Next Article in Journal
Synergistic Effect of Photocatalytic Degradation of Hexabromocyclododecane in Water by UV/TiO2/persulfate
Next Article in Special Issue
Ni-Mo Sulfide Semiconductor Catalyst with High Catalytic Activity for One-Step Conversion of CO2 and H2S to Syngas in Non-Thermal Plasma
Previous Article in Journal
Z-Schemed WO3/rGO/SnIn4S8 Sandwich Nanohybrids for Efficient Visible Light Photocatalytic Water Purification
Previous Article in Special Issue
A Facile Fabrication of Supported Ni/SiO2 Catalysts for Dry Reforming of Methane with Remarkably Enhanced Catalytic Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Combined Magnesia, Ceria and Nickel catalyst supported over γ-Alumina Doped with Titania for Dry Reforming of Methane

by
Ahmed Sadeq Al-Fatesh
1,*,
Samsudeen Olajide Kasim
1,
Ahmed Aidid Ibrahim
1,
Anis Hamza Fakeeha
1,
Ahmed Elhag Abasaeed
1,
Rasheed Alrasheed
2,
Rawan Ashamari
2 and
Abdulaziz Bagabas
2,*
1
Chemical Engineering Department, College of Engineering, King Saud University, P.O. Box 800, Riyadh 11421, Saudi Arabia
2
National Petrochemical Technology Center (NPTC), Materials Science Research Institute (MSRI), King Abdulaziz City for Science and Technology, P.O. Box 6086, Riyadh 11442, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(2), 188; https://doi.org/10.3390/catal9020188
Submission received: 21 January 2019 / Revised: 5 February 2019 / Accepted: 13 February 2019 / Published: 18 February 2019
(This article belongs to the Special Issue Catalysts for Syngas Production)

Abstract

:
This study investigated dry reforming of methane (DRM) over combined catalysts supported on γ-Al2O3 support doped with 3.0 wt. % TiO2. Physicochemical properties of all catalysts were determined by inductively coupled plasma/mass spectrometry (ICP-MS), nitrogen physisorption, X-ray diffraction, temperature programmed reduction/oxidation/desorption/pulse hydrogen chemisorption, thermogravimetric analysis, and scanning electron microscopy. Addition of CeO2 and MgO to Ni strengthened the interaction between the Ni and the support. The catalytic activity results indicate that the addition of CeO2 and MgO to Ni did not reduce carbon deposition, but improved the activity of the catalysts. Temperature programmed oxidation (TPO) revealed the formation of carbon that is mainly amorphous and small amount of graphite. The highest CH4 and CO2 conversion was found for the catalyst composed of 5.0 wt. % NiO-10.0 wt. % CeO2/3.0 wt. %TiO2-γ-Al2O3 (Ti-CAT-II), resulting in H2/CO mole ratio close to unity. The optimum reaction conditions in terms of reactant conversion and H2/CO mole ratio were achieved by varying space velocity and CO2/CH4 mole ratio.
Keywords:
CH4; CeO2; dry reforming; MgO; Ni; TiO2

1. Introduction

Global warming has become an alarming issue. Emissions of greenhouse gases including carbon dioxide (CO2) and methane (CH4) actively contribute to global warming. Methods of transforming these CO2(g) and CH4(g) into useful products are an important area of study to generate industrially important fuels and chemicals [1,2,3]. In this context, numerous reforming reactions of CH4 have been employed using several oxidants (e.g., H2O, CO2, O2, etc.) to produce H2(g) or synthesis gas (syngas, a mixture of H2(g) and CO(g)) with an equimolar ratio of H2(g)/CO(g). Methane reforming processes include steam reforming, auto thermal reforming, tri-reforming, etc. [4,5,6,7,8,9,10,11]. Methane reforming using CO2, known as dry reforming (DRM), is attractive because it mitigates the emission of CH4 and CO2, produces syngas, the starting material in the Fischer-Tropsch process to generate hydrocarbons and oxygenates, and generates clean energy through the combustion of hydrogen [12]. CH4 is a cost-effective feedstock for syngas production. The primary reaction that governs the process is as follows:
CH 4 + CO 2 2 H 2 + 2 CO Δ H 298 = + 247   kJ   mol 1
The reaction is energetically unfavorable, thus requiring high temperatures to achieve acceptable conversion. Both noble metals (i.e., Ru, Rh, or Pt) and first-row transition metals (i.e., Ni, Fe, Co) are common active elements in that catalyze CO2 reforming of CH4. Although noble metals display high activity and stability, their limited availability and high price have rendered them inappropriate for industrial use [13,14]. On the other hand, the first-row transition metals are cheaper and possess similar activity, but their stability is hampered by carbon deposition and particle sintering [15,16,17,18]. Therefore, development of Ni-based catalysts with high activity and resistance to deactivation due to carbon formation and metal sintering is essential for DRM. Catalytic performance can be influenced by many factors such as the active metal, support type, and texture. The support can enhance the catalyst selectivity, activity, and stability by increasing the surface area and dispersion of the active metal [19]. For example, Ni deposited on alumina supports result in high catalytic activity, but rapidly deactivates due to sintering, coke deposition, and formation of surface nickel aluminate phase. To increase the catalytic performance of Ni/γ-Al2O3, various parameters can be incorporated in the catalyst.
Titania (TiO2) is characterized by low specific surface area and poor mechanical strength, and undergoes a phase transformation from anatase to rutile at high temperatures, making it unsuitable for high temperature reactions [20]. Previous studies have shown enhanced thermal stabilization of TiO2 by introducing a thermally stable second metal oxide (i.e., SiO2, Al2O3, etc.) [21,22]. Incorporation of TiO2 in Al2O3 supports can improve metal dispersion, reduce particle sintering, increase thermal stability, and enhance oxygen storage capacity to assists in gasifying carbon produced in reforming reaction [23]. Tauster et al. investigated the effects of support modification on the oxidation state of Ru and the catalytic performance of Ru/TiO2 catalysts under conditions of partial oxidation of methane. It was found that doping of TiO2 with small amounts of WO3 favored oxygen adsorption on Ru under reaction conditions, resulting in a stabilization of a fraction of the catalyst in its oxidized form [24]. Addition of metal oxide promoters has been used to improve Ni metal catalysts. For instance, Shamskar et al. investigated the addition of CeO2, La2O3, and ZrO2 to Ni/Al2O3 catalyst used for DRM and found that ceria-promoted catalyst reduced the carbon formation [25]. Ni-MgO-Al2O3 catalysts were used for steam reforming of methane by Jang et al. [26]. Al-Fatesh et al. studied the promotional effect of ceria in the catalytic DRM and found that the Ni doping with ceria resulted in an excellent activity and lowered coke formation [27]. MgO promoters enhance CH4 conversion and mitigated the effect of the potassium poisoning of the Ni-based catalyst. The MgO promoter is beneficial in suppressing carbon formation.
In the present work, supported combinations of MgO, CeO2 and NiO catalysts were developed to retain high activity and stability while reducing the formation of coke during DRM. The effect of using MgO and CeO2 as separate and combined promoters, for 5.0 wt. % NiO supported over γ-Al2O3 doped with 3.0 wt. % TiO2 was studied. We determine the impact of each modifier on observed catalytic performance.

2. Results and Discussion

2.1. X-ray Powder Diffraction (XRD)

The XRD patterns of all the fresh catalysts are displayed in Figure S1 in the electronic supplementary information (ESI). All the patterns consisted of various metal oxides, where the presence metal oxide phases depended on the added components used to prepare the catalysts. Three metal oxides existed in all catalysts, where these metal oxides were the component of the support: cubic gamma-aluminum oxide, γ-(Al2O3)1.333 (PDF 01-075-0921), cubic synthesized honguiite titanium oxide, (TiO0.8)0.913 (PDF 01-085-1380), and aluminum silicate, Al0.5Si0.75O2.25 (PDF 00-037-1460). Rhombohedral nickel oxide, NiO (PDF 00-044-1159) was found in Ti-CAT-I, Ti-CAT-II, Ti-CAT-III, and Ti-CAT-V (these notations are defined in Section 3.2). When magnesium was added, cubic magnesium nickel oxide, MgNiO2 (PDF 00-024-0712) formed. Cubic synthesized cerianite (Ce) (ceria), CeO2 (PDF 00-034-0394), was detected in Ti-CAT-I, Ti-CAT-II, Ti-CAT-IV, and Ti-CAT-VI. Addition of magnesium strongly influenced the interaction of cerium with the other components of the catalyst. Monoclinic magnesium cerium oxide, MgCeO3 (PDF 00-004-0641), and cubic magnesium cerium titanium oxide, Mg2CeTiO6 (PDF 00-058-0550) were present in Ti-CAT-I and Ti-CAT-IV. Cubic periclase magnesium oxide, MgO (PDF 01-071-1176) was detected in Ti-CAT-I, Ti-CAT-III, and Ti-CAT-IV.

2.2. Inductively Coupled Plasma Mass Spectroscopy (ICP-MS)

ICP-MS analysis was carried out to quantify the metallic components as metal oxides for the best two catalysts. The results are shown in Table 1a,b.
Table 1 summarizes the results of ICP analysis of the metallic components in the prepared catalysts and compares it with the theoretical values. The experimental results were found to be in excellent agreement with the nominal values.

2.3. Temperature Programmed Desorption (CO2-TPD)

The CO2-TPD experiment was performed to study the basicity of the catalysts. The results obtained are shown in Figure 1. The basicity of the catalyst has a paramount influence on the catalytic performance in DRM due to the acidic nature of CO2. Thus, strong basic sites can enhance catalytic activity and increase the chemisorption and reaction of reacting gases [28]. Distribution of basic sites on the catalyst (i.e., weak, intermediate, strong, and very strong) correspond to the different desorption peaks in the temperature ranges of 20–150, 150–300, 300–450, and >450 °C, respectively, in the CO2-TPD profile [29,30].
All catalysts, except Ti-CAT-V and Ti-CAT-VI, showed the same basic site classification, because the CO2 desorption peaks appeared at almost the same different temperature ranges (Figure 1). Both Ti-CAT-V and Ti-CAT-VI have basic sites correspoding to site of high and strong basicity centered at a temperature around 310 °C.
For the peaks appearing at different temperature ranges, peaks in the temperature range of 50–125 °C correspond to weak basic sites, peaks at 160–185 °C fall under the category of intermediate strength basic sites, while the peaks at 260 °C correspond to strong basicity sites. An elbow peak was observed for all of the samples, except for Ti-CAT-V and Ti-CAT-VI, at temperature centered around 500 °C. This peak had no significant CO2 uptake.

2.4. Surface Characterization

The textural properties of the fresh catalysts were studied using nitrogen adsorption-desorption isotherms. The results obtained from the N2 physisorption are shown in Table 2 and that of the isotherms are presented in Figure 2. The results give an insight into the variations in the activities of the catalysts. In accordance to IUPAC classifications of isotherms, the isotherms in Figure 2 fall under the category of type II, with an H3-type hysteresis loop, which results from capillary condensation and evaporation at high relative pressures [31].
Type II isotherms constitute macroporous adsorbents, and for detailed study, the BJH pore size distribution is represented in Figure 2. All the Ti-CAT samples displayed a bimodal mesoporous/macroporous distribution curve with average pore size in the range 11.5–12.5 nm, typical for macroporous adsorbents with large surface area. For example, Jiang et al. [32] and Zhao et al. [33] synthesized macro-mesoporous bimodal titania with high surface areas. Here, the effect of surface area variation is observed when Mg, Ce, and Ni were combined. Table 2 shows that surface areas of the combined metal catalysts are reduced in relation to single-metal component catalysts. This observation is due to the combined metal deposition on the porous structure of the support and filling pores [34].

2.5. H2-TPR

The reduction behavior of the different catalyst samples was investigated using H2-TPR and the profiles are presented in Figure 3. The nickel reduction peaks for Ti-CAT-x (x = I, II, III) samples containing Ni combined with other metals, are characterized by three reduction regions at low, medium and high temperature ranges. Their ranges are dependent on the degree of dispersion and interaction of the active metal with the support. The nickel phase reducibility was influenced by the combination of the metal oxides. The reduction peak in the temperature range of 280–380 °C is assigned to the reduction of NiO having weak interaction with the support. Higher temperature peaks (600–700 °C) are likely due to the reduction of NiO species having strong interactions with the support. The reduction peak of Ni2+ derived from spinel is found at around 810 °C [35].
For Ti-CAT-V (the catalyst with only Ni), the NiO reduction peaks appeared narrower and more intense in temperature ranges lower than those of combined metal counterparts.
Only two reduction peaks are observed for Ti-CAT-VI at temperature ranges centered at 260 and 325 °C. Similar reduction peaks are expected for CeO2 promoted samples, but appear to have merged with the peaks for NiO that appeared around that temperature range.

2.6. Effect of MgO and CeO2 Combination on the Catalytic Performance

The effects of combining CeO2 and MgO on Ti-CAT-V and their catalytic performance were studied by comparing the activities of Ti-CAT-V catalyst with that of Ti-CAT-I, Ti-CAT-II, and Ti-CAT-III. CH4, CO2 conversions, H2/CO mole ratio, and their selectivity at 700 °C, for 7.0 h time-on-stream for DRM were calculated and plotted as shown in Figure 4. All the promoted catalysts have CH4 and CO2 conversions higher than that of the Ti-CAT-V catalyst except for Ti-CAT-VI and Ti-CAT-IV, which showed no sign of reaction during the DRM. Ti-CAT-II had the highest CH4 conversion at the start of the reaction (~55%) and maintained stability at around 52%. The high specific surface area of the catalyst (283 m2/g) enhanced the adsorption, diffusion, and contact of the reactant gases. The high average pore diameter and pore volume of Ti-CAT-II is a likely factor for the best-in-class performance. The Ce and Mg promoted catalysts enhanced the activity. The improvement of the activity is accompanied by the formation of graphitic carbon in comparison with the unpromoted catalysts, as depicted in the TG analysis.
The same trend was observed for CO2 conversion, with the Ti-Cat-V catalyst showing the least conversion. For all the catalysts under investigation, CO2 conversion was observed to be higher than CH4 conversion, which is suggestive of the occurrence of reverse water gas shift (RWGS) reaction. Wang et al. gave the same observation in their study on catalytic hydrogenation of carbon dioxide [36].
H 2 + CO 2 CO + H 2 O                     Δ H 298 = + 41.2   kj / mol
In addition, the H2/CO mole ratio showed values less than 1 for all the catalysts. The deviation from the stoichiometric ratio is also suggestive of the occurrence of RWGS reaction. Ti-CAT-II appeared to be the most selective towards H2 (~48%) and least selective towards CO (~52%), while Ti-CAT-V has the least H2 selectivity (~45%) but the highest CO selectivity (~55%). In all cases, the as-prepared catalysts showed higher CO selectivity than H2.
Ti-CAT-II catalyst resulted in a H2/CO mole ratio value closest to 1, compared to the tested catalysts. The desirable value of the syngas ratio suitable for downstream Fischer-Tropsch synthesis is unity [37], thus making it the best option for the dry reforming.

2.7. H2-Pulse Chemisorption

To understand the effect of the active Ni component on the catalytic performance of the best two catalysts, Ti-CAT-I and Ti-CAT-II, we carried out H2 pulse chemisorption to determine the degree of Ni dispersion on the surface of the support and Ni metallic surface area. The results of H2-pulse chemisorption are displayed in Table 3. We found that both catalysts had high Ni metallic surface areas of ~90% and good dispersion of ~13%, which is responsible for the good catalytic performance of the two catalysts. The small relative higher catalytic performance of Ti-CAT-II than that of Ti-CAT-I could be attributed to the slightly higher Ni metallic surface area and dispersion of Ti-CAT-II.

2.8. Temperature Programmed Oxidation (TPO) of the Spent Catalysts

TPO is a useful technique that can be employed to determine the nature of the carbon deposited onto the surface of the catalysts. Several forms of carbon deposition have been reported in dry reforming reactions—ranging from atomic carbon, to graphitic, and amorphous carbon. The carbon can undergo gasification to form CO2 under oxidative atmosphere and at different temperature ranges. The atomic carbon, amorphous, and graphitic carbon can be gasified at temperatures less than 250, 250–600, and >600 °C, respectively [38]. The TPO profiles of Ti-CAT-I and Ti-CAT-II spent catalysts are shown in Figure 5. Each of the catalysts exhibited a broad peak near 600 °C and a low-intensity shoulder at 100–250 °C. According to the TPO results, the carbon deposited on both Ti-CAT-I and Ti-CAT-II spent catalysts revealed the formation of carbon atoms, mostly amorphous carbon, and a small amount of extent graphitic carbon.

2.9. SEM and TG Analysis

We used scanning electron microscopy (SEM) to determine the change in morphology of the spent catalysts. Figure 6 shows the SEM micrographs for the best two catalysts: Ti-CAT-I and Ti-CAT-II. Similar morphology, based on agglomerated, spherical nanoparticles, was observed for both fresh catalysts (Figure 6A,B). Such an observation was expected, because both catalysts were synthesized using an identical preparation procedure and had similar components.
The morphology of the spent catalysts was similar to that one of the fresh samples, except for the presence of carbon nanotubes (CNTs) on the surface of the spent catalysts (Figure 6C,D). Detection of CNTs on the surface of the spent catalyst is in agreement with the TPO results (c.f. Figure 5) and confirms the results of TGA of spent catalysts (Figure 7). The presence of CNTs on the surface of the spent catalysts could be attributed to Boudouard reaction, which in turn would be responsible for reducing the catalytic performance.
After the 7 h reaction, we analyzed the used catalysts by thermal gravemetric analysis (TGA), a quantitative analysis that determines the amount of carbon deposition. Figure 7 shows the result of the analysis. Catalysts that showed no sign of reaction were not reported. The weight loss (%) for virtually all the catalysts began at around 620 °C. The TGA profiles revealed that both Ti-CAT-V and Ti-CAT-III catalysts had the lowest weight loss, ~15.0%, while the two most reactive catalysts (Ti-CAT-II and Ti-CAT-I) had the highest amount of carbon deposition, corresponding to a weight loss of 25.0%.
From these TGA results of spent catalysts, it can be inferred that the combined metal catalysts, namely Ti-CAT-II and Ti-CAT-I, enhance the feed conversion capacity of the catalysts and gasify the carbon deposited over the surface to a considerable extent.

2.10. Effect of Space Velocity

The effect of gas hourly space velocity (GHSV) was studied on the catalyst that showed the best performance in the previous section (i.e., Ti-CAT-II catalyst). GHSV of 19,500 and 78,000 feed   flow   rate   mass   of   cat .     ( mL g ·   h ) were considered at 700 °C and time-on-stream over 7.0 h for DRM, while keeping the mass of the catalyst constant. These GHSV values are half and twice as much as the initial GHSV of 39,000 mL g−1 h−1, respectively. The results, in terms of CO2 and CH4 conversions, as well as H2/CO mole ratio, were calculated and plotted in Figure 8A,B. As the GHSV increased, the CH4 and CO2 conversions decreased, with the highest conversions for both CH4 and CO2 being obtained at a GHSV of 19,500   feed   flow   rate   mass   of   cat .     ( mL g ·   h ) . The decrease in conversions can be attributed to the feed having less residence time at higher GHSV [39]. A similar trend was observed with H2/CO mole ratio, where it decreased from a ratio of 1 to ~0.8. However, the results at GHSV of 39,000 were the most stable in comparison to those obtained at other GHSV values.

2.11. Effect of GHSV on Carbon Deposition

Quantitative analysis of carbon deposition was performed on the catalyst Ti-CAT-II used in methane dry reforming at 3 different space velocities 19,500, 39,000 and 78,000 mL g−1 h−1.
The results obtained after the completion of the reactions are shown in Figure 9. The analysis for the reaction performed at 19,500 mL g−1 h−1 showed the least amount of carbon deposition of about 18%, which shows that, relatively, more contact time between the catalyst and the feed stream was allowed at this space velocity, giving room for gasification of the coke that was deposited during the reaction. The reactions carried out at 39,000 and 78,000 mL g−1 h−1 showed higher carbon deposition of about 26 and 25%, respectively. This is an indication that at the higher space velocities, the residence time was not enough for the gasification of the carbon deposit, which variably continued to pile up. Lalit et al. reported similar findings in their study of the effect of GHSV on the conversion of CH4 and CO2 [39].
From the results of the investigation, it can be inferred that deactivation is expected at all the space velocity investigated, since carbon deposit was evident; however, at 19,500 mL/(g·h) space velocity, the catalyst will stay active for a longer time than at 39,000 and 78,000 mL/(g·h).

2.12. Effect of Different CO2/CH4 Ratios

The mole ratio of CO2 to CH4 was changed at a fixed total flow rate to study the performance of Ti-CAT-II catalyst when CH4 was supposed to act as the limiting reagent in excess of CO2 at 700 °C and 39,000 mL g−1 h−1 GHSV. The results are shown in Figure 10a–c. The highest CH4 conversion of about 78% was obtained when CO2 was 20% in excess of CH4, while the least conversion of CH4 (~43%) resulted when the amount of CO2 was 50% of the required stoichiometric amount in the feed. This observation was expected, as CH4 would have enough CO2 to undergo dry reforming. On the other hand, the highest CO2 conversion of (~90%) was observed when CO2 was the limiting reagent. This observation could be due to excess CH4 present in the feed. The CO2 conversion was reduced with the reaction time-on-stream. Such observation could be ascribed to the disproportionation of carbon monoxide into CO2 and graphite, a transformation known as the Boudouard reaction:
2CO(g) = CO2(g) + C(s)
Comparing the different CO2/CH4 ratios, it was observed that CH4 conversion increased with the ratio up to 1.2 (i.e., 0.5 < 1.0 < 1.2) and then declined slightly at 1.5. However, the conversion for CO2 was observed to decrease as the ratio increased (i.e., 1.5 < 1.2 < 1.0 < 0.5).
Figure 10c displays the H2/CO mole ratio results. It was observed that at the lowest CO2/CH4 mole ratio, the H2/CO mole ratio was greater than one. This observation could be owing to the insufficient amount of CO2 for complete dry reforming of the available CH4 and to the thermal decomposition of unreformed CH4, giving more H2 than the stoichiometric amount. Moreover, the Boudouard reaction might contribute to the increase of hydrogen production, because the formed CO2 from Boudouard reaction would shift the DRM equilibrium to the product side.
On the other hand, H2/CO mole ratio was close to one of the cases where CO2 was in excess of CH4, where it was noticed that the H2/CO mole ratio increased with the reaction time-on-stream. Once again, the Boudouard reaction might be responsible for such observation.

3. Experimental Section

3.1. Materials

Nickel nitrate hexahydrate [Ni(NO3)2.6H2O, 98%, Alfa Aesar], cerium nitrate hexahydrate [Ce(NO3)3.6H2O, 99.0% assay on Ce basis, general purpose reagent, BDH], magnesium acetate tetra-hydrate [Mg(O2CCH3)2.4H2O, 99.5-102.0%, Merck & Co., Inc., Kenilworth, NJ, USA] were commercially available and were used without further purification. γ-Alumina doped with titania (3.0 wt. % TiO2/γ-Al2O3) in the shape of pellets, was a gift from Tiancun Xiao, Senior Research Fellow, Inorganic Chemistry Laboratory, Oxford University. Ultrapure deionized water (18.2 MΩ.cm) was obtained from a Milli-Q water purification system (Millipore, Burlington, MA, USA).

3.2. Catalyst Preparation

The required amounts of Ni(NO3)2.6H2O, Ce (NO3)3.6H2O, Mg (O2CCH3)2.4H2O, and support were mixed and were ground together to fine powder by pestle and mortar. A small amount of ultrapure water was used to convert the solid mixture into a paste, which was spun mechanically until dryness. The paste and spinning process was repeated three times. The final solid was calcined in a digital, programmed muffle furnace at 600 °C for three hours by ramping temperature from room temperature by a rate of 3.0 °C/min. The notation of the prepared catalyst samples and their wt. % loadings of nickel oxide, ceria, and magnesia at 600 °C calcination are given below in Table 4.

3.3. Catalyst Characterization

The metallic component composition of all catalysts was determined by an Agilent 7800 inductively coupled plasma mass spectrometry (ICP) at the laboratory of IDAC Merieux NutriSciences, Riyadh, Saudi Arabia. Carbon deposition on the used catalysts was measured by thermogravimetric analysis (TGA) under air by using a Shimadzu TGA-51(Shimadzu Corp., Kyoto, Japan). A certain amount from the spent catalyst (10 mg) was subjected to heat treatment within the temperature range 25 °C–1000 °C. Ramping temperature was maintained at 20 °C/min. Temperature programmed oxidation (TPO) was performed in an oxidative atmosphere to determine the kind of carbon deposited over the surface of the catalyst using Micromeritics AutoChem II over a temperature range of 50–800 °C under a flow of 10% O2/He mixture at 40 mL/min. The spent catalyst was first pretreated in the presence of high purity Argon at 150 ℃ for 30 min and subsequently cooled to room temperature. The Brunauer-Emmet-Teller technique was adopted in calculating the surface area per unit mass of the samples using a device that analyses surface area and porosity, i.e., Micromeritics Tristar II 3020 (Micromeritics Instrument Corporation, Norcross, GA, USA). For nitrogen physisorption measurements, an amount of 0.20–0.30 g weighed from the catalyst was subjected to degassing at 300 °C for three hours prior to analysis. The reducibility of the fresh catalysts was determined by the Micromeritics AutoChem II (Micromeritics Instrument Corporation, Norcross, GA, USA). A sample weight of 75.0 mg was analyzed. Samples were first heated under argon (99.9%) at 150 °C for 30 min, thereafter cooled to 25 °C. Afterwards, samples were heated to 1000 °C at 10 °C/min by allowing the flow of 10% H2/Ar gas at 40 mL/min. A thermal conductivity detector (TCD) was used to follow the H2 consumption. Temperature programmed desorption of carbon dioxide (CO2-TPD) and CO pulse chemisorption measurements were obtained from an automatic chemisorption equipment (Micromeritics AutoChem II 2920) with a TCD. At the start, a 70 mg sample was heated at 200 °C for 1 h under helium (He) flow to remove adsorbed components. Then, CO2 adsorption was carried out at 50 °C for 60 min in the flow of He/CO2 gas mixture (90/10 L/L) with a flow rate of 30.0 mL/min. Afterwards, a linear temperature rise at a rate of 10 °C/min until 800 °C was registered by the TCD of CO2 desorption signal. The nickel metallic surface area and dispersion were determined by H2 pulse chemisorption by using Micromeritics AutoChem II. A sample of 50.0 mg was heated to 150 °C under vacuum for sixteen hours. The sample was then transferred to the sample tube and was heated at temperature rate of 10.0 °C/min to 400 °C under flow rate of 10.0 mL/min of 10%H2/Ar for one hour. The sample was then flushed with highly pure Ar for one hour at 400 °C. The temperature was then reduced to 70.0 °C and pulses of H2 gas were introduced for one hour for determining the H2 uptake. X-ray powder diffraction patterns for the samples were recorded on a Bruker D8 Advance (Bruker, Billerica, MA, USA) XRD diffractometer by using Cu Kα radiation source and a nickel filter, operated at 40 kV and 40 mA. The step size and scanning range of 2θ for analysis was set to 0.01° and 5–100°, respectively. The present phases were documented using standard powder XRD cards (JCPDS). Catalyst morphology was studied using JEOL JSM-7100F (JEOL, Tokyo, Japan) (field emission scanning electron microscope, equipped with energy-dispersive X-ray spectroscopy (EDXS) for surface elemental analysis.

3.4. Catalytic Perfromance

Methane reforming reaction was accomplished in a fixed-bed tubular stainless-steel micro-reactor (ID = 9 mm) at atmospheric pressure. The reactor system was provided by Process Integral Development (Process Integral Development Eng & Tech SL, Madrid, Spain). Before performing the DRM reaction, a 0.10 g catalyst was activated by H2 flow of 40 mL/min at 700 °C for 60 min. N2 gas was then admitted to the reactor for 20 min to remove adsorbed H2 while the catalyst was kept at reaction temperature (700 °C). Afterwards, feed gases of CH4, CO2, and N2 were injected at flow rates of 30, 30 and 5 mL/min, respectively. The temperature, pressure and reaction variables were inspected through the reactor panel. A GC (Shimadzu Corp., Kyoto, Japan) unit having a thermal conductivity detector and two columns, Porapak Q and Molecular Sieve 5A, was connected in series/bypass connections in order to have a complete analysis of the reaction products. The following equations were used to calculate the CH4 and CO2 conversions respectively.
% CH 4   conversion = CH 4   in CH 4   out CH 4   in × 100
% CO 2   conversion = CO 2   in CO 2   out CO 2   in × 100

4. Conclusions

This paper investigated the dry reforming of methane, CH4, over Ti-CAT-V catalyst, and the effects of promoters such as CeO2 and MgO, on the catalytic activity and stability of the catalyst. The promoter loading was 10.0 wt. % and 1.0 wt. % for CeO2 and MgO, respectively. Promoted Ti-CAT-V catalyst showed better conversion of both CH4 and CO2 than the un-promoted counterpart. Ti-CAT-II had the highest CH4 and CO2 conversion of about 55% and 64% respectively, while no reaction was observed for Ti-CAT-VI and Ti-CAT-IV. It can be inferred from the improved performance of the promoted catalysts that the promoters had a positive influence on the textural properties, metal support interaction and reduction behavior of the catalyst. These impacts of promoters were well shown by the characterization techniques used. From the thermogravimetric analysis, un-promoted catalyst gave the lowest carbon deposition. The promoted catalyst, especially by Ce, with higher amounts than 10% was found to have the highest carbon formation. This result implied that the promoters enhanced the activity performance of the catalyst, resulting in the formation of graphitic carbon, and hence, were not effective in boosting the stability via reduction of carbon deposition relative to the un-promoted catalyst. The TPO investigation indicated the types of carbon formed, which include atomic, amorphous, and graphitic carbon.
Ti-CAT-V was selected for further investigation at different GHSVs and subsequently at various CO2/CH4 ratios. An inverse relationship between GHSV and catalytic activity was observed. A GHSV of 19,500   feed   flow   rate   mass   of   cat .     ( mL g ·   h ) and CO2/CH4 ratio of 0.5 gave the best results.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/2/188/s1, Figure S1: XRD patterns of fresh catalysts.

Author Contributions

A.S.F., A.H.F, S.O.K., R.A. (Rasheed Alrasheed), R.A. (Rawan Ashamari) and A.B. carried out all experiments and characterization tests as well as shared in the analysis of the data and shared in the writing of the manuscript. A.S.F, S.O.K, A.E.A and A.B. wrote the paper and shared data analysis. A.H.F and A.A.I. contributed in writing the paper and edited it.

Funding

This research was funded by the Deanship of Scientific Research at King Saud University, Project No. RGP-119.

Acknowledgments

The authors would like to express their sincere appreciation to the Deanship of Scientific Research at King Saud University for its funding for this research group project No. (RGP-119).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Albarazi, A.; Beaunier, P.; Da Costa, P. Hydrogen and syngas production by methane dry reforming on SBA-15 supported nickel catalysts: On the effect of promotion by Ce0.75Zr0.25O2 mixed oxide. Int. J. Hydrogen Energy 2013, 38, 127–139. [Google Scholar] [CrossRef]
  2. Bao, Z.; Lu, Y.; Han, J.; Li, Y.; Yu, F. Highly active and stable Ni based bimodal pore catalysts for dry reforming of methane. Appl. Catal. A Gen. 2015, 491, 116–126. [Google Scholar] [CrossRef]
  3. Buelens, L.C.; Galvita, V.V.; Poelman, H.; Detavernier, C.; Marin, G.B. Super-dry reforming of methane intensifies CO2 utilization via Le Chatelier’s principle. Science 2016, 354, 449–452. [Google Scholar] [CrossRef] [PubMed]
  4. Julián-Durán, L.M.; Ortiz-Espinoza, A.P.; El-Halwagi, M.M.; Jiménez-Gutiérrez, A. Techno-economic assessment and environmental impact of shale gas alternatives to methanol. ACS Sustain. Chem. Eng. 2014, 2, 2338–2344. [Google Scholar]
  5. Schulz, H. Short history and present trends of Fischer-Tropsch synthesis. Appl. Catal. A Gen. 1999, 186, 3–12. [Google Scholar] [CrossRef]
  6. Sheu, E.J.; Mokheimer, E.M.A.; Ghoniem, A.F. A review of solar methane reforming systems. Int. J. Hydrogen Energy 2015, 40, 12929–12955. [Google Scholar] [CrossRef] [Green Version]
  7. Barelli, L.; Bidini, G.; Gallorini, F.; Servili, S. Hydrogen production through sorption-enhanced steam methane reforming and membrane technology. Energy 2008, 33, 554–570. [Google Scholar] [CrossRef]
  8. Hyunjin, J.; Junghun, L.; Eunyeong, C.; Ilsung, S. Hydrogen production from steam reforming using an indirect heating method. Int. J. Hydrogen Energy 2018, 43, 3655–3663. [Google Scholar]
  9. Cyril, P.; Wenhao, F.; Mickaël, C.; Sébastien, P.; Louise, J.D. Steam reforming, partial oxidation and oxidative steam reforming for hydrogen production from ethanol over cerium nickel based oxyhydride catalyst. Appl. Catal. A Gen. 2016, 518, 78–86. [Google Scholar]
  10. Yousri, M.A.W.; Mohamed, M.E.G.; Nader, R.A. Steam and partial oxidation reforming options for hydrogen production from fossil fuels for PEM fuel cells. AEJ 2012, 51, 69–75. [Google Scholar] [Green Version]
  11. Devin, M.W.; Sandra, L.P.; John, T.W.; John, N.K. Synthesis gas production to desired hydrogen to carbon monoxide ratios by tri-reforming of methane using Ni–MgO–(Ce,Zr)O2 catalysts. Appl. Catal. A Gen. 2012, 445–446, 61–68. [Google Scholar]
  12. Linus, A.S.; Lea, C.S.K.; Karla, H.D.; Stephan, A.S.; Johannes, A.L. On the coke deposition in dry reforming of methane at elevated pressures. Appl. Catal. A Gen. 2015, 504, 599–607. [Google Scholar]
  13. Drif, A.; Bion, N.; Brahmi, R.; Ojala, S.; Pirault-Roy, L.; Turpeinen, E.; Seelam, P.K.; Keiski, R.L.; Epron, F. Study of the dry reforming of methane and ethanol using Rh catalysts supported on doped alumina. Appl. Catal. A Gen. 2015, 504, 576–584. [Google Scholar] [CrossRef]
  14. Jabbour, K.; El Hassan, N.; Casale, S.; Estephane, J.; El Zakhem, H. Promotional effect of Ru on the activity and stability of Co/SBA-15 catalysts in dry reforming of methane. Int. J. Hydrogen Energy 2014, 39, 7780–7787. [Google Scholar] [CrossRef]
  15. Du, X.; France, L.J.; Kuznetsov, V.L.; Xiao, T.; Edwards, P.P.; AlMegren, H.; Bagabas, A. Dry Reforming of Methane Over ZrO2-Supported Co–Mo Carbide Catalyst. Appl. Petrochem. Res. 2014, 4, 137–144. [Google Scholar] [CrossRef]
  16. France, L.J.; Du, X.; Almuqati, N.; Kuznetzov, V.L.; Zhao, Y.; Zheng, J.; Xiao, T.; Bagabas, A.; Almegren, H.; Edwards, P.P. The Effect of Lanthanum Addition on The Catalytic Activity of γ-Alumina Supported Bimetallic Co–Mo Carbides for Dry Methane Reforming. Appl. Petrochem. Res. 2014, 4, 145–156. [Google Scholar] [CrossRef]
  17. Tao, X.; Wang, G.; Huang, L.; Li, X.; Ye, Q. Effect of Cu-Mo Activities on the NiCuMo/Al2O3 Catalyst for CO2 Reforming of methane. Catal. Lett. 2016, 146, 2129–2138. [Google Scholar] [CrossRef]
  18. Budiman, A.W.; Song, S.-H.; Chang, T.-S.; Shin, C.-H.; Choi, M.-J. Dry reforming of methane over cobalt catalysts: A literature review of catalyst development. Catal. Surv. Asia 2012, 16, 183–197. [Google Scholar] [CrossRef]
  19. Park, J.H.; Lee, D.; Lee, H.C.; Park, E.D. Steam reforming of liquid petroleum gas over Mn-promoted Ni/γ-Al2O3 catalysts. Korean J. Chem. Eng. 2010, 27, 1132–1138. [Google Scholar] [CrossRef]
  20. Gao, X.; Bare, S.R.; Fierro, J.L.G.; Banares, M.A.; Wachs, I.E. Preparation and in-Situ Spectroscopic Characterization of Molecularly Dispersed Titanium Oxide on Silica. J. Phys. Chem. B 1998, 102, 5653–5666. [Google Scholar] [CrossRef]
  21. Klein, S.; Thorimbert, S.; Maier, W.F. Amorphous microporous titania–silica mixed oxides: Preparation, characterization, and catalytic redox properties. J. Catal. 1996, 163, 476–488. [Google Scholar] [CrossRef]
  22. Miller, J.B.; Ko, E.I. Control of mixed oxide textural and acidic properties by the sol-gel method. Catal. Today 1997, 35, 269–292. [Google Scholar] [CrossRef]
  23. Niu, F.; Li, S.; Zong, Y.; Yao, Q. Catalytic Behavior of Flame-Made Pd/TiO2 Nanoparticles in Methane Oxidation at Low Temperatures. J. Phys. Chem. C 2014, 118, 19165–19171. [Google Scholar] [CrossRef]
  24. Tauster, S.J.; Fung, S.C.; Baker, R.T.K.; Horsely, J.A. Partial Oxidation of Methane to Synthesis Gas over Ru/TiO2 Catalysts: Effects of Modification of the Support on Oxidation State and Catalytic Performance. Science 1981, 211, 1121–1125. [Google Scholar] [CrossRef]
  25. Shamskar, F.R.; Meshkani, F.; Rezaei, M. Preparation and characterization of ultrasound-assisted co-precipitated nanocrystalline La-, Ce-, Zr–promoted Ni-Al2O3 catalysts for dry reforming reaction. J. CO2 Util. 2017, 22, 124–134. [Google Scholar] [CrossRef]
  26. Jang, W.J.; Jung, Y.S.; Shim, J.O.; Roh, H.S.; Yoon, W.L. Preparation of a Ni-MgO-Al2O3 catalyst with high activity and resistance to potassium poisoning during direct internal reforming of methane in molten carbonate fuel cells. J. Power Sources 2018, 378, 597–602. [Google Scholar] [CrossRef]
  27. Ahmed, S.A.A.; Anis, H.F.; Ahmed, E.A. Effects of Selected Promoters on Ni/Y-Al2O3 Catalyst Performance in Methane Dry Reforming. Chin. J. Catal. 2011, 32, 1604–1609. [Google Scholar]
  28. Zhu, F.; Zhang, H.; Yan, X.; Yan, J.; Ni, M.; Li, X.; Tu, X. Plasma-catalytic reforming of CO2-rich biogas over Ni/γ-Al2O3 catalysts in a rotating gliding arc reactor. Fuel 2017, 199, 430–437. [Google Scholar] [CrossRef]
  29. Mei, D.; Ashford, B.; He, Y.-L.; Tu, X. Plasma-catalytic reforming of biogas over supported Ni catalysts in a dielectric barrier discharge reactor: Effect of catalyst supports. Plasma Process. Polym. 2017, 14, 1600076. [Google Scholar] [CrossRef]
  30. Akbari, E.; Alavi, S.M.; Rezaei, M. Synthesis gas production over highly active and stable nanostructured NiMgOAl2O3 catalysts in dry reforming of methane: Effects of Ni contents. Fuel 2017, 194, 171–179. [Google Scholar] [CrossRef]
  31. Sudarsanam, P.; Hillary, B.; Deepa, D.K.; Amin, M.H.; Mallesham, B.; Reddy, B.M.; Bhargava, S.K. Highly efficient cerium dioxide nanocube-based catalysts for low temperature diesel soot oxidation: The cooperative effect of cerium- and cobalt-oxides. Catal. Sci. Technol. 2015, 5, 3496–3500. [Google Scholar] [CrossRef]
  32. Li, R.; Qian, R.; Jiang, Z.-T. Synthesis of high specific surface area titania monolith by addition of poly (ethylene oxide). Mater. Res. Innov. 2015, 19, S8–S136. [Google Scholar] [CrossRef]
  33. Liu, Y.; Lan, K.; Bagabas, A.A.; Zhang, P.; Gao, W.; Wang, J.; Sun, Z.; Fan, J.; Elzatahry, A.A.; Zhao, D. Ordered Macro/Mesoporous TiO2 Hollow Microspheres with Highly Crystalline Thin Shells for High-Efficiency Photoconversion. Small 2016, 12, 860–867. [Google Scholar] [CrossRef] [PubMed]
  34. Pudukudy, M.; Yaakob, Z.; Akmal, Z.S. Direct decomposition of methane over SBA-15 supported Ni, Co and Fe based bimetallic catalysts. Appl. Surf. Sci. 2015, 330, 418–430. [Google Scholar] [CrossRef]
  35. Siew, K.W.; Lee, H.C.; Gimbun, J.; Chin, S.Y.; Khan, M.R.; Taufiq-Yap, Y.H.; Cheng, C.K. Syngas production from glycerol-dry (CO2) reforming over La-promoted Ni/Al2O3 catalyst. Renew. Energy 2015, 74, 441–447. [Google Scholar] [CrossRef]
  36. Wang, W.; Wang, S.P.; Ma, X.B.; Gong, J.L. Recent advances in catalytic hydrogenation of carbon dioxide. Chem. Soc. Rev. 2011, 40, 3703–3727. [Google Scholar] [CrossRef] [PubMed]
  37. Wu, J.; Fang, Y.; Wang, Y.; Zhang, D.K. Combined coal gasification and methane reforming for production of syngas in a fluidized-bed reactor. Energy Fuel 2005, 19, 512–516. [Google Scholar] [CrossRef]
  38. Hao, Z.; Zhu, Q.; Jiang, Z.; Hou, B.; Li, H. Characterization of aerogel Ni/Al2O3 catalysts and investigation on their stability for CH4-CO2 reforming in a fluidized bed. Fuel Process. Technol. 2009, 90, 113–121. [Google Scholar] [CrossRef]
  39. Lalit, S.G.; Guido, S.J.S.; Valero-Romero, M.J.; Mallada, R.; Santamaria, J.; Stankiewicz, A.I.; Stefanidis, G.D. Synthesis, characterization, and application of ruthenium-doped SrTiO3 perovskite catalysts for microwave-assisted methane dry reforming. Chem. Eng. Process. Process Intensif. 2018, 127, 178–190. [Google Scholar]
Figure 1. CO2-TPD profiles of the synthesized catalysts.
Figure 1. CO2-TPD profiles of the synthesized catalysts.
Catalysts 09 00188 g001
Figure 2. N2 adsorption-desorption isotherms and BJH desorption pore size distribution curves for Ti-CAT samples.
Figure 2. N2 adsorption-desorption isotherms and BJH desorption pore size distribution curves for Ti-CAT samples.
Catalysts 09 00188 g002aCatalysts 09 00188 g002b
Figure 3. TPR profiles of the promoted and un-promoted catalysts.
Figure 3. TPR profiles of the promoted and un-promoted catalysts.
Catalysts 09 00188 g003
Figure 4. Catalytic performance of Ti-CAT-I, Ti-CAT-II, Ti-CAT-III, and Ti-CAT-V (a) CH4 conversion (b) CO2 conversion and (c) H2/CO ratio and (d) H2, CO selectivity.
Figure 4. Catalytic performance of Ti-CAT-I, Ti-CAT-II, Ti-CAT-III, and Ti-CAT-V (a) CH4 conversion (b) CO2 conversion and (c) H2/CO ratio and (d) H2, CO selectivity.
Catalysts 09 00188 g004
Figure 5. TPO profiles for both Ti-CAT-I and Ti-CAT-II.
Figure 5. TPO profiles for both Ti-CAT-I and Ti-CAT-II.
Catalysts 09 00188 g005
Figure 6. SEM micrographs for fresh catalysts (A) Ti-CAT-I, (B) Ti-CAT-II, and spent catalysts (C) Ti-CAT-I, (D) Ti-CAT-II. White circles are for some areas where CNTs are present.
Figure 6. SEM micrographs for fresh catalysts (A) Ti-CAT-I, (B) Ti-CAT-II, and spent catalysts (C) Ti-CAT-I, (D) Ti-CAT-II. White circles are for some areas where CNTs are present.
Catalysts 09 00188 g006
Figure 7. TGA profiles for the spent catalysts.
Figure 7. TGA profiles for the spent catalysts.
Catalysts 09 00188 g007
Figure 8. (A) CO2 conversion for Ti-CAT-II at different gas hourly space velocity (B) CH4 and H2/CO ratio for Ti-CAT-II catalyst at different space velocities.
Figure 8. (A) CO2 conversion for Ti-CAT-II at different gas hourly space velocity (B) CH4 and H2/CO ratio for Ti-CAT-II catalyst at different space velocities.
Catalysts 09 00188 g008
Figure 9. TGA curves for Ti-Cat-II at 19,500, 39,000 and 78,000 mL/(g·h) GHSV values.
Figure 9. TGA curves for Ti-Cat-II at 19,500, 39,000 and 78,000 mL/(g·h) GHSV values.
Catalysts 09 00188 g009
Figure 10. (a) CH4 (b) CO2 conversions, and (c) H2/CO ratio for different CO2/CH4 ratio over Ti-CAT-II.
Figure 10. (a) CH4 (b) CO2 conversions, and (c) H2/CO ratio for different CO2/CH4 ratio over Ti-CAT-II.
Catalysts 09 00188 g010
Table 1. ICP metal oxide microanalysis of Ti-CAT.
Table 1. ICP metal oxide microanalysis of Ti-CAT.
(a)
CatalystTi-CAT-I
ComponentNiOCeO2MgOTiO2SiO2Al2O3
Theoretical, wt/wt. %5.0010.001.003.002.0079.00
Experimental, wt/wt. %5.219.911.022.861.9378.05
(b)
CatalystTi-CAT-II
ComponentNiOCeO2MgOTiO2SiO2Al2O3
Theoretical, wt/wt. %5.0010.000.003.002.0079.00
Experimental, wt/wt. %4.9810.110.002.922.0781.03
Table 2. N2 physisorption results for the different catalysts.
Table 2. N2 physisorption results for the different catalysts.
CatalystBET Surface Area (m2/g)Av. Pore Diameter (nm)Pore Volume (cm3/g)
Ti-CAT-I28411.50.40
Ti-CAT-II28312.40.43
Ti-CAT-III32611.80.43
Ti-CAT-IV25612.30.39
Ti-CAT-V33412.40.43
Ti-CAT-VI29912.50.40
Table 3. Ni metallic surface area and dispersion obtained by H2 chemisorption.
Table 3. Ni metallic surface area and dispersion obtained by H2 chemisorption.
CatalystNi Metallic Surface Area, m2/gNi Dispersion, %
Ti-CAT-I8913.3
Ti-CAT-II9113.6
Table 4. Prepared catalyst samples and the wt. % of their composition.
Table 4. Prepared catalyst samples and the wt. % of their composition.
CatalystConcentration, wt. %
NiOCeO2MgO
Ti-CAT-I5.010.01.0
Ti-CAT-II5.010.00.0
Ti-CAT-III5.00.01.0
Ti-CAT-IV0.010.01.0
Ti-CAT-V5.00.00.0
Ti-CAT-VI0.010.00.0

Share and Cite

MDPI and ACS Style

Sadeq Al-Fatesh, A.; Olajide Kasim, S.; Aidid Ibrahim, A.; Hamza Fakeeha, A.; Elhag Abasaeed, A.; Alrasheed, R.; Ashamari, R.; Bagabas, A. Combined Magnesia, Ceria and Nickel catalyst supported over γ-Alumina Doped with Titania for Dry Reforming of Methane. Catalysts 2019, 9, 188. https://doi.org/10.3390/catal9020188

AMA Style

Sadeq Al-Fatesh A, Olajide Kasim S, Aidid Ibrahim A, Hamza Fakeeha A, Elhag Abasaeed A, Alrasheed R, Ashamari R, Bagabas A. Combined Magnesia, Ceria and Nickel catalyst supported over γ-Alumina Doped with Titania for Dry Reforming of Methane. Catalysts. 2019; 9(2):188. https://doi.org/10.3390/catal9020188

Chicago/Turabian Style

Sadeq Al-Fatesh, Ahmed, Samsudeen Olajide Kasim, Ahmed Aidid Ibrahim, Anis Hamza Fakeeha, Ahmed Elhag Abasaeed, Rasheed Alrasheed, Rawan Ashamari, and Abdulaziz Bagabas. 2019. "Combined Magnesia, Ceria and Nickel catalyst supported over γ-Alumina Doped with Titania for Dry Reforming of Methane" Catalysts 9, no. 2: 188. https://doi.org/10.3390/catal9020188

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