmol vs. 0.18 kJ/mmol). It can also be seen that in the major part of the Cu-involving DRM related works, methanol is the main product among the oxygenates produced and in our research, methanol content increases when the ZSM-12 zeolite is introduced into the catalyst composition. It should be noted that the Cu/CeO2 catalyst from the recent study [38] shows the worst CO2 conversion results among other catalysts. The authors explain this phenomenon by the interaction of Cu+ species with CeO2 oxygen vacancies, which led to CO + H2O reaction activation with subsequent CO2 formation. Thus, controlling the valence state of the Cu species is essential for plasma-catalytic DRM reactions. In contrast, a CuO-based catalyst on CeAl support may lead to 13.5% CO2 conversion if the CH4 is absent, which means no side reactions occur. In the presence of our CuCe catalytic systems, the CO2 conversion is lower due to the acidic nature of the zeolite-containing support.
From the literature, it is known that acetic acid is one of the desired products in plasma-catalytic dry methane reforming. However, in our system, no amount of this acid was observed when analyzing the liquid phase. One of the possible reasons for the absence of acetic acid in the liquid products can be the limited desorption of the acid from the zeolite pores, as reported in the literature [40]. In order to overcome this limitation, reactor modification is needed in future work. For example, reactor cooling or co-reagent injection (such as O2 or water vapor) may take place.
Reusability tests of the CuCe-12 were carried out in order to investigate the stability of the catalyst. Each run duration was 2 h in order to obtain three samples of the gas. No catalyst pretreatment was made between the runs. It was revealed that the obtained sample was stable in the current conditions within the investigated time (10 h) and no drastic changes in gas conversion and the yields of H2 and CO were observed (Figure 10). It may be preliminarily concluded that catalyst deactivation did not occur in the present experiment during the 10 h of the total activity of the tests and thus the catalyst may be used in several cycles. The only problem with a continuous run is related to reactor heating, so additional technological solutions need to be implemented in future work.

4. Conclusions

In the present work, the conversion of greenhouse gases was conducted via the plasma-catalytic dry methane reforming approach using zeolite-containing catalysts. Under the studied conditions, the conversion of CH4 increases in the presence of the catalysts, but CO2 conversion slightly decreases which can be explained by the decrease of the input power in the presence of packing material and the acidic properties of the catalysts. Based on analysis results, the ZSM-5-based catalyst (CuCe-5) had a higher specific surface area, higher concentration of acid centers, and better CuO distribution. Nevertheless, CH4 conversion was higher and liquid phase content increased in the presence of CuCe-12. These effects are related to lower acidity as well as the bulk Cu phase which bridges the CH3 species and allows them to recombine with CH3OH and CH3C(O)CH3 molecules. There are still challenges to overcome in plasma-catalytic dry methane reforming. The design and optimization of the plasma-catalyst system requires careful consideration of factors such as plasma power, catalyst type, and reactor configuration. Further research and development efforts are ongoing for a better understanding of the complex interactions between plasma and catalysts and to improve the overall performance of this process.

Author Contributions

Conceptualization, A.L.M. and O.V.G.; methodology, O.V.G.; validation, O.V.G. and D.E.T.; investigation, O.V.G.; resources, O.V.G. and D.E.T.; writing—original draft preparation, O.V.G. and D.E.T.; writing—review and editing, A.L.M.; visualization, O.V.G.; supervision, A.L.M.; project administration, O.V.G. All authors have read and agreed to the published version of the manuscript.

Funding

The research was funded by the Russian Science Foundation, grant number 22-79-00259.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors thank Golubev K.B. for the chromatographic determination of the liquid samples. This work was performed using the equipment of the Shared Research Center, the Analytical Center of Deep Oil Processing and Petrochemistry of the A.V. Topchiev Institute of Petrochemical Synthesis, RAS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kostyuchenko, N.; Smolennikov, D. Greenhouse Gas Emissions. In Responsible Consumption and Production; Filho, W.L., Azul, A.M., Brandli, L., Özuyar, P.G., Wall, T., Eds.; Springer: Cham, Switzerland, 2022; pp. 324–335. [Google Scholar] [CrossRef]
  2. Lalsare, A.; Sivri, A.; Egan, R.; Vukmanovich, R.J.; Dumitrescu, C.E.; Hu, J. Biomass—Flare gas synergistic co-processing in the presence of carbon dioxide for the controlled production of syngas (H2:CO~2–2.5). Chem. Eng. J. 2020, 385, 123783. [Google Scholar] [CrossRef]
  3. Cai, X.; Hu, Y.H. Advances in catalytic conversion of methane and carbon dioxide to highly valuable products. Energy Sci. Eng. 2019, 7, 4–29. [Google Scholar] [CrossRef]
  4. Hussien, A.G.S.; Polychronopoulou, K.A. Review on the Different Aspects and Challenges of the Dry Reforming of Methane (DRM) Reaction. Nanomaterials 2022, 12, 3400. [Google Scholar] [CrossRef] [PubMed]
  5. Lavoie, J.-M. Review on dry reforming of methane, a potentially more environmentally-friendly approach to the increasing natural gas exploitation. Front. Chem. 2014, 2, 81. [Google Scholar] [CrossRef] [PubMed]
  6. An, Y.; Lin, T.; Yu, F.; Yang, Y.; Zhong, L.; Wu, M.; Sun, Y. Advances in direct production of value-added chemicals via syngas conversion. Sci. China Chem. 2017, 60, 887–903. [Google Scholar] [CrossRef]
  7. Wender, I. Synthesis Gas as a Source of Fuels and Chemicals: C-1 Chemistry. Annu. Rev. Energy 1986, 11, 295–314. [Google Scholar] [CrossRef]
  8. Han, J.W.; Park, J.S.; Choi, M.S.; Lee, H. Uncoupling the size and support effects of Ni catalysts for dry reforming of methane. Appl. Catal. B 2017, 203, 625–632. [Google Scholar] [CrossRef]
  9. Jabbour, K.; El Hassan, N.; Davidson, A.; Massiani, P.; Casale, S. Characterizations and performances of Ni/diatomite catalysts for dry reforming of methane. Chem. Eng. J. 2015, 264, 351–358. [Google Scholar] [CrossRef]
  10. Zafarnak, S.; Rahimpour, M.R. Co-Ni bimetallic supported on mullite as a promising catalyst for biogas dry reforming toward hydrogen production. Mol. Catal. 2023, 534, 112803. [Google Scholar] [CrossRef]
  11. Xu, D.-J.; Wu, J.; Liu, Z.; Qiao, L.-Y.; Zong, S.-S.; Zhou, Z.-F.; Yao, Y.-G. Doping low amount of Zirconium in Rh-LTO to prepare durable catalysts for dry reforming of methane. Mol. Catal. 2023, 535, 112822. [Google Scholar] [CrossRef]
  12. Bai, Y.; Sun, K.; Wu, J.; Zhang, M.; Zhao, S.; Kim, Y.D.; Liu, Y.; Gao, J.; Liu, Z.; Peng, Z. The Ga-promoted Ni/CeO2 catalysts for dry reforming of methane with high stability induced by the enhanced CO2 activation. Mol. Catal. 2022, 530, 112577. [Google Scholar] [CrossRef]
  13. Kushida, M.; Yamaguchi, A.; Miyauchi, M. Photocatalytic dry reforming of methane by rhodium supported monoclinic TiO2-B nanobelts. J. Energy Chem. 2022, 71, 562–571. [Google Scholar] [CrossRef]
  14. Tahir, M.; Ali Khan, A.; Bafaqeer, A.; Kumar, N.; Siraj, M.; Fatehmulla, A. Highly Stable Photocatalytic Dry and Bi-Reforming of Methane with the Role of a Hole Scavenger for Syngas Production over a Defective Co-Doped g-C3N4 Nanotexture. Catalysts 2023, 13, 1140. [Google Scholar] [CrossRef]
  15. Zhang, Z.-Y.; Li, T.; Yao, J.-L.; Xie, T.; Xiao, Q. Mechanism and kinetic characteristics of photo-thermal dry reforming of methane on Pt/mesoporous-TiO2 catalyst. Mol. Catal. 2023, 535, 112828. [Google Scholar] [CrossRef]
  16. Mehta, P.; Barboun, P.; Go, D.B.; Hicks, J.C.; Schneider, W.F. Catalysis Enabled by Plasma Activation of Strong Chemical Bonds: A Review. ACS Energy Lett. 2019, 4, 1115–1133. [Google Scholar] [CrossRef]
  17. Vadikkeettil, Y.; Subramaniam, Y.; Murugan, R.; Ananthapadmanabhan, P.V.; Mostaghimi, J.; Pershin, L.; Batiot-Dupeyrat, C.; Kobayashi, Y. Plasma assisted decomposition and reforming of greenhouse gases: A review of current status and emerging trends. Renew. Sustain. Energy Rev. 2022, 161, 112343. [Google Scholar] [CrossRef]
  18. Abiev, R.S.; Sladkovskiy, D.A.; Semikin, K.V.; Murzin, D.Y.; Rebrov, E.V. Non-Thermal Plasma for Process and Energy Intensification in Dry Reforming of Methane. Catalysts 2020, 10, 1358. [Google Scholar] [CrossRef]
  19. Malik, M.I.; Achouri, I.E.; Abatzoglou, N.; Gitzhofer, F. Intensified performance of methane dry reforming based on non-thermal plasma technology: Recent progress and key challenges. Fuel Process. Technol. 2023, 245, 107748. [Google Scholar] [CrossRef]
  20. Liu, S.; Winter, L.R.; Chen, J.G. Review of Plasma-Assisted Catalysis for Selective Generation of Oxygenates from CO2 and CH4. ACS Catal. 2020, 10, 2855–2871. [Google Scholar] [CrossRef]
  21. Wang, L.; Yi, Y.; Wu, C.; Guo, H.; Tu, X. One-Step Reforming of CO2 and CH4 into High-Value Liquid Chemicals and Fuels at Room Temperature by Plasma-Driven Catalysis. Angew. Chem. Int. Ed. 2017, 56, 13679–13683. [Google Scholar] [CrossRef]
  22. Ban, T.; Yu, X.-Y.; Kang, H.-Z.; Zhang, H.-X.; Gao, X.; Huang, Z.-Q.; Chang, C.-R. Design of Single-Atom and Frustrated-Lewis-Pair dual active sites for direct conversion of CH4 and CO2 to acetic acid. J. Catal. 2022, 408, 206–215. [Google Scholar] [CrossRef]
  23. Lia, D.; Rohania, V.; Fabrya, F.; Ramaswamya, A.P.; Sennourb, M.; Fulcheri, L. Direct conversion of CO2 and CH4 into liquid chemicals by plasma-catalysis. Appl. Catal. B 2020, 261, 118228. [Google Scholar] [CrossRef]
  24. Puliyalil, H.; Jurkovic, D.L.; Dasireddy, V.D.B.C.; Likozar, B. A review of plasma-assisted catalytic conversion of gaseous carbon dioxide and methane into value added platform chemicals and fuels. RSC Adv. 2018, 8, 27481. [Google Scholar] [CrossRef] [PubMed]
  25. Shi, C.; Wang, S.; Ge, X.; Deng, S.; Chen, B.; Shen, J. A review of different catalytic systems for dry reforming of methane: Conventional catalysis-alone and plasma-catalytic system. J. CO2 Util. 2021, 46, 101462. [Google Scholar] [CrossRef]
  26. Marinho, A.L.A.; Toniolo, F.S.; Noronha, F.B.; Epron, F.; Duprez, D.; Bion, N. Highly active and stable Ni dispersed on mesoporous CeO2-Al2O3 catalysts for production of syngas by dry reforming of methane. Appl. Catal. B 2021, 281, 119459. [Google Scholar] [CrossRef]
  27. Bouchoul, N.; Touati, H.; Fourr’e, E.; Clacens, J.-M.; Batonneau-Gener, I.; Batiot-Dupeyrat, C. Plasma-catalysis coupling for CH4 and CO2 conversion over mesoporous macroporous Al2O3: Influence of the physico-chemical properties. Appl. Catal. B 2021, 295, 120262. [Google Scholar] [CrossRef]
  28. Tao, X.; Yang, C.; Huang, L.; Xu, D. DBD plasma combined with catalysts derived from NiMgAlCe hydrotalcite for CO2 reforming of CH4. Mater. Chem. Phys. 2020, 250, 123118. [Google Scholar] [CrossRef]
  29. Wang, Y.; Fan, L.; Xu, H.; Du, X.; Xiao, H.; Qian, J.; Zhu, Y.; Tu, X.; Wang, L. Insight into the synthesis of alcohols and acids in plasma-driven conversion of CO2 and CH4 over copper-based catalysts. Appl. Catal. B 2022, 315, 121583. [Google Scholar] [CrossRef]
  30. Dou, L.; Liu, Y.; Gao, Y.; Li, J.; Hu, X.; Zhang, S.; Ostrikov, K.; Shao, T. Disentangling metallic cobalt sites and oxygen vacancy effects in synergistic plasma-catalytic CO2/CH4 conversion into oxygenates. Appl. Catal. B 2022, 318, 121830. [Google Scholar] [CrossRef]
  31. Wang, J.; Zhang, K.; Bogaerts, A.; Meynen, V. 3D porous catalysts for Plasma-Catalytic dry reforming of Methane: How does the pore size affect the Plasma-Catalytic Performance? Chem. Eng. J. 2023, 464, 142574. [Google Scholar] [CrossRef]
  32. Bouchoul, N.; Fourr´e, E.; Duarte, A.; Tanchoux, N.; Louste, C.; Batiot-Dupeyrat, C. Plasma-metal oxides coupling for CH4-CO2 transformation into syngas and/or hydrocarbons, oxygenates. Catal. Today 2021, 369, 62–68. [Google Scholar] [CrossRef]
  33. Li, D.; Rohani, V.; Ramaswamy, A.P.; Sennour, M.; Georgi, F.; Dupont, P.; Fulcheri, L. Orientating the plasma-catalytic conversion of CO2 and CH4 toward liquid products by using a composite catalytic bed. Appl. Catal. 2023, 650, 119015. [Google Scholar] [CrossRef]
  34. Li, J.; Dou, L.; Gao, Y.; Hei, X.; Yu, F.; Shao, T. Revealing the active sites of the structured Ni-based catalysts for one-step CO2/CH4 conversion into oxygenates by plasma-catalysis. J. CO2 Util. 2021, 52, 101675. [Google Scholar] [CrossRef]
  35. Mei, D.; Sun, M.; Li, S.; Zhang, P.; Fang, Z.; Tu, X. Plasma-enabled catalytic dry reforming of CH4 into syngas, hydrocarbons and oxygenates: Insight into the active metals of γ-Al2O3 supported catalysts. J. CO2 Util. 2023, 67, 102307. [Google Scholar] [CrossRef]
  36. Krawczyk, K.; Młotek, M.; Ulejczyk, B.; Schmidt-Szałowski, K. Methane conversion with carbon dioxide in plasma-catalytic system. Fuel 2014, 117, 608–617. [Google Scholar] [CrossRef]
  37. Li, J.; Dou, L.; Liu, Y.; Gao, Y.; Hu, X.; Yu, F.; Li, J.; Zhang, S.; Shao, T. One-step plasma reforming of CO2–CH4 into hydrogen and liquid fuels: The roles of Cu and Fe sites on products distribution. Fuel Process. Technol. 2023, 242, 107648. [Google Scholar] [CrossRef]
  38. Wang, L.; Wang, Y.; Fan, L.; Xu, H.; Liu, B.; Zhang, J.; Zhu, Y.; Tu, X. Direct conversion of CH4 and CO2 to alcohols using plasma catalysis over Cu/Al(OH)3 catalysts. Chem. Eng. J. 2023, 466, 143347. [Google Scholar] [CrossRef]
  39. Wang, S.; Guo, S.; Luo, Y.; Qin, Z.; Chen, Y.; Dong, M.; Li, J.; Fan, W.; Wang, J. Direct synthesis of acetic acid from carbon dioxide and methane over Cu-modulated BEA, MFI, MOR and TON zeolites: A density functional theory study. Catal. Sci. Technol. 2019, 9, 6613. [Google Scholar] [CrossRef]
  40. Montejo-Valencia, B.D.; Pagán-Torres, Y.J.; Martínez-Iñesta, M.M.; Curet-Arana, M.C. Density Functional Theory (DFT) Study to Unravel the Catalytic Properties of M-Exchanged MFI, (M = Be, Co, Cu, Mg, Mn, Zn) for the Conversion of Methane and Carbon Dioxide to Acetic Acid. ACS Catal. 2017, 7, 6719–6728. [Google Scholar] [CrossRef]
  41. Feng, G.; Wen, Z.-H.; Wang, J.; Lu, Z.-H.; Zhou, J.; Zhang, R. Guiding the design of practical MTW zeolite catalysts: An integrated experimental-theoretical perspective. Microporous Mesoporous Mater. 2021, 312, 110810. [Google Scholar] [CrossRef]
  42. Golubev, O.V.; Il’chuk, P.S.; Tsaplin, D.E.; Maximov, A.L. Application of Zeolite-Containing Catalysts For Plasma-Catalytic Carbon Dioxide Methane Reforming. Russ. J. Appl. Chem. 2022, 95, 1749–1755. [Google Scholar] [CrossRef]
  43. Tsaplin, D.E.; Makeeva, D.A.; Kulikov, L.A.; Maksimov, A.L.; Karakhanov, E.A. Synthesis of ZSM-12 zeolites with new templates based on salts of ethanolamines. Russ. J. Appl. Chem. 2018, 91, 1957–1962. [Google Scholar] [CrossRef]
  44. Groen, J.C.; Peffer, L.A.A.; Pérez-Ramírez, J. Pore size determination in modified micro- and mesoporous materials. Pitfalls and limitations in gas adsorption data analysis. Microporous Mesoporous Mater. 2003, 60, 1–17. [Google Scholar] [CrossRef]
  45. Sandoval-Díaz, L.-E.; Gonzalez-Amaya, J.-A.; Trujillo, C.-A. General aspects of zeolite acidity characterization. Microporous Mesoporous Mater. 2015, 215, 229–243. [Google Scholar] [CrossRef]
  46. Wang, H.; Xu, R.; Jin, Y.; Zhang, R. Zeolite structure effects on Cu active center, SCR performance and stability of Cu-zeolite catalysts. Catal. Today 2019, 327, 295–307. [Google Scholar] [CrossRef]
  47. Fedorov, A.V.; Kukushkin, R.G.; Yeletsky, P.M.; Bulavchenko, O.A.; Chesalov, Y.A.; Yakovlev, V.A. Temperature-programmed reduction of model CuO, NiO and mixed CuO-NiO catalysts with hydrogen. J. Alloys Compd. 2020, 844, 156135. [Google Scholar] [CrossRef]
  48. Mierczynski, P.; Mierczynska, A.; Ciesielski, R.; Mosinska, M.; Nowosielska, M.; Czylkowska, A.; Maniukiewicz, W.; Szynkowska, M.I.; Vasilev, K. High Active and Selective Ni/CeO2–Al2O3 and Pd–Ni/CeO2–Al2O3 Catalysts for Oxy-Steam Reforming of Methanol. Catalysts 2018, 8, 380. [Google Scholar] [CrossRef]
  49. The Luong, N.; Okumura, H.; Yamasue, E.; Ishihara, K.N. Structure and catalytic behaviour of CuO–CeO2 prepared by high-energy ball milling. R. Soc. Open Sci. 2019, 6, 181861. [Google Scholar] [CrossRef]
  50. Sun, H.; Wang, H.; Qu, Z. Construction of CuO/CeO2 Catalysts via the Ceria Shape Effect for Selective Catalytic Oxidation of Ammonia. ACS Catal. 2023, 13, 1077–1088. [Google Scholar] [CrossRef]
  51. Luo, M.-F.; Zhong, Y.-J.; Yuan, X.-X.; Zheng, X.-M. TPR and TPD studies of CuO/CeO2 catalysts for low temperature CO oxidation. Appl. Catal. 1997, 162, 121–131. [Google Scholar] [CrossRef]
  52. Zhua, P.; Liu, M.; Zhou, R. Effect of interaction between CuO and CeO2 on the performance of CuO-CeO2 catalysts for selective oxidation of CO in H2-rich streams. Indian J. Chem. 2012, 51A, 1529–1537. [Google Scholar]
  53. Rosenthal, L.A.; Davis, D.A. Electrical Characterization of a Corona Discharge for Surface Treatment. IEEE Trans. Ind. Appl. 1975, IA-11, 328–335. [Google Scholar] [CrossRef]
  54. Golubev, O.V.; Maksimov, A.L. Plasma-Assisted Catalytic Decomposition of Carbon Dioxide. Russ. J. Appl. Chem. 2022, 95, 617–630. [Google Scholar] [CrossRef]
  55. Wang, X.L.; Gao, Y.; Zhang, S.; Sun, H.; Li, J.; Shao, T. Nanosecond pulsed plasma assisted dry reforming of CH4: The effect of plasma operating parameters. Appl. Energy 2019, 243, 132–144. [Google Scholar] [CrossRef]
  56. Munir, S.; Varzeghaniab, A.R.; Kaya, S. Electrocatalytic reduction of CO2 to produce higher alcohols. Sustain. Energy Fuels 2018, 2, 2532–2541. [Google Scholar] [CrossRef]
  57. Andersen, J.A.; Christensen, J.M.; Østberg, M.; Bogaerts, A.; Jensen, A.D. Plasma-catalytic dry reforming of methane: Screening of catalytic materials in a coaxial packed-bed DBD reactor. Chem. Eng. J. 2020, 397, 125519. [Google Scholar] [CrossRef]
  58. Zeng, Y.; Zhu, X.; Mei, D.; Ashford, B.; Tu, X. Plasma-catalytic dry reforming of methane over γ-Al2O3 supportedmetal catalysts. Catal. Today 2015, 256, 80–87. [Google Scholar] [CrossRef]
  59. Ray, D.; Chawdhury, P.; Bhargavi, K.V.S.S.; Thatikonda, S.; Lingaiah, N.; Subrahmanyam, C. Ni and Cu oxide supported γ-Al2O3 packed DBD plasma reactor for CO2 activation. J. CO2 Util. 2021, 44, 101400. [Google Scholar] [CrossRef]
Figure 1. Scheme of the plasma-catalytic unit.
Figure 1. Scheme of the plasma-catalytic unit.
Gases 03 00012 g001
Figure 2. Low-temperature N2 adsorption isotherms of the zeolites, supports, and catalysts.
Figure 2. Low-temperature N2 adsorption isotherms of the zeolites, supports, and catalysts.
Gases 03 00012 g002
Figure 3. Pore size distribution based on desorption branch of the BJH plot.
Figure 3. Pore size distribution based on desorption branch of the BJH plot.
Gases 03 00012 g003
Figure 4. NH3-TPD profiles of the prepared supports and catalysts.
Figure 4. NH3-TPD profiles of the prepared supports and catalysts.
Gases 03 00012 g004
Figure 5. X-ray diffractograms of the supports and obtained catalysts.
Figure 5. X-ray diffractograms of the supports and obtained catalysts.
Gases 03 00012 g005
Figure 6. Temperature programmed reduction profiles of the CuCe-5 and CuCe-12 samples.
Figure 6. Temperature programmed reduction profiles of the CuCe-5 and CuCe-12 samples.
Gases 03 00012 g006
Figure 7. Lissajous figure evolution during the power increase in a blank experiment (empty reactor, CH4:CO2 = 1:1, total gas flow 47 mL/min).
Figure 7. Lissajous figure evolution during the power increase in a blank experiment (empty reactor, CH4:CO2 = 1:1, total gas flow 47 mL/min).
Gases 03 00012 g007
Figure 8. The results of the plasma-catalytic processing of the CH4 and CO2 using the prepared catalysts: (a) the conversion of CH4 and CO2; (b) yield of the gaseous products; (c) selectivity of the gaseous products; (d) energy efficiency and power.
Figure 8. The results of the plasma-catalytic processing of the CH4 and CO2 using the prepared catalysts: (a) the conversion of CH4 and CO2; (b) yield of the gaseous products; (c) selectivity of the gaseous products; (d) energy efficiency and power.
Gases 03 00012 g008
Figure 9. The results of the obtained liquid phase chromatographic analysis.
Figure 9. The results of the obtained liquid phase chromatographic analysis.
Gases 03 00012 g009
Figure 10. The CuCe-12 catalyst stability test during five runs (Absorbed power = 9 W, CH4:CO2 = 1:1, flow rate = 47 mL/min).
Figure 10. The CuCe-12 catalyst stability test during five runs (Absorbed power = 9 W, CH4:CO2 = 1:1, flow rate = 47 mL/min).
Gases 03 00012 g010
Table 1. Reagents used for catalyst preparation.
Table 1. Reagents used for catalyst preparation.
ReagentPurityManufacturer
colloid silica Ludox HS-4040 wt% solutionSigma-Aldrich, St. Louis, MO, USA
Al2(SO4)3·18H2O98%Sigma-Aldrich
N-Methyldiethanolamine99%Sigma-Aldrich
bromoethane99%Sigma-Aldrich
NaOH>97%LLC “Komponent-Reaktiv”, Moscow, Russia
NH4Cl98%Reachem, Chennai, India
boehmite Pural SB99%Sasol, Hamburg, Germany
ZSM-5 zeolite>90%JSC “NZHK”, Novosibirsk, Russia
HNO365 wt% solutionLLC “NevaReaktiv”, Saint-Petersburg, Russia
Ce(NO3)2·6H2O99%LLC “Tsentr Tekhnologii Lantan”, Moscow, Russia
Cu(NO3)2·3H2O99.9%JSC “Lenreaktiv”, Saint-Petersburg, Russia
Table 2. Elemental composition and the textural characteristics of the obtained supports and catalysts.
Table 2. Elemental composition and the textural characteristics of the obtained supports and catalysts.
SampleElemental Composition, wt%Textural Characteristics
AlSiCeCuSBET, m2/gVpores, cm3/gdpores, nm
ZSM-5/Al2O315.732.7--2870.216.1
CuCe-513.030.33.44.62500.216.4
ZSM-12/Al2O31236--1810.199.1
CuCe-1210.632.33.84.51470.1810.2
Table 3. Quantity of the acid sites determined by NH3–TPD.
Table 3. Quantity of the acid sites determined by NH3–TPD.
SampleAcid Site Concentration, μmol/g
Weak Sites
~200 °C
Medium Sites
~250 °C
Strong Sites
~400 °C
Total
ZSM-5/Al2O3118159254531
CuCe-54922345317
ZSM-12/Al2O3232565113
CuCe-122410638168
Table 4. The average crystallite size of the CeO2 and CuO phases.
Table 4. The average crystallite size of the CeO2 and CuO phases.
CatalystPhaseCrystallite Size, nm
CuCe-5CeO24.4 ± 0.5
CuOn/d *
CuCe-12CeO24.6 ± 0.4
CuO18.3 ± 0.5
* n/d = no data.
Table 5. DBD plasma-catalytic DRM and CO2 decomposition performance data using Cu catalysts.
Table 5. DBD plasma-catalytic DRM and CO2 decomposition performance data using Cu catalysts.
CatalystInput Power, WGas Flow, mL/minη (Total), mmol·kJ−1X (CO2),
%
X (CH4),
%
Oxygenates
Distribution
Source
Cu/Al2O31040n/d *716acetic acid, methanol, ethanol, formic acid[21]
Cu/Al(OH)31240n/d617R–OH[29]
Cu/Mg(OH)2919R–OH
Cu/SiO2720R–OH, R–COOH
Cu/HZSM-5919R–OH, R–COOH
Cu/TiO2414R–OH, R–COOH
Cu/CeO2540n/d113methanol, ethanol, Pr-OH, Bu-OH, acetic acid, acetone
Cu/CeO2—lowest oxygenates selectivity
Cu/Al(OH)3—highest oxygenates selectivity
[38]
Cu/TiO245
Cu/γ-Al2O32.511
Cu/Al(OH)3716
Cu/Al2O345500.181632methanol, ethanol, DME, formic acid[57]
Cu/γ-Al2O37.5500.56815n/d[58]
CuCe-58.7470.651220methanol, acetoneThis work
CuCe-120.81130
CuO/CeAl2.2301.413.5n/a **n/a[59]
CuO/Al2O31.615.7
* n/d = no data, ** n/a = not applicable since only CO2 decomposition to CO is considered in the study.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Golubev, O.V.; Tsaplin, D.E.; Maximov, A.L. Greenhouse Gas Conversion into Hydrocarbons and Oxygenates Using Low Temperature Barrier Discharge Plasma Combined with Zeolite Catalysts. Gases 2023, 3, 165-180. https://doi.org/10.3390/gases3040012

AMA Style

Golubev OV, Tsaplin DE, Maximov AL. Greenhouse Gas Conversion into Hydrocarbons and Oxygenates Using Low Temperature Barrier Discharge Plasma Combined with Zeolite Catalysts. Gases. 2023; 3(4):165-180. https://doi.org/10.3390/gases3040012

Chicago/Turabian Style

Golubev, Oleg V., Dmitry E. Tsaplin, and Anton L. Maximov. 2023. "Greenhouse Gas Conversion into Hydrocarbons and Oxygenates Using Low Temperature Barrier Discharge Plasma Combined with Zeolite Catalysts" Gases 3, no. 4: 165-180. https://doi.org/10.3390/gases3040012

Article Metrics

Back to TopTop