Next Article in Journal
Photocatalytic Degradation of Losartan with Bismuth Oxychloride: Batch and Pilot Scale Demonstration
Previous Article in Journal
Photocatalytic Systems Based on ZnO Produced by Supercritical Antisolvent for Ceftriaxone Degradation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Effect of the Ni-to-CaO Ratio on Integrated CO2 Capture and Direct Methanation

1
Department of Chemical Engineering, Kyungpook National University, Daegu 41566, Republic of Korea
2
Department of Chemical and Environmental Engineering, University of California–Riverside, Riverside, CA 92521, USA
3
Research Institute of Advanced Energy Technology, Kyungpook National University, Daegu 41566, Republic of Korea
4
Korea Institute of Energy Research, Daejeon 34129, Republic of Korea
5
Department of Material Science and Engineering, University of California–Riverside, Riverside, CA 92521, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 24 June 2023 / Revised: 21 July 2023 / Accepted: 29 July 2023 / Published: 31 July 2023

Abstract

:
Direct methanation in an integrated CO2 capture and utilization system has recently gained considerable attention as a promising approach owing to its simplified process and lower requirement of total thermal energy as compared to conventional CO2 capture and utilization techniques. This study formulated macroporous structured Ni/CaO catal-sorbents by controlling the Ni-to-CaO ratio. The influence of this ratio on the CO2 capture (capacity and kinetics) and direct methanation performances (productivity and kinetics) was evaluated at 500 °C. CO2 capture combined with direct methanation experiments revealed that 10Ni/CaO exhibited the best CO2 capture capacity, kinetics, and CH4 productivity with the thermal stability of Ni and CaO species.

Graphical Abstract

1. Introduction

Carbon dioxide (CO2) emissions are an important cause of global climate change and a major challenge our planet is facing today. The burning of fossil fuels, such as coal, oil, and gas, for energy and transportation is the primary source of CO2 emissions. These emissions trap heat in the atmosphere and lead to global warming, which has various adverse effects on the environment, including sea-level rise, extreme weather events, and the loss of biodiversity [1,2,3,4,5]. CO2 capture and utilization (CCU) technology captures CO2 emissions from various sources and converts them into value-added products. This technology has gained considerable attention as a means to mitigate climate change and greenhouse gas emissions. The captured CO2 can be used to produce fuels, chemicals, and other products through various conversion methods, such as biological, chemical, and electrochemical processes [3,6]. By utilizing the captured CO2, CCU technology not only reduces emissions but also contributes to the circular economy by producing sustainable products. In addition, CO2 utilization may decrease the dependency on fossil fuels, thereby promoting a sustainable future. With continued research and development, CCU technology may become a feasible solution to the challenges of climate change while supporting economic growth [3,7].
Recently, there has been a growing interest in the integrated CO2 capture and utilization (ICCU) technology that uses dual functional materials to combine CO2 capture and utilization in a single reaction system. ICCU technology is promising for mitigating climate change by reducing CO2 emissions while producing valuable products. This technology involves capturing CO2 from industrial processes, transport, and power plants, followed by its utilization in various green technologies [8,9,10,11,12,13,14,15,16,17,18,19,20,21]. Calcium looping technology is one of the effective methods for CO2 capture, which employs a calcium oxide (CaO)-based solid sorbent to chemically absorb CO2 [12,22,23,24]. However, the sintering-induced decrease in CO2 capture capacity during carbonation and regeneration cycles limits the efficiency of these sorbents [23]. To improve their multicycle stability, metal oxide stabilizers are added to the intraparticle region of CaO, which increases the length of diffusional paths, decreases the sintering rate, and enhances the textural properties of these sorbents [12,22,23,24].
Owing to the characteristics of ICCU technology, where CO2 capture and conversion must occur continuously in the same temperature range, only CO2 methanation through the Sabatier reaction is suitable for intermediate temperature conversion. The methanation reaction is the simplest and the most efficient catalytic CO2 utilization process, which converts CO2 and H2 to CH4, which is an energy carrier with the least CO2 emission per unit of energy among fossil fuels [17,18,19,20,25,26,27]. Direct methanation after CO2 capture in flue gas conditions through Ni-CaO occurs as follows [28,29]:
CO2 capture: NiO/CaO(s) + CO2(g) ↔ NiO/CaCO3(s),
Direct methanation: NiO/CaCO3(s) + 4H2(g) ↔ Ni0/CaO(s) + CH4(g) + 2H2O(g).
In our previous paper, an ICCU-direct methanation process using a 10Ni/CaO catal-sorbent was proposed, and the optimum temperature was determined to be 500 °C, considering the high CO2 capture capacity, CH4 selectivity, and thermal stability of the catal-sorbent [28,29]. Herein, Ni/CaO catal-sorbents with different Ni-to-CaO ratios were synthesized by the sol–gel method to determine the effect of these ratios on the macroporous of CaO and Ni dispersion. Textural structure of the Ni/CaO catal-sorbents is critically important since the pore structure enables rapid CO2 diffusion for CO2 capture capacity and the Ni dispersion promotes catalytic activity. Their CO2 capture performance (capacity and kinetics) and direct methanation performance (productivity and kinetics) according to the textural structure were studied. Moreover, the cycle stability of the catal-sorbents in CO2 capture and direct methanation was revealed. The optimal Ni/CaO ratio was proposed through CO2 capture and direct methanation performance results and analysis.

2. Results and Discussion

2.1. Characterization

The ICP-OES metal content, BET surface area, nanoparticle size, and porosity of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents are summarized in Table 1. The Ni metal contents of the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents prepared by the citrate sol–gel method are 1.6, 9.3, and 18.0 wt.%, respectively, which are consistent with the desired metal loading. The BET surface area of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents are 4.59, 9.90, and 7.77 m2/g, respectively, and 10Ni/CaO has the smallest nanoparticle size of 605.8 nm. As can be seen in Figure 1, 2Ni/CaO has a dense morphology, with interconnected nanoparticle structures and poor porosity. Notably, as the Ni-to-CaO ratio increases the porous structures expand, and the nanoparticle size decreases dramatically; 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO exhibited porosities 72.9, 74.7, and 74.4%, respectively. In addition, the Hg porosimeter revealed that macroporous structures > 100 nm increase as the Ni-to-CaO ratio increases (Figure S1). 10Ni/CaO exhibited exceptional BET surface area, nanoparticle size, and porosity among xNi/CaO catal-sorbents. Figure 1d–f present the TEM images and the Ni crystallite size distribution of the catal-sorbents. The energy dispersive spectroscopy (EDS) mapping on the catal-sorbents revealed that Ni nanoparticles of uniform size are dispersed uniformly throughout CaO (Figures S2–S4). Individual Ni particles in 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO had an average crystallite size of 20.2, 27.0, and 45.4 nm, respectively (Table 2). 2Ni/CaO exhibited the highest Ni dispersion (2.43%) and Ni surface area (16.19 m2/gNi), and Ni dispersion and Ni surface area decreased with the increase in the Ni-to-CaO ratio, leading to poor catalytic activity.
In Figure 2a, the XRD patterns of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO in the fresh state display NiO (JCPDS No. 89-7131) and CaO phase (JCPDS No. 48-1467). In addition to the NiO phase, a very small and broad peak of metallic Ni is observed for 20Ni/CaO. 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO have CaO crystallite sizes of 49.7, 50.1, and 32.2 nm, respectively. The crystallite size of CaO decreased with increasing Ni-to-CaO ratio, and the smallest CaO in 20Ni/CaO is expected to be the fastest kinetics in CO2 capture. In Figure 2b, the XPS spectra of Ni 2p3/2 revealed that Ni2+ exists in two states within the catal-sorbents. The peak with the lowest binding energy (853.5–854.5 eV) corresponds to the bulk NiO (Ni2+ (I)), whereas the peak with a higher binding energy (855.5–856 eV) is attributed to the NiO interacting with CaO (Ni2+ (II)) [29,30,31]. The intensity ratio of Ni2+ (I) and Ni2+ (II) increased as the Ni concentration increased from 2 to 20 wt.%. This is because for catal-sorbents with larger Ni crystallite sizes (e.g., 20Ni/CaO), there may be a greater increase in the amount of bulk NiO relative to the amount of NiO interacting with CaO. In the XRD results, very small peaks of Ni0 were observed for 20Ni/CaO. The H2-TPR profiles of Ni/CaO catal-sorbents (Figure S5) revealed that NiO phases can be converted to metallic Ni under direct methanation conditions. The H2-TPR curves for 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO exhibit two groups of peaks at maximum temperatures of ~380 °C and 450–500 °C. The lower temperature event is attributed to the bulk reduction of NiO, indicating minimal or no interaction with the CaO phase, whereas the higher temperature event corresponds to NiO interacting strongly with CaO [31].

2.2. Evaluation of Catal-Sorbent Performance

CO2 capture and direct methanation in the ICCU process during five cycles under 10 vol.% CO2 and 10 vol.% H2O at 500 °C using the catal-sorbents and the corresponding profiles are depicted in Figure 3. The breakthrough curves of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO are shown in Figure 3a. The breakthrough time slightly increased with increasing Ni-to-CaO ratio because of the enhanced macroporous structures and the smaller size of CaO, despite the decrease in CaO amount. In the initial cycle, the total CO2 capture capacities of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO, which were calculated from the breakthrough curves, were 5.89, 8.96, and 9.87 mmol CO2/g catal-sorbent, respectively (Table 3). 20Ni/CaO showed the highest CO2 capture capacity owing to the smallest crystallite size of CaO and improved macroporous structure. Figure 3b shows the CH4 concentration for 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO during direct methanation. Further, it was observed that the maximum CH4 concentration in the direct methanation step decreased with increasing Ni-to-CaO ratio. However, 20Ni/CaO exhibited the highest CH4 productivity at the initial cycle as it has the highest CO2 capture capacity, and all CO2 was converted to CH4 (i.e., 100% CH4 selectivity) during direct methanation (Table 3).
The catalytic activity of the catal-sorbents in the direct methanation step was quantified by calculating the apparent turnover frequency (TOF), as shown below [32]:
TOF   ( s 1 ) = r C H 4 · M W · D · x
where r C H 4 denotes the CH4 production rate during direct methanation (mol/s), M is the atomic mass of Ni (g/mol), W indicates the weight of the catalyst packed into the reactor (g), D denotes the dispersion of Ni particles on the surface of the catalyst, and x denotes the loading weight ratio of Ni in the catalyst. Figure 3c shows the apparent TOF of the catal-sorbents obtained at the highest CH4 concentration during direct methanation in the initial cycle. The TOF increased from 0.1707 s−1 for 2Ni/CaO to 0.4621 s−1 for 10Ni/CaO and then decreased to 0.2232 s−1 for 20Ni/CaO. Figure 3d shows the CH4 productivity over five cycles of ICCU methanation steps for the catal-sorbents. The CO2 capture capacities of 2Ni/CaO and 20Ni/CaO catal-sorbents decreased gradually over multiple cycles, resulting in a gradual decline in CH4 productivity at the subsequent direct methanation step (Figure S6). Further, it was observed that the breakthrough time of 2Ni/CaO slightly decreased as the cycle test progressed. Nevertheless, the kinetics of direct methanation in 2Ni/CaO were stable throughout the cycles. In contrast to 2Ni/CaO, the breakthrough time of 20Ni/CaO was stable. However, the CO2 capture capacity decreased in subsequent reaction cycles because of the incomplete regeneration of CaCO3 by slow direct methanation. 20Ni/CaO exhibited a relatively sluggish direct methanation rate, and the rate was decreasing as the number of reaction cycles increased, leading to a decrease in CH4 productivity from 9.72 to 6.97 mmol CH4/g catal-sorbent. 10Ni/CaO exhibited excellent thermal stability with high CH4 productivity (8.95–8.75 mmol CH4/g catal-sorbent) over five cycles of ICCU methanation because of its high CO2 capture capacity, fast carbonation, and direct methanation by the improved macroporous CaO structure and Ni dispersion. The CaO phase in the catal-sorbents is converted to CaCO3 by CO2 absorption after the first carbonation at 500 °C for 2 h, and the CaO conversion ratio increases with increasing Ni-to-CaO ratio (Figure 4a), corresponding to the CO2 capture performance presented in Figure 3a. As shown in Figure 1b, NiO is reduced to Ni, and CaCO3 is regenerated to CaO/Ca(OH)2 phases during direct methanation. However, for 20Ni/CaO, not all CaCO3 is regenerated after the fifth direct methanation because of the slow methanation rate, as mentioned above. After the sixth CO2 capture, the Ni phase is converted back to the NiO phase by the oxidation of metallic Ni by CO2 or H2O in the CO2 capture step [28,29]. It is observed that the CaO conversion ratio in 2Ni/CaO decreased, compared to the first CO2 capture. This might be because 2Ni/CaO has the largest CaO crystallite size and the poorest macroporous structures among the catal-sorbents, and CO2 cannot diffuse through the thick layer of CaCO3 to come in contact with the inner CaO phase. For 10Ni/CaO, the conversion of CaO to CaCO3 increased, which may be due to the redispersion of the CaO phase. However, for 20Ni/CaO, the increase in CaO conversion after the sixth CO2 capture is caused by incomplete decarbonation during direct methanation.
After five consecutive cycles of CO2 capture and direct methanation, the crystallite size of Ni species in 2Ni/CaO and 10Ni/CaO increased slightly from 20.2 and 26.9 to 32.3 and 37.8 nm, respectively (Figure 5). Despite the increase in the crystallite size of Ni in 2Ni/CaO and 10Ni/CaO catal-sorbents, the kinetics of direct methanation maintained stability because of the relatively high Ni dispersion. 20Ni/CaO shows high bulk NiO (Ni2+ (I)) in the XPS results, which explains the large Ni particle size. The large size of the Ni particles means that the dispersion of Ni is low. This exhibited a sluggish direct methanation rate (Figure S6). 20Ni/CaO has a higher amount of bulk NiO compared to the amount of NiO interacting with CaO (Ni2+ (II)) due to the large particle size of Ni. Many bulk NiO produced rapid sintering of Ni species after five consecutive cycles in CO2 capture and direct methanation (Figure 5f). The increase in the Ni crystallite size correlates with the gradual deactivation of CH4 productivity as the cycle progresses.

3. Materials and Methods

3.1. Catal-Sorbent Synthesis

Nickel (II) nitrate hexahydrate (Ni(NO3)2·6H2O) and calcium nitrate tetrahydrate (Ca(NO3)2·4H2O) were used as Ni and Ca precursors, respectively, and citric acid was used as a chelating agent. The 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents were prepared by citrate sol–gel techniques [28,29]. Metal precursors and citric acid were dissolved in deionized water, and the solution was vigorously stirred at 50 °C for 1 h. Then, the solution was dried overnight in an oven at 100 °C to obtain a gel. The resulting dry gel was finely ground and calcined in a furnace at 800 °C for 5 h (temperature ramp rate of 10 °C/min). The resulting catal-sorbents were designated as xNi/CaO, where x represents the weight percentage of the metallic Ni based on the catal-sorbents.

3.2. Catal-Sorbent Characterization

The metal contents (Ni and Ca) of the catal-sorbents were determined using inductively coupled plasma–optical emission spectrometry (ICP-OES, OPTIMA 4300 DV, PerkinElmer, Waltham, MA, USA). Textural properties such as the Brunauer, Emmett, and Teller (BET) surface area, pore volume, average pore size, and nanoparticle size were estimated from the N2 adsorption–desorption isotherms at −196 °C using Micromeritics ASAP 2020 (Norcross, GA, USA) analyzer. The Ni dispersion and surface area were determined by a CO chemisorption experiment using a Micromeritics ASAP 2020 analyzer. The average pore size was calculated by the Barrett–Joyner–Halenda method, and macroporosity and pore size distribution were measured by a Hg porosimeter (AutoPore V 9620, Micromeritics, Norcross, GA, USA). The morphology of the catal-sorbents was determined through field emission scanning electron microscopy (FE-SEM, S-4800, Hitachi, Tokyo, Japan), and the crystalline phase was determined through field emission transmission electron microscopy (FE-TEM, JEM-2100F, JEOL, Tokyo, Japan). The chemical composition of the surface of the catal-sorbents was obtained using an X-ray photoelectron spectrometer (XPS) using a NEXSA (ThermoFisher, Waltham, MA, USA) equipped with an Al Kα source. Differential thermal gravity curves on an SDT Q600 thermogravimetric analyzer were used to evaluate the reduction properties. The crystallographic structure of the catal-sorbents was determined by X-ray diffraction (XRD) analysis (Phillips X’PERT X-ray diffractometer, PANalytical, Amsterdam, Netherlands) using a Cu Kα radiation source. The crystallite size of CaO in the catal-sorbents was calculated by the Scherrer equation. In addition, thermogravimetric analysis (TGA, N-1000, Scinco, Seoul, Korea) was utilized to assess the cyclic regeneration properties of the catal-sorbents during CO2 capture and direct methanation (CO2 capture: pure CO2 and direct methanation: pure H2 at 500 °C).

3.3. Reaction Process

The CO2 capture and direct methanation in the ICCU process was carried out using 0.5 g of catal-sorbents in a packed-bed reactor (inner diameter of ½ inch). The reactor was placed in an electric furnace at atmospheric pressure. The reaction temperatures during CO2 capture and direct methanation were maintained at 500 °C. At the CO2 capture step, 10 vol.% CO2, 10 vol.% H2O, and balance N2 were made to flow through the packed bed with 40 mL/min of total flow rate. At the direct methanation step, 90 vol.% H2 and balance N2 were passed after purging the reactor. N2 was used as the balancing and standard gas. The outlet gases were analyzed through gas chromatography (GC, Agilent 6890, Agilent Technologies, Santa Clara, CA, USA) equipped with two thermal conductivity detectors (TCD). A Porapak Q packed column was used to separate N2, CO2, and H2O during CO2 capture, whereas a Carboxen-1000 packed column was used to separate H2, N2, CH4, and CO2 during direct methanation step.

4. Conclusions

Herein, Ni/CaO catal-sorbents were prepared by the sol–gel method to determine the effect of the Ni-to-CaO ratio on CO2 capture and direct methanation performances. With increasing Ni-to-CaO ratio, the macroporous structures and crystallite size of Ni species increased, whereas the extent of nanoparticle agglomeration and the crystallite size of CaO decreased. The CO2 capture breakthrough time slightly increased with increasing Ni-to-CaO ratio because of the enhanced macroporous structures and the smaller size of CaO. In the direct methanation step, 20Ni/CaO showed the lowest methane production due to the large Ni particle size. 10Ni/CaO showed the best TOF results (0.4621 s−1) in the initial cycle with appropriate Ni dispersion and macroporous structure. The CO2 capture and CH4 production performances (capacity and kinetics) in 2Ni/CaO and 20Ni/CaO catal-sorbents decreased gradually over five consecutive cycles. After five consecutive cycles of CO2 capture and direct methanation, despite the increase in the crystallite size of Ni species in 2Ni/CaO and 10Ni/CaO catal-sorbents, the kinetics of direct methanation maintained stability because of the relatively high Ni dispersion. 10Ni/CaO exhibited excellent thermal stability with high CH4 productivity over five cycles in ICCU methanation because of its high CO2 capture capacity, fast carbonation, and direct methanation by the improved macroporous CaO structure and Ni dispersion. Overall, the 10Ni/CaO catal-sorbent showed promise for cost-effective CO2 capture and conversion value-added products. In the future study, long-term stability tests will be conducted to study deactivation and its effect on catalyst performance.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal13081174/s1, Figure S1. Pore size distribution from Hg porosimetry curves of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents; Figure S2. TEM-EDS mapping of the 2Ni/CaO catal-sorbent; Figure S3. TEM-EDS mapping of the 10Ni/CaO catal-sorbent; Figure S4. TEM-EDS mapping of the 20Ni/CaO catal-sorbent; Figure S5. H2-TPR profiles of the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents under pure H2 from 100 °C to 800 °C; Figure S6. Breakthrough curves of the (a) 2Ni/CaO, (b) 10Ni/CaO, and (c) 20Ni/CaO catal-sorbents over five consecutive cycles at 500 °C.

Author Contributions

Conceptualization, J.-H.W. and S.J.; formal analysis, J.-E.K., T.-Y.K., H.-D.S., H.-J.R. and B.H.; investigation, J.-H.W., S.J., J.-E.K., T.-Y.K. and H.-D.S.; supervision, S.-C.L., J.-C.K. and K.L.G.-A.; writing—original draft, J.-H.W. and S.J.; writing—review and editing, J.-H.W. and S.-C.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by National R&D Program through the National Research Foundation of Korea (NRF) funded by Ministry of Science and ICT (No.2022M3J2A1085551).

Data Availability Statement

The data presented in this study are available on request from the corresponding author. Data are contained within the article or Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Guo, Z.-L.; Bian, X.-L.; Du, Y.-B.; Zhang, W.-C.; Yao, D.-D.; Yang, H.-P. Recent advances in integrated carbon dioxide capture and methanation technology. J. Fuel Chem. Technol. 2023, 51, 293–302. [Google Scholar] [CrossRef]
  2. Yang, N.; Kang, F.; Liu, Z.; Ge, X.; Zhou, Y. An integrated CCU-plant scheme and assessment for conversion of captured CO2 into methanol. Int. J. Low-Carbon Technol. 2022, 17, 550–562. [Google Scholar] [CrossRef]
  3. Ho, H.-J.; Iizuka, A.; Shibata, E. Carbon capture and utilization technology without carbon dioxide purification and pressurization: A review on its necessity and available technologies. Ind. Eng. Chem. Res. 2019, 58, 8941–8954. [Google Scholar] [CrossRef]
  4. Fernández, J.; Sotenko, M.; Derevschikov, V.; Lysikov, A.; Rebrov, E.V. A radiofrequency heated reactor system for post-combustion carbon capture. Chem. Eng. Process. 2016, 108, 17–26. [Google Scholar] [CrossRef]
  5. Chronopoulos, T.; Fernandez-Diez, Y.; Maroto-Valer, M.M.; Ocone, R.; Reay, D.A. CO2 desorption via microwave heating for post-combustion carbon capture. Microporous Mesoporous Mat. 2014, 197, 288–290. [Google Scholar] [CrossRef]
  6. Varvoutis, G.; Lykaki, M.; Stefa, S.; Binas, V.; Marnellos, G.E.; Konsolakis, M. Deciphering the role of Ni particle size and nickel-ceria interfacial perimeter in the low-temperature CO2 methanation reaction over remarkably active Ni/CeO2 nanorods. Appl. Catal. B-Environ. 2021, 297, 120401. [Google Scholar] [CrossRef]
  7. Rezaei Sadr, N.; Bahrdo, T.; Taghizadeh, R. Impacts of Paris agreement, fossil fuel consumption, and net energy imports on CO2 emissions: A panel data approach for three West European countries. Clean Technol. Environ. Policy 2022, 24, 1521–1534. [Google Scholar] [CrossRef]
  8. Wang, G.; Guo, Y.; Yu, J.; Liu, F.; Sun, J.; Wang, X.; Wang, T.; Zhao, C. Ni-CaO dual function materials prepared by different synthetic modes for integrated CO2 capture and conversion. Chem. Eng. J. 2022, 428, 132110. [Google Scholar] [CrossRef]
  9. Chen, J.; Xu, Y.; Liao, P.; Wang, H.; Zhou, H. Recent progress in integrated CO2 capture and conversion process using dual function materials: A state-of-the-art review. Carbon Capture Sci. Technol. 2022, 4, 100052. [Google Scholar] [CrossRef]
  10. Zhao, Y.; Li, Y.; Jin, B.; Liang, Z. Layered double hydroxide derived bifunctional Ca-Fe-Mg material for integrated CO2 capture and utilization via chemical looping strategy. Chem. Eng. J. 2022, 431, 133826. [Google Scholar] [CrossRef]
  11. Al-Mamoori, A.; Lawson, S.; Rownaghi, A.A.; Rezaei, F. Oxidative dehydrogenation of ethane to ethylene in an integrated CO2 capture-utilization process. Appl. Catal. B-Environ. 2020, 278, 119329. [Google Scholar] [CrossRef]
  12. Lawson, S.; Baamran, K.; Newport, K.; Alghamadi, T.; Jacobs, G.; Rezaei, F.; Rownaghi, A.A. Integrated direct air capture and oxidative dehydrogenation of propane with CO2 at isothermal conditions. Appl. Catal. B-Environ. 2022, 303, 120907. [Google Scholar] [CrossRef]
  13. Sasayama, T.; Kosaka, F.; Liu, Y.; Yamaguchi, T.; Chen, S.-Y.; Mochizuki, T.; Urakawa, A.; Kuramoto, K. Integrated CO2 capture and selective conversion to syngas using transition-metal-free Na/Al2O3 dual-function material. J. CO2 Util. 2022, 60, 102049. [Google Scholar] [CrossRef]
  14. Sun, S.; Sun, H.; Guan, S.; Xu, S.; Wu, C. Integrated CO2 capture and methanation on Ru/CeO2-MgO combined materials: Morphology effect from CeO2 support. Fuel 2022, 317, 123420. [Google Scholar] [CrossRef]
  15. Gu, H.; Gao, Y.; Iftikhar, S.; Li, F. Ce stabilized Ni–SrO as a catalytic phase transition sorbent for integrated CO2 capture and CH4 reforming. J. Mater. Chem. A 2022, 10, 3077–3085. [Google Scholar] [CrossRef]
  16. Guo, Y.; Wang, G.; Yu, J.; Huang, P.; Sun, J.; Wang, R.; Wang, T.; Zhao, C. Tailoring the performance of Ni-CaO dual function materials for integrated CO2 capture and conversion by doping transition metal oxides. Sep. Purif. Technol. 2023, 305, 122455. [Google Scholar] [CrossRef]
  17. Sun, Z.; Shao, B.; Zhang, Y.; Gao, Z.; Wang, M.; Liu, H.; Hu, J. Integrated CO2 capture and methanation from the intermediate-temperature flue gas on dual functional hybrids of AMS/CaMgO|| NixCoy. Sep. Purif. Technol. 2023, 307, 122680. [Google Scholar] [CrossRef]
  18. Jo, S.; Cruz, L.; Shah, S.; Wasantwisut, S.; Phan, A.; Gilliard-AbdulAziz, K.L. Perspective on Sorption Enhanced Bifunctional Catalysts to Produce Hydrocarbons. ACS Catal. 2022, 12, 7486–7510. [Google Scholar] [CrossRef]
  19. Jo, S.; Son, H.D.; Kim, T.-Y.; Woo, J.H.; Kim, J.C.; Lee, S.C.; Gilliard-AbdulAziz, K.L. Ru/K2CO3–MgO catalytic sorbent for integrated CO2 capture and methanation at low temperatures. Chem. Eng. J. 2023, 469, 143772. [Google Scholar] [CrossRef]
  20. Jo, S.; Lee, J.H.; Kim, T.Y.; Woo, J.H.; Ryu, H.-J.; Hwang, B.; Lee, S.C.; Kim, J.C.; Gilliard-AbdulAziz, K.L. A fundamental study of CO2 capture and CH4 production in a rapid cyclic system using nickel-lithium-silicate as a catal-sorbent. Fuel 2022, 311, 122602. [Google Scholar] [CrossRef]
  21. Jo, S.; Lee, J.H.; Woo, J.H.; Kim, T.-Y.; Ryu, H.-J.; Hwang, B.; Kim, J.C.; Lee, S.C.; Gilliard-AbdulAziz, K.L. Coke-promoted Ni/CaO catal-sorbents in the production of cyclic CO and syngas. Sustain. Energ. Fuels 2022, 6, 81–88. [Google Scholar] [CrossRef]
  22. Wei, S.; Han, R.; Su, Y.; Gao, J.; Zhao, G.; Qin, Y. Pore structure modified CaO-based sorbents with different sized templates for CO2 capture. Energy Fuels 2019, 33, 5398–5407. [Google Scholar] [CrossRef]
  23. Yoon, H.J.; Lee, K.B. Introduction of chemically bonded zirconium oxide in CaO-based high-temperature CO2 sorbents for enhanced cyclic sorption. Chem. Eng. J. 2019, 355, 850–857. [Google Scholar] [CrossRef]
  24. Zhang, C.; Li, Y.; He, Z.; Zhao, J.; Wang, D. Microtubular Fe/Mn-promoted CaO-Ca12Al14O33 bi-functional material for H2 production from sorption enhanced water gas shift. Appl. Catal. B-Environ. 2022, 314, 121474. [Google Scholar] [CrossRef]
  25. Frontera, P.; Macario, A.; Ferraro, M.; Antonucci, P. Supported catalysts for CO2 methanation: A review. Catalysts 2017, 7, 59. [Google Scholar] [CrossRef]
  26. Tsiotsias, A.I.; Charisiou, N.D.; Yentekakis, I.V.; Goula, M.A. The role of alkali and alkaline earth metals in the CO2 methanation reaction and the combined capture and methanation of CO2. Catalysts 2020, 10, 812. [Google Scholar] [CrossRef]
  27. Wang, S.; Schrunk, E.T.; Mahajan, H.; Farrauto, R.J. The role of ruthenium in CO2 capture and catalytic conversion to fuel by dual function materials (DFM). Catalysts 2017, 7, 88. [Google Scholar] [CrossRef] [Green Version]
  28. Jo, S.B.; Woo, J.H.; Lee, J.H.; Kim, T.Y.; Kang, H.I.; Lee, S.C.; Kim, J.C. CO2 green technologies in CO2 capture and direct utilization processes: Methanation, reverse water-gas shift, and dry reforming of methane. Sustain. Energ. Fuels 2020, 4, 5543–5549. [Google Scholar] [CrossRef]
  29. Jo, S.B.; Woo, J.H.; Lee, J.H.; Kim, T.Y.; Kang, H.I.; Lee, S.C.; Kim, J.C. A novel integrated CO2 capture and direct methanation process using Ni/CaO catal-sorbents. Sustain. Energ. Fuels 2020, 4, 4679–4687. [Google Scholar]
  30. Hu, J.; Hongmanorom, P.; Galvita, V.V.; Li, Z.; Kawi, S. Bifunctional Ni-Ca based material for integrated CO2 capture and conversion via calcium-looping dry reforming. Appl. Catal. B-Environ. 2021, 284, 119734. [Google Scholar] [CrossRef]
  31. Tian, S.; Yan, F.; Zhang, Z.; Jiang, J. Calcium-looping reforming of methane realizes in situ CO2 utilization with improved energy efficiency. Sci. Adv. 2019, 5, eaav5077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Lim, H.S.; Kim, G.; Kim, Y.; Lee, M.; Kang, D.; Lee, H.; Lee, J.W. Ni-exsolved La1-xCaxNiO3 perovskites for improving CO2 methanation. Chem. Eng. J. 2021, 412, 127557. [Google Scholar] [CrossRef]
Figure 1. SEM images, TEM images, and Ni crystallite size of (a,d,g) 2Ni/CaO, (b,e,h) 10Ni/CaO, and (c,f,i) 20Ni/CaO catal-sorbents prepared by the citrate sol–gel method.
Figure 1. SEM images, TEM images, and Ni crystallite size of (a,d,g) 2Ni/CaO, (b,e,h) 10Ni/CaO, and (c,f,i) 20Ni/CaO catal-sorbents prepared by the citrate sol–gel method.
Catalysts 13 01174 g001
Figure 2. (a) XRD patterns of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents in fresh states: () NiO and () Ni0, () CaO and () Ca(OH)2. (b) XPS spectra of Ni 2p3/2 in the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents.
Figure 2. (a) XRD patterns of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents in fresh states: () NiO and () Ni0, () CaO and () Ca(OH)2. (b) XPS spectra of Ni 2p3/2 in the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents.
Catalysts 13 01174 g002
Figure 3. CO2 capture and direct methanation capacity and behavior of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal−sorbents (a,b) in the initial cycle, (c) apparent TOF of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents in the initial cycle, and (d) over five consecutive cycles. (CO2 capture: 10 vol.% CO2 and 10 vol.% H2O; direct methanation: 90 vol.% H2 at 500 °C).
Figure 3. CO2 capture and direct methanation capacity and behavior of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal−sorbents (a,b) in the initial cycle, (c) apparent TOF of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents in the initial cycle, and (d) over five consecutive cycles. (CO2 capture: 10 vol.% CO2 and 10 vol.% H2O; direct methanation: 90 vol.% H2 at 500 °C).
Catalysts 13 01174 g003
Figure 4. XRD patterns of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents after (a) the first and sixth CO2 capture, and (b) fifth direct methanation in the ICCU process: () NiO and () Ni0, () CaO and () Ca(OH)2, and () CaCO3.
Figure 4. XRD patterns of 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents after (a) the first and sixth CO2 capture, and (b) fifth direct methanation in the ICCU process: () NiO and () Ni0, () CaO and () Ca(OH)2, and () CaCO3.
Catalysts 13 01174 g004
Figure 5. TEM images and Ni crystallite size of (a,d) 2Ni/CaO, (b,e) 10Ni/CaO, and (c,f) 20Ni/CaO catal-sorbents after five cycles in CO2 capture and direct methanation in the ICCU process.
Figure 5. TEM images and Ni crystallite size of (a,d) 2Ni/CaO, (b,e) 10Ni/CaO, and (c,f) 20Ni/CaO catal-sorbents after five cycles in CO2 capture and direct methanation in the ICCU process.
Catalysts 13 01174 g005
Table 1. Metal contents, BET surface area, nanoparticle size, and porosity of the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents in their fresh states.
Table 1. Metal contents, BET surface area, nanoparticle size, and porosity of the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents in their fresh states.
Metal ContentsBET SurfaceNanoparticle Size (nm)Porosity
(wt.%)Area (m2/g)(%)
2Ni/CaONi: 1.64.59130772.9
CaO: 98.4
10Ni/CaONi: 9.39.9605.874.68
CaO: 90.7
20Ni/CaONi: 18.07.77771.674.39
CaO: 82.0
Table 2. Ni crystallite size, Ni dispersion, and Ni surface area of the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents.
Table 2. Ni crystallite size, Ni dispersion, and Ni surface area of the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents.
Ni Crystallite Size a
(nm)
Dispersion b
(%)
Ni surface Area b
(m2/gNi)
2Ni/CaO20.22.4316.19
10Ni/CaO27.00.110.72
20Ni/CaO45.40.090.58
a Measured by TEM. b Calculated from CO chemisorption.
Table 3. CO2 capture and direct methanation performances of the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents over five consecutive cycles at 500 °C.
Table 3. CO2 capture and direct methanation performances of the 2Ni/CaO, 10Ni/CaO, and 20Ni/CaO catal-sorbents over five consecutive cycles at 500 °C.
CO2 CaptureDirect Methanation
CycleCO2 Captured (mmol/g)CH4 Produced (mmol/g)CH4 Selectivity (%)Carbon Balance (%)
2Ni/CaO15.895.8810099.8
25.755.75100100
35.335.33100100
44.844.8310099.8
54.454.44100100
10Ni/CaO18.968.9510099.9
28.948.9310099.9
39.249.24100100
48.708.70100100
58.758.75100100
20Ni/CaO19.879.7210098.5
28.708.6210099.1
38.167.9810097.8
47.397.2110097.6
57.156.9710097.5
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

Woo, J.-H.; Jo, S.; Kim, J.-E.; Kim, T.-Y.; Son, H.-D.; Ryu, H.-J.; Hwang, B.; Kim, J.-C.; Lee, S.-C.; Gilliard-AbdulAziz, K.L. Effect of the Ni-to-CaO Ratio on Integrated CO2 Capture and Direct Methanation. Catalysts 2023, 13, 1174. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13081174

AMA Style

Woo J-H, Jo S, Kim J-E, Kim T-Y, Son H-D, Ryu H-J, Hwang B, Kim J-C, Lee S-C, Gilliard-AbdulAziz KL. Effect of the Ni-to-CaO Ratio on Integrated CO2 Capture and Direct Methanation. Catalysts. 2023; 13(8):1174. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13081174

Chicago/Turabian Style

Woo, Jin-Hyeok, Seongbin Jo, Ju-Eon Kim, Tae-Young Kim, Han-Dong Son, Ho-Jung Ryu, Byungwook Hwang, Jae-Chang Kim, Soo-Chool Lee, and Kandis Leslie Gilliard-AbdulAziz. 2023. "Effect of the Ni-to-CaO Ratio on Integrated CO2 Capture and Direct Methanation" Catalysts 13, no. 8: 1174. https://0-doi-org.brum.beds.ac.uk/10.3390/catal13081174

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