Next Article in Journal
Deep Control of Linear Oligomerization of Glycerol Using Lanthanum Catalyst on Mesoporous Silica Gel
Next Article in Special Issue
Selective Hydrogenation of Naphthalene over γ-Al2O3-Supported NiCu and NiZn Bimetal Catalysts
Previous Article in Journal
A Bioorthogonally Synthesized and Disulfide-Containing Fluorescence Turn-On Chemical Probe for Measurements of Butyrylcholinesterase Activity and Inhibition in the Presence of Physiological Glutathione
Previous Article in Special Issue
Electrocatalytic Degradation of Azo Dye by Vanadium-Doped TiO2 Nanocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Aqueous Miscible Organic LDH Derived Ni-Based Catalysts for Efficient CO2 Methanation

1
College of Environmental Science and Engineering, Beijing Forestry University, 35 Qinghua East Road, Haidian District, Beijing 100083, China
2
Department of Chemical and Process Engineering, University of Surrey, Guildford GU2 7XH, UK
3
Chemistry Department, Heterogeneous Catalysis Lab, University of Cyprus, 1 University Ave., University Campus, Nicosia 2109, Cyprus
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(10), 1168; https://doi.org/10.3390/catal10101168
Submission received: 2 August 2020 / Revised: 6 October 2020 / Accepted: 8 October 2020 / Published: 12 October 2020

Abstract

:
Converting CO2 to methane via catalytic routes is an effective way to control the CO2 content released in the atmosphere while producing value-added fuels and chemicals. In this study, the CO2 methanation performance of highly dispersed Ni-based catalysts derived from aqueous miscible organic layered double hydroxides (AMO-LDHs) was investigated. The activity of the catalyst was found to be largely influenced by the chemical composition of Ni metal precursor and loading. A Ni-based catalyst derived from AMO-Ni3Al1-CO3 LDH exhibited a maximum CO2 conversion of 87.9% and 100% CH4 selectivity ascribed to both the lamellar catalyst structure and the high Ni metal dispersion achieved. Moreover, due to the strong Ni metal–support interactions and abundant oxygen vacancy concentration developed, this catalyst also showed excellent resistance to carbon deposition and metal sintering. In particular, high stability was observed after 19 h in CO2/H2 reaction at 360 °C.

1. Introduction

The concentration of CO2 in the atmosphere has risen from 270 to 385 ppm in the past 200 years which has been causing serious problems such as the greenhouse effect, global climate change, and glaciers melting [1,2,3]. Regarding these environmental issues, three ways to reduce CO2 content in the atmosphere have been suggested [4,5,6,7], namely CO2 emission reduction, CO2 capture and storage (CCS), and CO2 conversion and utilization.
CO2 is chemically stable and it needs to be activated in the presence of hydrogen in order to convert it to CH4. Although CO2 activation is an endothermic process, thus requiring energy input, it can usually be integrated with fuel production from renewable energy, making CO2 conversion in the Power-to-Gas (PtG) process a very promising green technology [4,8,9]. Due to a series of technological issues of storage and transportation, renewable energy can hardly be a primary source of energy for the society [10].
CO2 methanation appears to be one of the largely investigated scientific topics [11,12]. Herein, CO2 is hydrogenated to methane at relatively high temperatures, ca. 300–400 °C, known as the Sabatier reaction, which is a very promising approach for the storage of renewable energy [6]. Based on the PtG concept, electricity generated by renewable energy is then used to produce hydrogen by electrolysis. The thus formed hydrogen is then used to produce selectively CH4 by the reaction with CO2 [13,14,15].
CO2(g)+ 4 H2(g) → CH4(g) + 2 H2O(g) (ΔH° = −165 kJ/mol)
Presently, CO2 activation catalysts are under extensive investigation. Commonly used catalysts are group VIII noble metal-based catalysts, such as Rh, Ru, etc [16,17]. Moreover, non-precious metal-based in the group VIII, like Ni, Co, and Fe can also be used [18,19]. Noble metal-based catalysts show the best methanation catalytic performance, but their high cost limits large-scale utilization [20,21]. Among all the non-precious metals, although Fe-based catalysts are preferable because of their very low cost, they suffer from problems of low catalytic activity, coking, and deactivation after long time-on-stream [22]. On the other hand, Ni-based catalysts are currently used in Sabatier reaction due to their high activity and low cost [23,24,25]. For Ni-based catalysts, γ-Al2O3 has been widely used as support for its large specific surface area (SSA), high-temperature thermal resistance, and good chemical stability [26,27]. Benefiting from these features, Ni/γ-Al2O3 composite is the most broadly used catalytic system for CO2 methanation.
LDHs is a family of ionic lamellar solid compounds consisting of positively charged brucite-like layers and exchangeable interlayer anions [28,29]. Group VIII metal ions can be inserted into the LDH precursor and then be reduced to form catalytically active metal nanoparticles [30,31,32]. Thus, the LDHs derived materials can be tuned for the carbon dioxide methanation to fulfill the following requirements: high basicity that is crucial for CO2 activation (adsorption strength and capacity), and appropriate nickel dispersion and reducibility, necessary for this redox process [33,34,35]. Liu et al. [36] demonstrated that the interaction between metal nanoparticles and support can be strengthened due to chemical bonding, thus improving the anti-coking and anti-sintering properties of the supported metal catalysts. In addition, a novel method called aqueous miscible organic (AMO) treatment was described in our previous work to increase the SSA and total pore volume of the LDHs materials [30]. In this method, the LDHs were synthesized by the conventional co-precipitation strategy, but the final wet slurry was washed with an AMO solvent. If organic solvents are completely miscible with water, they will then replace surface bound water, which can often lead to dispersion into thin nanosheets or exfoliation to even single layers [37,38]. AMO treatment gives a much-diminished driving force for aggregation to dense agglomerates of LDHs, while improving the SSA and surface basic site density of catalysts, and ultimately enhancing the catalytic activity.
Herein, we report the use of AMO-LDH as precursors for the preparation of solids suitable for the CO2 methanation reaction. Ni-based CO2 methanation catalysts derived from NiAl LDHs precursors have been prepared by co-precipitation and further treatment with the AMO method [30,39,40]. Precursors and catalysts were characterized by powder X-ray diffraction (XRD), scanning electron microscope (SEM), and surface texture. The impact of Ni loading (wt%) and the factors influencing the CO2 conversion and CH4 selectivity over these AMO-NiAl layered double oxides (LDOs) catalysts, as well as their stability during a long-term CO2 methanation reaction were examined.

2. Results and Discussion

2.1. Influence of Interlayer Anions of AMO-Ni3Al1 LDHs on CO2 Methanation

Figure 1 shows the powder XRD patterns of LDH precursors with different interlayer anions (NO3 and CO32−) before and after reduction. LDH precursors present typical characteristic peaks at 11.7°, 22.8° and 35.1°, corresponding to the reflections of (003), (006) and (012) facets of the LDH crystal phase, respectively, indicating the successful synthesis of LDHs [41]. The peak intensity of Ni3Al1-CO3 LDH is higher than that of Ni3Al1-NO3 LDH, suggesting that CO32− intercalated LDH is better crystallized than the NO3 intercalated one. The XRD patterns of reduced samples show two intensive peaks at 2θ = 44.4 and 51.8°, corresponding to the reflections of (111) and (200) facets of metallic Ni, and no diffraction peaks due to the NiO phase could be observed [42]. This result shows that NiO is rather completely reduced into metallic Ni. However, if strong Ni support interactions prevail, NiO might not be completely reduced into metallic Ni since small NiO particles (<4 nm) cannot be detected by XRD. Moreover, the γ-Al2O3 phase was not detected, result that might be related to its very low crystallinity. According to literature reports [30], the Al2O3 phase derived from LDH precursors generally exists in an amorphous phase. In order to better understand the dispersion of Ni on the catalyst surface, the TEM image and EDX element mapping images of the reduced AMO-Ni3Al-CO3 LDH solid are shown in Figure 2 and Figure 3, respectively. Black dots in the TEM image represent the Ni particles, where a mean diameter of ~10 nm was estimated. From Figure 3b, it can be observed that Ni is evenly dispersed in the sample. Overall, it can be concluded that highly dispersed Ni-based catalysts were successfully synthesized from Ni-Al LDH precursors.
The CO2 conversion and CH4 selectivity over the AMO-Ni3Al1-CO3-LDO and AMO-Ni3Al1-NO3-LDO solids were compared and results are presented in Figure 4. These two catalysts show extremely high CH4 selectivity, ca. 100% at T < 400 °C, although the conversion level is not remarkable in the low-temperature window, i.e., temperatures lower than 260 °C. Above 450 °C, the CH4 selectivity starts to decrease, due to the competition with the reverse water-gas shift reaction (RWGS) which converts CO2 to CO through hydrogenation (CO2 + H2 → CO + H2O) [43]. For both catalysts, the CO2 conversion reaches maximum value at 360 °C, namely 87.9 and 84.2% over AMO-Ni3Al1-CO3-LDO and AMO-Ni3Al1-NO3-LDO, respectively. According to our previous study [44], interlayer anions influence the morphology and the interlayer distance of LDH, and the LDH with CO32− interlayer has a larger SSA than the LDH with NO3 interlayer. The values of SSA of the pre-reduced catalysts, namely 158 m2/g (AMO-Ni3Al1-CO3-LDO) and 134 m2/g (AMO-Ni3Al1-NO3-LDO) are in good agreement with the previous study. Owing to the larger SSA, Ni produced after reduction will be likely better exposed, which might be one of the intrinsic reasons for the better activity performance of AMO-Ni3Al1-CO3-LDO catalyst.

2.2. The Influence of Catalysts Preparation Procedure on the CO2 Methanation

The N2 adsorption-desorption isotherms were measured for the as-synthesized carbonate intercalated LDHs derived catalysts, AMO-Ni3Al1-CO3-LDO and C-Ni3Al1-CO3-LDO, by coprecipitation and AMO treatment. These pre-reduced catalysts show different SSA, namely 117 m2/g (C-Ni3Al1-CO3-LDO) and 158 m2/g (AMO-Ni3Al1-CO3-LDO). AMO-Ni3Al1-CO3-LDO, indicating that AMO treatment has a positive effect on increasing the SSA of LDH-derived solids. As previously reported [30,37], AMO treatment greatly reduces the driving force for aggregation to dense agglomerates, resulting in the increase of SSA.
The CO2 methanation catalytic performance (CO2-conversion and CH4-selectivity) of AMO-Ni3Al1-CO3-LDO and C-Ni3Al1-CO3-LDO solids is illustrated in Figure 5. There is a significant difference in the CO2-conversion (%) over the two catalysts between 250 and 350 °C. The optimal CO2 conversion of AMO-Ni3Al1-CO3-LDO is 88%, while that of C-Ni3Al1-CO3-LDO is 82.5%. As discussed above, the AMO treatment is beneficial to increase the SSA of the solid, and this could lead to better Ni dispersion and thus to an increased CO2 methanation activity per gram of solid basis.

2.3. Influence of Ni/Al Ratio on Catalyst CO2 Methanation Activity

A series of NiAl-LDHs with different Ni/Al atom ratios were synthesized and their XRD patterns are shown in Figure 6. The X-ray diffraction peaks at 11.76, 22.84 and 35.2° prove that LDH precursors were successfully synthesized, and there are no obvious differences in the crystal structure of LDHs with different Ni/Al ratios. The precursors were then calcined and reduced (see Section 2.1) before CO2 methanation activity tests. The CO2 conversion (%) and CH4 selectivity (%) vs temperature profiles obtained over these NiAl-LDHs catalysts are presented in Figure 7. AMO-Ni2Al1-CO3-LDO presented the lowest CO2 conversion among all the samples, especially at low temperatures. Upon increasing of the Ni/Al ratio to 3:1, the CO2 conversion increase. As the Ni/Al ratio continues to increase to 4:1, the CO2 conversion levels off, and becomes almost constant. This is likely because overloaded Ni can cause aggregation of Ni particles [45]. Hence, we identify a threshold for the positive impact of increasing Ni content on the catalytic CO2 methanation activity being the Ni/Al ratio 4:1 the experimental limit to boost conversion. As for the selectivity to CH4, by increasing the Ni/Al ratio no change is observed except a slight decrease at high temperatures (ca. 500 °C, Figure 7) due to the RWGS reaction.

2.4. Influence of Reduction Temperature in Hydrogen on Catalyst CO2 Methanation Activity

Since the active component of the as-synthesized CO2 methanation catalysts is the Ni metal, NiO (formed after calcination) should be reduced to metallic Ni (H2 is used) before catalytic activity testing. The SEM images of AMO-Ni3Al1-CO3 LDH, AMO-Ni3Al1-CO3 LDO, and reduced AMO-Ni3Al1-CO3 LDO solids are presented in Figure 8. Stacked and interlayered platelet-like crystals of LDH particles of a crystal particle size of ~60 nm are observed in Figure 8a, which is consistent with previous reports [46,47]. This typical platelet-like morphology remains after calcination. However, the morphology greatly changed after reduction in hydrogen.
As shown in Figure 8c, spherical particles with a diameter of 100 ~ 200 nm are formed. The hydrogen treatment conditions had a great influence on the particle dispersion, morphology and structure, and these features would eventually affect the activity performance of the catalyst [30,48]. The powder XRD patterns of samples reduced at 500 and 600 °C are presented in Figure 9. After hydrogen treatment at 600 °C for 1 h, the sample shows relatively strong diffraction peaks at 44.48 and 51.67° corresponding to the Ni phase. The main diffraction peaks for the NiO phase (ca. 37.25 and 44.48°) were not detected, indicating that NiO has been fully reduced to Ni [49]. However, the sample exhibited relatively weak peaks of Ni phase and clear peaks of NiO phase after H2 treatment at 500 °C.
The H2-TPR profile of the AMO-Ni3Al1-CO3-LDO solid (Figure 10) shows its reducibility characteristics. It can be inferred that complete reduction of NiO cannot be accomplished at 500 °C, and it is envisaged that the full reduction temperature is at least 550 °C [50]. This is consistent with the powder XRD results that the sample reduced at 500 °C is not as complete as that reduced at 600 °C. It can be concluded that there exist strong metal–support interactions between NiO and Al2O3 surface [36,51,52,53].
The effect of reduction temperature on the CO2 conversion and CH4 selectivity over the AMO-Ni3Al1-CO3-LDO solid was investigated and results are presented in Figure 11. In all case, practically 100% CH4-selectivity (no CO formation) was detected below 400 °C. The CO2 conversion reaches its maximum value at 360 °C, namely 87.9%, 86.3%, 85.5%, and 77.5% for the catalysts reduced at 600, 550, 500, and 400 °C, respectively. A significant improvement of CO2 conversion can be seen after increasing the hydrogen reduction temperature from 400 to 600 °C, especially for the CO2 methanation reaction occurred at 260 °C. As discussed above, with the increase of reduction temperature, more NiO was converted into Ni, which leads to a high activity. Based on the results of powder XRD, H2-TPR and catalytic tests, it can be stated that the degree of supported Ni reduction is influenced by the hydrogen reduction temperature and it has an important impact on the CO2 conversion. It appears that 600 °C is the best reduction temperature for the present LDHs derived Ni-based CO2 methanation catalysts. However, according to the H2-TPR profile, the catalyst should be reduced completely at 550 °C. Since the H2-TPR experiment is conducted dynamically, the reduction peak maximum may not appear at the exact temperature recorded. To ensure the thorough conversion of NiO into Ni, the hydrogen reduction temperature of the sample was set at 600 °C.

2.5. Stability Performance of AMO-Ni3Al1-CO3-LDO

A stability test was performed on AMO-Ni3Al1-CO3-LDO at 360 °C for 19 h and results are shown in Figure 12. Before any measurements, the catalyst was reduced at 600 °C for 1 h. The results obtained indicate that AMO-Ni3Al1-CO3-LDO derived catalyst showed enhanced stability for long time-on-stream (TOS) since no obvious deactivation was observed (Figure 12). The CO2 conversion remained above 87% and no CO was detected in the effluent gas stream. After the stability test, the spent catalyst was characterized by SEM. Figure 13 indicates the surface morphology of the spent catalyst. Compared with Figure 8c, the SEM images of spent catalysts show no obvious change in the particle configuration, which means that there were no structural changes in the solid catalyst. It was reported [22] that deactivation of CO2 methanation catalysts is mainly caused by possible structural changes, such as sintering of the metal, and the generation of carbon deposits. By using AMO-Ni3Al1-CO3 LDH as precursor, the obtained supported Ni nanoparticles exhibit a high degree of dispersion and strong metal–support interactions. Furthermore, the LDO support may provide oxygen vacancies [54], which are beneficial to the improvement of anti-sintering and anti-coking properties by providing an alternative route for CO2 dissociation into CO and O (lattice oxygen), the latter being able to oxidize carbon (formed on Ni) formed by the reverse Boudouard reaction [55], so as to ensure its high activity during long-term operation.

3. Experimental

3.1. Preparation of Catalysts

LDHs were first synthesized by the co-precipitation method. An aqueous solution containing nitrates of the metallic salts, ca. Ni(NO3)2·6H2O and Al(NO3)3·9H2O were added dropwise into a vigorously stirred Na2CO3 (NaNO3 for NiAl-NO3 LDH) solution. The pH of the resulting solution was kept constant at 10 by the addition of NaOH solution (4 M). The resulting slurry was stirred continuously for ~12 h at 30 °C, then filtered and washed several times with deionized water until pH reached 7, followed by drying at 60 °C in an oven. For AMO treatment, after the above-mentioned washing step, the resulting slurry was re-dispersed in ethanol for 2 h. The final solids were collected by filtration and then dried at 60 °C for further characterization. To investigate the influence of Ni content, NiAl LDHs with different Ni/Al atom ratios (x = Ni/Al) were synthesized, which were denoted as NixAl1-LDH (x = 2, 3, and 4). The obtained LDHs were calcined at 400 °C for 5 h and reduced in 10 vol% H2/Ar gas atmosphere (1 bar) before being used in any catalytic experiment.

3.2. Catalysts Characterization

Powder X-ray diffraction (PXRD) patterns of the solid AMO-NiAl LDOs samples were recorded using a Shimadzu XRD-7000 instrument in reflection mode and a Cu Kα radiation. The X-ray tube was operated at 40 kV and 40 mA while the accelerating voltage was set at 40 kV with 30 mA current (λ = 1.542 Å). Diffraction patterns were recorded within the range of 2θ = 5–75° with a scanning rate of 5°/min and a step size of 0.02°. Textural properties of the solids were analyzed using nitrogen adsorption-desorption isotherms obtained from a physisorption analyzer (SSA–7000, Builder, Beijing, China). SSA was estimated using the Brunauer−Emmett−Teller (BET) method. Before each measurement, ~0.1 g catalyst sample was degassed in a N2/He gas mixture at 220 °C for 4 h.
Hydrogen temperature-programmed reduction (H2-TPR) experiments were carried out in a multifunction chemisorption analyzer (PCA-1200, Builder, Beijing, China) equipped with a quartz U-tube reactor and a thermal conductivity detector (TCD) for gas analysis. For each test, a ~0.05 g sample was utilized. Before switching to the H2/Ar gas stream, the sample was pretreated in Ar (40 mL/min) at 200 °C for 30 min and then cooled in Ar gas flow to 50 °C. The sample was then heated from 50 to 800 °C with a ramping rate of 10 °C/min in 5 vol% H2/Ar gas mixture flow (30 mL/min). A field emission scanning electron microscope (SU-8010, Hitachi, Tokyo, Japan) was used to characterize the morphology of Ni33Al LDO (secondary agglomerated particles). A carbon tape was used to stick the powder to the SEM stage. Transmission electron microscopy (TEM) images were obtained on a FEI Tecnai G2 F30 (Hillsboro, USA) which was operated at 300 kV. Energy-dispersive X-ray (EDX) mapping characterizations of the composite oxides-based catalysts was made to investigate the distribution of the elements in the samples using the same TEM instrument.

3.3. Catalytic Activity Tests

The CO2 methanation catalytic activity tests of the synthesized AMO-NiAl LDOs were carried out at 1 atm pressure in a fixed-bed stainless-steel micro-reactor. The stainless-steel micro-reactor was installed in a vertical split-tube furnace equipped with a proportional-integral-derivative (PID) temperature controller and several mass flow controllers. Before the catalytic performance test, a catalyst amount of Wcat = 0.1 g was first pretreated in 20 vol% H2/N2 gas mixture (50 mL/min) at 600 °C for 1 h, and then cooled to 200 °C in N2 gas flow. Subsequently, a gas mixture of H2, CO2 and N2 with a molar ratio of H2/CO2/Ar = 4:1:5 (ca. 40 vol% H2/10 vol% CO2/50 vol% Ar) was introduced into the reactor with a total flow rate of 80 mL/min corresponding to a space velocity of 48,000 mL gcat−1 h−1. The methanation reaction was tested in the 200–500 °C range. After the reaction rate was stabilized at a given temperature after about 30 min, the composition of the effluent gas stream was analyzed online using a gas chromatography (SP-7890, Shandong Lunan, Zaozhuang, China) equipped with a TCD and FID detectors. The CO2 conversion and CH4 selectivity were estimated from the following Equations (2) and (3), respectively:
C O 2   c o n v e r s i o n ( % ) = [ [ C O 2 , i n ] [ C O 2 , o u t ] [ C O 2 , i n ] ] × 100 %
C H 4   s e l e c t i v i t y ( % ) = [ C H 4 ] [ C H 4 ] + [ C O ] × 100 %
In Equation (3), the [CO] concentration at the outlet of reactor refers to the non-selective CO2 hydrogenation reaction towards CO and H2O products (reverse water-gas shift reaction). Also, given the CO2 feed composition used (10 vol%), the change in the outlet total molar flow rate of the methanation reaction was small with respect to the inlet total molar flow rate, and this was considered in the estimation of CH4 selectivity (Equation (3)).

4. Conclusions

Selective CO2 methanation reaction was conducted over NiAl LDH derived catalysts which were prepared by coprecipitation and AMO treatment. AMO treated LDH exhibits higher SSA, which benefits Ni dispersion, than the conventional one. Highly dispersed Ni particles on the catalyst surface led to higher catalytic activity for CO2 methanation. The Ni loading was found to play a key role in determining the catalytic activity by influencing the concentration of active sites along with the SSA. However, excessive Ni loading may result in Ni particle agglomeration, thus reducing the catalytic activity. As shown by the performed experiments, an optimal Ni/Al atom ratio of 3:1 was found. The catalyst reduction temperature by hydrogen was found to also play a role in its activity, which should be higher than 550 °C to ensure complete reduction of NiO to Ni. After hydrogen reduction at 600 °C, AMO-Ni3Al1-CO3 LDH derived catalyst showed the highest CO2 conversion (87.9%) and 100% CH4 selectivity at 360 °C, due to the enhanced dispersion of Ni nanoparticles and SSA. Moreover, this AMO-Ni3Al1-CO3 LDH derived catalyst shows remarkable stability in a 19-h activity test (above 87%) as the result of maintaining a highly dispersed Ni state, strong metal–support interactions, and abundant oxygen vacancy concentration.

Author Contributions

Conceptualization, L.H. and A.M.E.; Formal analysis, Z.W. and A.M.E.; Funding acquisition, Q.W.; Investigation, Z.W.; Methodology, Z.W. and L.H.; Project administration, Q.W.; Resources, T.R.R.; Supervision, Q.W.; Validation, L.H. and A.M.E.; Visualization, T.R.R.; Writing—original draft, Z.W.; Writing—review & editing, L.H., T.R.R., A.M.E. and Q.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Fundamental Research Funds for the Central Universities (BLX201935, 2019JQ03015), the International Science and Technology Cooperation Project of Bingtuan (2018BC002), the Beijing Municipal Education Commission through the Innovative Transdisciplinary Program “Ecological Restoration Engineering”, and International Science and Technology Cooperation Project of Bingtuan (No.2018BC002).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mikkelsen, M.; Jørgensen, M.; Krebs, F.C. The teraton challenge. A review of fixation and transformation of carbon dioxide. Energy Environ. Sci. 2010, 3, 43–81. [Google Scholar] [CrossRef]
  2. Smol, J.P. Climate Change: A planet in flux. Nature 2012, 483, S12–S15. [Google Scholar] [CrossRef] [PubMed]
  3. Liang, S.; Altaf, N.; Huang, L.; Gao, Y.; Wang, Q. Electrolytic cell design for electrochemical CO2 reduction. J. CO2 Util. 2020, 35, 90–105. [Google Scholar] [CrossRef]
  4. Li, W.; Nie, X.; Jiang, X.; Zhang, A.; Ding, F.; Liu, M. ZrO2 Support Imparts Superior Activity and Stability of Co Catalysts for CO2 Methanation. Appl. Catal. B Environ. 2018, 220, 397–408. [Google Scholar] [CrossRef]
  5. Rahmani, S.; Rezaei, M.; Meshkani, F. Preparation of promoted nickel catalysts supported on mesoporous nanocrystalline gamma alumina for carbon dioxide methanation reaction. J. Ind. Eng. Chem. 2014, 20, 4176–4182. [Google Scholar] [CrossRef]
  6. Renda, S.; Ricca, A.; Palma, V. Study of the effect of noble metal promotion in Ni-based catalyst for the Sabatier reaction. Int. J. Hydrog. Energy 2020. [Google Scholar] [CrossRef]
  7. Branco, J.B.; Brito, P.E.; Ferreira, A.C. Methanation of CO2 over nickel-lanthanide bimetallic oxides supported on silica. Chem. Eng. J. 2020, 380, 122465. [Google Scholar] [CrossRef]
  8. García-Gutiérrez, P.; Cuéllar-Franca, R.M.; Reed, D.; Dowson, G.; Styring, P.; Azapagic, A. Environmental sustainability of cellulosesupported solid ionic liquids for CO2 capture. Green Chem. 2019, 21, 4100–4114. [Google Scholar] [CrossRef] [Green Version]
  9. Zhou, Z.; Sun, N.; Wang, B.; Han, Z.; Cao, S.; Hu, D.; Zhu, T.; Shen, Q.; Wei, W. 2D-Layered Ni-MgO-Al2O3 Nanosheets for Integrated Capture and Methanation of CO2. ChemSusChem 2019, 13, 360–368. [Google Scholar] [CrossRef]
  10. Dou, L.; Yan, C.; Zhong, L.; Zhang, D.; Zhang, J.; Li, X.; Xiao, L. Enhancing CO2 methanation over a metal foam structured catalyst by electric internal heating. Chem. Commun. 2020, 56, 205–208. [Google Scholar] [CrossRef]
  11. Pastor-Pérez, L.; Patel, V.; Le Saché, E.; Reina, T.R. CO2 methanation in the presence of methane: Catalysts design and effect of methane concentration in the reaction mixture. J. Energy Inst. 2020, 93, 415–424. [Google Scholar] [CrossRef]
  12. Pastor-Pérez, L.; Saché, E.L.; Jones, C.; Gu, S.; Arellano-Garcia, H.; Reina, T.R. Synthetic natural gas production from CO2 over Ni-x/CeO2-ZrO2 (x = Fe, Co) catalysts: Influence of promoters and space velocity. Catal. Today 2018, 317, 108–113. [Google Scholar] [CrossRef]
  13. Jarvis, S.M.; Samsatli, S. Technologies and infrastructures underpinning future CO2 value chains: A comprehensive review and comparative analysis. Renew. Sustain. Energy Rev. 2018, 85, 46–68. [Google Scholar] [CrossRef]
  14. Thema, M.; Bauer, F.; Sterner, M. Power-to-Gas: Electrolysis and methanation status review. Renew. Sustain. Energy Rev. 2019, 112, 775–787. [Google Scholar] [CrossRef]
  15. Dannesboe, C.; Hansen, J.B.; Johannsen, I. Catalytic methanation of CO2 in biogas: Experimental results from a reactor at full scale. React. Chem. Eng. 2020, 5, 183–189. [Google Scholar] [CrossRef] [Green Version]
  16. Beuls, A.; Swalus, C.; Jacquemin, M.; Heyen, G.; Karelovic, A.; Ruiz, P. Methanation of CO2: Further insight into the mechanism over Rh/γ-Al2O3 catalyst. Appl. Catal. B Environ. 2012, 113-114, 2–10. [Google Scholar] [CrossRef]
  17. Wang, F.; He, S.; Chen, H.; Wang, B.; Zheng, L.; Wei, M.; Evans, D.G.; Duan, X. Active Site Dependent Reaction Mechanism over Ru/CeO2 Catalyst toward CO2 Methanation. J. Am. Chem. Soc. 2016, 138, 6298–6305. [Google Scholar] [CrossRef]
  18. Aziz, M.A.A.; Jalil, A.A.; Triwahyono, S.; Ahmad, A. CO2 methanation over heterogeneous catalysts: Recent progress and future prospects. Green Chem. 2015, 17, 2647–2663. [Google Scholar] [CrossRef]
  19. Sun, J.; Wang, Y.; Zou, H.; Guo, X.; Wang, Z.-J. Ni catalysts supported on nanosheet and nanoplate γ-Al2O3 for carbon dioxide methanation. J. Energy Chem. 2019, 29, 3–7. [Google Scholar] [CrossRef] [Green Version]
  20. Ab Halim, A.Z.; Ali, R.; Wan Abu Bakar, W.A. CO2/H2 methanation over M*/Mn/Fe-Al2O3 (M*: Pd, Rh, and Ru) catalysts in natural gas; optimization by response surface methodology-central composite design. Clean Technol. Environ. Policy 2014, 17, 627–636. [Google Scholar] [CrossRef]
  21. Zamani, A.H.; Ali, R.; Abu Bakar, W.A.W. Optimization of CO2 methanation reaction over M*/Mn/Cu–Al2O3 (M*: Pd, Rh and Ru) catalysts. J. Ind. Eng. Chem. 2015, 29, 238–248. [Google Scholar] [CrossRef]
  22. Lee, W.J.; Li, C.; Prajitno, H.; Yoo, J.; Patel, J.; Yang, Y.; Lim, S. Recent trend in thermal catalytic low temperature CO2 methanation: A critical review. Catal. Today 2020. [Google Scholar] [CrossRef]
  23. Chein, R.-Y.; Wang, C.-C. Experimental Study on CO2 Methanation over Ni/Al2O3, Ru/Al2O3, and Ru-Ni/Al2O3 Catalysts. Catalysts 2020, 10, 1112. [Google Scholar] [CrossRef]
  24. Ghaib, K.; Ben-Fares, F.-Z. Power-to-Methane: A state-of-the-art review. Renew. Sustain. Energy Rev. 2018, 81, 433–446. [Google Scholar] [CrossRef]
  25. Li, S.; Liu, G.; Zhang, S.; An, K.; Ma, Z.; Wang, L.; Liu, Y. Cerium-modified Ni-La2O3/ZrO2 for CO2 methanation. J. Energy Chem. 2020, 43, 155–164. [Google Scholar] [CrossRef] [Green Version]
  26. Solis-Garcia, A.; Fierro-Gonzalez, J.C. Mechanistic Insights into the CO2 Methanation Catalyzed by Supported Metals: A Review. J. Nanosci. Nanotechnol. 2019, 19, 3110–3123. [Google Scholar] [CrossRef] [PubMed]
  27. Frontera, P.; Macario, A.; Monforte, G.; Bonura, G.; Ferraro, M.; Dispenza, G.; Antonucci, V.; Aricò, A.S.; Antonucci, P.L. The role of Gadolinia Doped Ceria support on the promotion of CO2 methanation over Ni and Ni Fe catalysts. Int. J. Hydrog. Energy 2017, 42, 26828–26842. [Google Scholar] [CrossRef]
  28. Wang, J.; Mei, X.; Huang, L.; Zheng, Q.; Qiao, Y.; Zang, K.; Mao, S.; Yang, R.; Zhang, Z.; Gao, Y.; et al. Synthesis of layered double hydroxides/graphene oxide nanocomposite as a novel high-temperature CO2 adsorbent. J. Energy Chem. 2015, 24, 127–137. [Google Scholar] [CrossRef]
  29. Huang, L.; Wang, J.; Gao, Y.; Qiao, Y.; Zheng, Q.; Guo, Z.; Zhao, Y.; O’Hare, D.; Wang, Q. Synthesis of LiAl2-layered double hydroxides for CO2 capture over a wide temperature range. J. Mater. Chem. A 2014, 2, 18454–18462. [Google Scholar] [CrossRef]
  30. Wang, Q.; O’Hare, D. Large-scale synthesis of highly dispersed layered double hydroxide powders containing delaminated single layer nanosheets. Chem. Commun. 2013, 49, 6301. [Google Scholar] [CrossRef]
  31. Climent, M.J.; Corma, A.; Iborra, S.; Epping, K.; Velty, A. Increasing the basicity and catalytic activity of hydrotalcites by different synthesis procedures. J. Catal. 2004, 225, 316–326. [Google Scholar] [CrossRef]
  32. Gao, Y.; Zhao, Y.; Qiu, L.; Guo, Z.; O’Hare, D.; Wang, Q. Preparation of 4,4′-diaminostilbene-2,2′-disulfonic acid intercalated LDH/polypropylene nanocomposites with enhanced UV absorption property. Polym. Compos. 2017, 38, 1937–1947. [Google Scholar] [CrossRef]
  33. Bian, L.; Wang, W.; Xia, R.; Li, Z. Ni-based catalyst derived from Ni/Al hydrotalcite-like compounds by the urea hydrolysis method for CO methanation. RSC Adv. 2016, 6, 677–686. [Google Scholar] [CrossRef]
  34. Marocco, P.; Morosanu, E.A.; Giglio, E.; Ferrero, D.; Mebrahtu, C.; Lanzini, A.; Abate, S.; Bensaid, S.; Perathoner, S.; Santarelli, M.; et al. CO2 methanation over Ni/Al hydrotalcite-derived catalyst: Experimental characterization and kinetic study. Fuel 2018, 225, 230–242. [Google Scholar] [CrossRef]
  35. Abelló, S.; Berrueco, C.; Montané, D. High-loaded nickel-alumina catalyst for direct CO2 hydrogenation into synthetic natural gas (SNG). Fuel 2013, 113, 598–609. [Google Scholar] [CrossRef]
  36. Liu, J.; Bing, W.; Xue, X.; Wang, F.; Wang, B.; He, S.; Zhang, Y.; Wei, M. Alkaline-assisted Ni nanocatalysts with largely enhanced low-temperature activity toward CO2 methanation. Catal. Sci. Technol. 2016, 6, 3976–3983. [Google Scholar] [CrossRef]
  37. Chen, C.; Yang, M.; Wang, Q.; Buffet, J.-C.; O’Hare, D. Synthesis and characterisation of aqueous miscible organic-layered double hydroxides. J. Mater. Chem. A 2014, 2, 15102–15110. [Google Scholar] [CrossRef]
  38. Ruengkajorn, K.; Erastova, V.; Buffet, J.-C.; Greenwell, H.C.; O’Hare, D. Aqueous immiscible layered double hydroxides: Synthesis, characterisation and molecular dynamics simulation. Chem. Commun. 2018, 54, 4394–4397. [Google Scholar] [CrossRef] [Green Version]
  39. Chen, C.; Wangriya, A.; Buffet, J.-C.; O’Hare, D. Tuneable ultra high specific surface area Mg/Al-CO3 layered double hydroxides. Dalton Trans. 2015, 44, 16392–16398. [Google Scholar] [CrossRef]
  40. Cermelj, K.; Ruengkajorn, K.; Buffet, J.-C.; O’Hare, D. Layered double hydroxide nanosheets via solvothermal delamination. J. Energy Chem. 2019, 35, 88–94. [Google Scholar] [CrossRef] [Green Version]
  41. Guo, X.; He, H.; Traitangwong, A.; Gong, M.; Meeyoo, V.; Li, P.; Li, C.; Peng, Z.; Zhang, S. Ceria imparts superior low temperature activity to nickel catalysts for CO2 methanation. Catal. Sci. Technol. 2019, 9, 5636–5650. [Google Scholar] [CrossRef]
  42. Yan, Y.; Dai, Y.; He, H.; Yu, Y.; Yang, Y. A novel W-doped Ni-Mg mixed oxide catalyst for CO2 methanation. Appl. Catal. B Environ. 2016, 196, 108–116. [Google Scholar] [CrossRef]
  43. Yang, L.; Pastor-Pérez, L.; Gu, S.; Sepúlveda-Escribano, A.; Reina, T.R. Highly efficient Ni/CeO2-Al2O3 catalysts for CO2 upgrading via reverse water-gas shift: Effect of selected transition metal promoters. Appl. Catal. B Environ. 2018, 232, 464–471. [Google Scholar] [CrossRef]
  44. Wang, Q.; Wu, Z.; Tay, H.H.; Chen, L.; Liu, Y.; Chang, J.; Zhong, Z.; Luo, J.; Borgna, A. High temperature adsorption of CO2 on Mg-Al hydrotalcite: Effect of the charge compensating anions and the synthesis pH. Catal. Today 2011, 164, 198–203. [Google Scholar] [CrossRef]
  45. Zhen, W.; Li, B.; Lu, G.; Ma, J. Enhancing catalytic activity and stability for CO2 methanation on Ni@MOF-5 via control of active species dispersion. Chem. Commun. 2015, 51, 1728–1731. [Google Scholar] [CrossRef]
  46. Mak Yu, T.; Caroline Reis Meira, A.; Cristina Kreutz, J.; Effting, L.; Mello Giona, R.; Gervasoni, R.; Amado de Moura, A.; Maestá Bezerra, F.; Bail, A. Exploring the surface reactivity of the magnetic layered double hydroxide lithium-aluminum: An alternative material for sorption and catalytic purposes. Appl. Surf. Sci. 2019, 467–468, 1195–1203. [Google Scholar] [CrossRef]
  47. Yan, Q.; Nie, Y.; Yang, R.; Cui, Y.; O’Hare, D.; Wang, Q. Highly dispersed CuyAlOx mixed oxides as superior low-temperature alkali metal and SO2 resistant NH3 -SCR catalysts. Appl. Catal. A Gen. 2017, 538, 37–50. [Google Scholar] [CrossRef]
  48. Wang, L.; Niu, X.; Chen, J. SiO2 supported Ni-In intermetallic compounds: Efficient for selective hydrogenation of fatty acid methyl esters to fatty alcohols. Appl. Catal. B Environ. 2020, 278, 119293. [Google Scholar] [CrossRef]
  49. Li, S.; Yang, Y.; Huang, S.; He, Z.; Li, C.; Li, D.; Ke, B.; Lai, C.; Peng, Q. Adsorption of humic acid from aqueous solution by magnetic Zn/Al calcined layered double hydroxides. Appl. Clay Sci. 2020, 188, 105414. [Google Scholar] [CrossRef]
  50. Ye, R.-P.; Gong, W.; Sun, Z.; Sheng, Q.; Shi, X.; Wang, T.; Yao, Y.; Razink, J.J.; Lin, L.; Zhou, Z.; et al. Enhanced stability of Ni/SiO2 catalyst for CO2 methanation: Derived from nickel phyllosilicate with strong metal-support interactions. Energy 2019, 188, 116059. [Google Scholar] [CrossRef]
  51. Xu, M.; He, S.; Chen, H.; Cui, G.; Zheng, L.; Wang, B.; Wei, M. TiO2-x-Modified Ni Nanocatalyst with Tunable Metal–Support Interaction for Water–Gas Shift Reaction. ACS Catal. 2017, 7, 7600–7609. [Google Scholar] [CrossRef]
  52. Mori, K.; Taga, T.; Yamashita, H. Isolated Single-Atomic Ru Catalyst Bound on a Layered Double Hydroxide for Hydrogenation of CO2 to Formic Acid. ACS Catal. 2017, 7, 3147–3151. [Google Scholar] [CrossRef]
  53. Zhang, J.; Wang, H.; Dalai, A. Development of stable bimetallic catalysts for carbon dioxide reforming of methane. J. Catal. 2007, 249, 300–310. [Google Scholar] [CrossRef]
  54. Li, Z.; Yan, Q.; Jiang, Q.; Gao, Y.; Xue, T.; Li, R.; Liu, Y.; Wang, Q. Oxygen vacancy mediated CuyCo3-yFe1Ox mixed oxide as highly active and stable toluene oxidation catalyst by multiple phase interfaces formation and metal doping effect. Appl. Catal. B Environ. 2020, 269, 118827. [Google Scholar] [CrossRef]
  55. Damaskinos, C.M.; Vasiliades, M.A.; Efstathiou, A.M. The effect of Ti4+ dopant in the 5 wt% Ni/Ce1-xTixO2-δ catalyst on the carbon pathways of dry reforming of methane studied by various transient and isotopic techniques. Appl. Catal. A Gen. 2019, 579, 116–129. [Google Scholar] [CrossRef]
Figure 1. The powder XRD patterns of AMO-Ni3Al-NO3 LDH and AMO-Ni3Al-CO3 LDH precursors and their corresponding reduced samples.
Figure 1. The powder XRD patterns of AMO-Ni3Al-NO3 LDH and AMO-Ni3Al-CO3 LDH precursors and their corresponding reduced samples.
Catalysts 10 01168 g001
Figure 2. TEM image of reduced AMO-Ni3Al-CO3 LDH solid.
Figure 2. TEM image of reduced AMO-Ni3Al-CO3 LDH solid.
Catalysts 10 01168 g002
Figure 3. TEM image of reduced AMO-Ni3Al-CO3 LDH solid (a) and corresponding EDX elemental mapping of AMO-Ni3Al-CO3 LDH derived catalyst for Ni (b), Al (c), and O (d).
Figure 3. TEM image of reduced AMO-Ni3Al-CO3 LDH solid (a) and corresponding EDX elemental mapping of AMO-Ni3Al-CO3 LDH derived catalyst for Ni (b), Al (c), and O (d).
Catalysts 10 01168 g003
Figure 4. CO2 conversion and CH4 selectivity vs T profiles of AMO-Ni3Al1-CO3-LDO and AMO-Ni3Al1-NO3-LDO catalysts for the CO2 methanation reaction.
Figure 4. CO2 conversion and CH4 selectivity vs T profiles of AMO-Ni3Al1-CO3-LDO and AMO-Ni3Al1-NO3-LDO catalysts for the CO2 methanation reaction.
Catalysts 10 01168 g004
Figure 5. CO2 conversion (%) and CH4 selectivity (%) vs T profiles over AMO-Ni3Al1-CO3-LDO and C-Ni3Al1-CO3-LDO catalysts for the CO2 methanation reaction.
Figure 5. CO2 conversion (%) and CH4 selectivity (%) vs T profiles over AMO-Ni3Al1-CO3-LDO and C-Ni3Al1-CO3-LDO catalysts for the CO2 methanation reaction.
Catalysts 10 01168 g005
Figure 6. The powder XRD patterns of LDHs of different Ni/Al atom ratios (2:1, 3:1 and 4:1).
Figure 6. The powder XRD patterns of LDHs of different Ni/Al atom ratios (2:1, 3:1 and 4:1).
Catalysts 10 01168 g006
Figure 7. CO2 conversion (%) and CH4 selectivity (%) vs T profiles over AMO-Ni2Al1-CO3-LDO, AMO-Ni3Al1-CO3-LDO and AMO-Ni4Al1-CO3-LDO catalysts for the CO2 methanation.
Figure 7. CO2 conversion (%) and CH4 selectivity (%) vs T profiles over AMO-Ni2Al1-CO3-LDO, AMO-Ni3Al1-CO3-LDO and AMO-Ni4Al1-CO3-LDO catalysts for the CO2 methanation.
Catalysts 10 01168 g007
Figure 8. SEM images of AMO-Ni3Al1-CO3 LDH (a), AMO-Ni3Al1-CO3 LDO (b), and reduced AMO-Ni3Al1-CO3 LDO (c) solids.
Figure 8. SEM images of AMO-Ni3Al1-CO3 LDH (a), AMO-Ni3Al1-CO3 LDO (b), and reduced AMO-Ni3Al1-CO3 LDO (c) solids.
Catalysts 10 01168 g008
Figure 9. Powder XRD patterns of AMO-Ni3Al1-CO3-LDO reduced in hydrogen at different temperatures, ca. 500 and 600 °C.
Figure 9. Powder XRD patterns of AMO-Ni3Al1-CO3-LDO reduced in hydrogen at different temperatures, ca. 500 and 600 °C.
Catalysts 10 01168 g009
Figure 10. H2-TPR profile of AMO-Ni3Al1-CO3-LDO solid.
Figure 10. H2-TPR profile of AMO-Ni3Al1-CO3-LDO solid.
Catalysts 10 01168 g010
Figure 11. CO2 conversion (%) and CH4 selectivity (%) vs temperature profiles for the CO2 methanation reaction over AMO-Ni3Al1-CO3-LDO reduced in hydrogen at different temperatures, ca. 400, 500, 550 and 600 °C.
Figure 11. CO2 conversion (%) and CH4 selectivity (%) vs temperature profiles for the CO2 methanation reaction over AMO-Ni3Al1-CO3-LDO reduced in hydrogen at different temperatures, ca. 400, 500, 550 and 600 °C.
Catalysts 10 01168 g011
Figure 12. Stability performance of the AMO-Ni3Al1-CO3-LDO catalyst at 360 °C in the CO2 methanation reaction.
Figure 12. Stability performance of the AMO-Ni3Al1-CO3-LDO catalyst at 360 °C in the CO2 methanation reaction.
Catalysts 10 01168 g012
Figure 13. SEM images of used (19 h of CO2 methanation) AMO-Ni3Al1-CO3-LDO catalyst.
Figure 13. SEM images of used (19 h of CO2 methanation) AMO-Ni3Al1-CO3-LDO catalyst.
Catalysts 10 01168 g013

Share and Cite

MDPI and ACS Style

Wang, Z.; Huang, L.; Reina, T.R.; Efstathiou, A.M.; Wang, Q. Aqueous Miscible Organic LDH Derived Ni-Based Catalysts for Efficient CO2 Methanation. Catalysts 2020, 10, 1168. https://doi.org/10.3390/catal10101168

AMA Style

Wang Z, Huang L, Reina TR, Efstathiou AM, Wang Q. Aqueous Miscible Organic LDH Derived Ni-Based Catalysts for Efficient CO2 Methanation. Catalysts. 2020; 10(10):1168. https://doi.org/10.3390/catal10101168

Chicago/Turabian Style

Wang, Ziling, Liang Huang, Tomas Ramirez Reina, Angelos M. Efstathiou, and Qiang Wang. 2020. "Aqueous Miscible Organic LDH Derived Ni-Based Catalysts for Efficient CO2 Methanation" Catalysts 10, no. 10: 1168. https://doi.org/10.3390/catal10101168

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