Next Article in Journal
Optical Model and Optimization for Coherent-Incoherent Hybrid Organic Solar Cells with Nanostructures
Next Article in Special Issue
Preparation of Zeolitic Imidazolate Framework-8-Based Nanofiber Composites for Carbon Dioxide Adsorption
Previous Article in Journal
Recent Issues and Configuration Factors in Perovskite-Silicon Tandem Solar Cells towards Large Scaling Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Multicycle Performance of CaTiO3 Decorated CaO-Based CO2 Adsorbent Prepared by a Versatile Aerosol Assisted Self-Assembly Method

1
Department of Geosciences, National Taiwan University, Taipei 106, Taiwan
2
Research Center for Future Earth, National Taiwan University, Taipei 106, Taiwan
3
Department of Environmental Engineering, National Ilan University, Yilan 260, Taiwan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(12), 3188; https://doi.org/10.3390/nano11123188
Submission received: 19 October 2021 / Revised: 17 November 2021 / Accepted: 18 November 2021 / Published: 24 November 2021
(This article belongs to the Special Issue Advanced Nanomaterials for Adsorption Purposes)

Abstract

:
Calcium oxide (CaO) is a promising adsorbent to separate CO2 from flue gas. However, with cycling of carbonation/decarbonation at high temperature, the serious sintering problem causes its capture capacity to decrease dramatically. A CaTiO3-decorated CaO-based CO2 adsorbent was prepared by a continuous and simple aerosol-assisted self-assembly process in this work. Results indicated that CaTiO3 and CaO formed in the adsorbent, whereas CaO gradually showed a good crystalline structure with increased calcium loading. Owing to the high thermal stability of CaTiO3, it played a role in suppressing the sintering effect and maintaining repeated high-temperature carbonation and decarbonation processes. When the calcium and titanium ratio was 3, the CO2 capture capacity was as large as 7 mmol/g with fast kinetics. After 20 cycles under mild regeneration conditions (700 °C, N2), the performance of CO2 capture of CaTiO3-decorated CaO-based adsorbent nearly unchanged. Even after 10 cycles under severe regeneration conditions (920 °C, CO2), the performance of CO2 capture still remained nearly 70% compared to the first cycle. The addition of CaTiO3 induced good and firm CaO dispersion on its surface. Excellent kinetics and stability were evident.

Graphical Abstract

1. Introduction

The global climate-change phenomenon has become an important concern in recent years because of excessive CO2 emissions, and this situation will continue because our energy supply originates mostly from fossil-fuel combustion now and in the next few years. Therefore, capturing CO2 from flue gas and transporting it to a suitable site for storage is a solution for CO2-emission reduction [1,2,3]. Calcium is an abundant element on earth, and its oxide form, CaO, can capture CO2 and change it to the carbonate form, CaCO3, via carbonation. Then the oxide form can be regenerated back by CO2 removal through decarbonation. Therefore, CaO-based adsorbents can be used repeatedly by combining carbonation and decarbonation processes [4,5,6,7,8]. The nontoxicity and low cost of CaO-based adsorbents also enable them to become excellent potential candidate for CO2-adsorption application.
However, the sintering problem that causes CO2-capture capacity loss dramatically after few cycles remains a challenge during operation in high-temperature environments (600–900 °C) [4,9,10,11]. The process of CO2 capture by calcium oxide can be divided into two stages: the surface chemical-reaction controlled stage and the inner-diffusion-controlled stage [12,13]. The former has higher capture efficiency than the latter. The sintering effect increases particle size. Larger particle sizes result in the difficult use of inner calcium oxide because the latter stage is more important. Thus, the capture performance is reduced. Introducing additives with high thermal stability into CaO-based adsorbents has been a common strategy to mitigate sintering effects and improve cyclic stability. This strategy involves adding another metal (Zr, Al, Mg, Si, etc.) precursor during CaO-based adsorbent preparation, and then high-thermal-stability additives form independently (e.g., MgO, Al2O3, SiO2) or react with partial calcium (e.g., Ca9Al6O18, CaZrO3) as the sorbent is obtained [13,14,15,16,17,18,19], or introducing it into high-thermal-stability porous silica solid (e.g., SBA-15, KIT-6) via the impregnation process [9,20,21]. The granules form sorbents for enhancing mechanical stability, or acoustic sound-assisted technique are also considered to reduce sorbent deactivation after several repeated uses [22,23].
The incorporation of titanium additives can introduce calcium titanium oxide (CaTiO3) into CaO-based adsorbents. This introduction can be expected to alleviate the CaO-based adsorbent sintering problem, owing to its thermal stability at the working temperature of CaO-based adsorbents [24,25]. Yu et al. used the precipitation and deposition method to prepare Ca/Al/Ti sorbents. The capture capacity was reduced less than 5% after 10 cycles of the capture–regeneration experiment as small amount of titania were introduced. Wu et al. prepared CaTiO3/nano-CaO by the hydrolysis method. The authors found that the high melting point of CaTiO3 can improve the cyclic stability of CaO-based adsorbents. The sorbent almost retained its capacity after 40 repeated uses.
Our scope is to prepare well sintering-resisted CaO-based CO2 adsorbent through an easily scaled up process. In the current work, CaTiO3-decorated CaO-based CO2 adsorbent was prepared by the aerosol-assisted self-assembly method. This method can be used to quickly and continuously prepare multicomponent metal oxide [26,27]. The properties and CO2-capture performance of sorbents with different calcium and titanium mole ratios were investigated.

2. Materials and Methods

2.1. Adsorbents Preparation

In the typical preparation, 11.5 mL of acetic acid (Nacalai, Nacalai Tesque, Kyoto, Japan, 99%), 5 mL of hydrochloric acid (Nacalai, Nacalai Tesque, Kyoto, Japan, 35%), 60 mL of alcohol (Nacalai, Nacalai Tesque, Kyoto, Japan, >99.5%), calcium nitrate 4-hydrate (Ca(NO)3 · 4H2O, Macron, Avantor Inc., Radnor, PA, USA), and 6 g of F127 (Sigma-Aldrich, Sigma Ltd., Saint Louis, MO, USA) were added in 17 mL of tetrabutyl orthotitanate (Ti(OC4H9)4, Sigma-Aldrich, Sigma Ltd., Saint Louis, MO, USA, 97%) sequentially. The mole ratio of Ca(NO)3 · 4H2O to Ti(OC4H9)4 was 1.2/2/3/5. The mixture was stirring until all chemicals totally dissolved. Then, the transparent mixture solution was transformed to aerosol by nozzle first, and then passed through a 400 °C furnace. The as-prepared product was collected on filter paper. The sorbent was obtained by calcination at 700 °C for 30 min with a 1 °C/min heating rate. The received sorbent was named as CaTi-x, where x represents the mole ratio of calcium to titanium.

2.2. Adsorption Characterization

An X-ray diffraction (XRD) measurement was completed by PANalytical X’ Pert PRO (Cu Kα λ = 0.154 nm, voltage: 45 kV, current: 40 mA). A nitrogen adsorption/desorption experiment was completed by Micromeritics ASAP 2020 (degas condition: 130 °C overnight, analysis condition: liquid nitrogen environment). Surface area was calculated from the adsorption isotherm in the relative pressure range of 0.05–0.3 using BET theory. Scanning electron microscopy (SEM) was done using a JEOL JSM-6500F.

2.3. CO2 Capture Experiments

CO2 capture experiments were evaluated by the gravimetric method (TA SDT-Q600). Approximately 5 mg of sorbent was used and pretreated at 700 °C under pure nitrogen gas flowing for 0.5 h. The capture environment was set at 600 °C under pure CO2 gas flowing for 1 h. A cyclic stability experiment was done by using the gravimetric method. Each cycle combined carbonation and regeneration processes, and a pretreatment process was added before the cycle experiment started. The capture condition was set at 600 °C under pure CO2 gas flowing for 5 min, and the regeneration condition was set at 700 °C under pure N2 gas flowing for 50 min (20 cycles) and 920 °C under pure CO2 gas flowing for 15 min (10 cycles).

3. Results and Discussion

3.1. Properties of Adsorbents

The SEM images of obtained sorbents with different calcium to titanium mole ratios are shown in Figure 1. Owing to fixed titanium precursor addition amounts, the larger ratio corresponded with more calcium-precursor loading. At low calcium-precursor loading (CaTi-1.2 and CaTi-2), the morphology showed a more aggregate pattern than high calcium-precursor loading (CaTi-3 and CaTi-5).
The XRD result, as shown in Figure 2, revealed that CaTi-1.2 and CaTi-2 had diffraction signals assigned as CaTiO3 contribution (JCPDS #220153) only. With increased calcium loading, the diffraction signals’ intensities contributed by CaTiO3 decreased. The CaO diffraction signal was not observed in the two sorbents because less calcium was available to form a highly crystalline structure. CaTi-3 presented diffraction signals contributed by CaTiO3 and CaO (JCPDS #772376). The sample with higher calcium-precursor loading sorbent, CaTi-5, presented two other diffraction signals from Ca(OH)2 (JCPDS #841263) compared with CaTi-3. This result indicated that only certain amounts of calcium can form high-thermal-stability oxide with titanium, and the others formed oxide itself. The highly crystalline hydroxide form demanded more calcium to participate than did the oxide form.
The nitrogen adsorption/desorption analysis results are shown in Figure 3. All four sorbents showed surface areas of 5–17 m2/g (Table 1). Similar value indicated that calcium-precursor loading had little influence on the change in surface area. The appearance of hysteresis loops indicated a porous structure, possibly due to nanoparticle aggregation or stacking. The pore size distribution of the sorbents in this work, CaTi-2, CaTi-3, and CaTi-5, were primarily in the range from 2 to 10 nm, whereas CaTi-1.2 was mainly in the range from 8 to 22 nm (Figure S1).

3.2. CO2 Capture Performance of Adsorbents

The CO2-capture performance of calcium-based sorbents is presented in Figure 4. All sorbents showed rapidly increasing capture in less than 3 min before reaching saturation. CO2 capture by calcium oxide involved two stages: the first was a rapidly increasing capture amount controlled by chemical reaction, and the second was slowly increasing capture amount controlled by an inner-diffusion mechanism [6,13]. However, these calcium-based sorbents were mainly observed first rapidly increasing in the capture-amount stage, while the inner-diffusion stage was unclear. This result indicated that the calcium-oxide particle size in these sorbents was very small, so the calcium oxide in inner sites could be utilized easily. Wei et al. concluded that pore size located in the range from 2 to 10 nm is beneficial for the chemical reaction controlled stage [28]. Therefore, CaTi-3 and CaTi-5 more rapidly increased than the other two sorbents, which may be related to the reason that most of their pore sizes were in the range from 2 to 10 nm. The saturated capture amount increased with increased calcium-precursor loading because a higher amount of calcium oxide was present (Table 1).

3.3. Multicycle Capture and Regeneration Experiment of Adsorbents

The sorbents with the highest and second highest saturated capture amounts, CaTi-3 and CaTi-5, were selected for evaluation of their multicycle stability. The results are shown in Figure 5a,b, respectively. After 20 cycles of repeated use, the capture capacity of CaTi-3 was almost unchanged, whereas that of CaTi-5 decreased by ~17% compared with that in the first cycle (Figure 5c). CaTi-5 showed capacity loss because too much calcium oxide was present in this sorbent. The carbonation behavior of CaTi-3 and CaTi-5 at the twentieth cycle were different than the first cycle (Figure 5d). The two stage carbonation behavior due to more importance of the inner-diffusion mechanism at the twentieth cycle in comparison to the first cycle carbonation reaction revealed that slight sintering occurred in both sorbents during 20 cycles of repeated use. The sintering phenomenon occurred at the CaCO3 phase due to its low Tammann temperature (~530 °C) [13]. Thus, the inner-diffusion controlled mechanism became important as repeated times increased. This result indicated that after 20 cycles of repeated use, calcium species particle sizes increased. Additionally, the capture capacity of CaTi-3 in the twentieth-cycle carbonation was closer to its first cycle than the CaTi-5 case at 10 min, indicating that calcium species particles suffered a greater degree of sintering in CaTi-5 than in CaTi-3.
The XRD results of CaTi-3 and CaTi-5 sorbents after 1 and 20 cycles of repeated use are presented in Figure 6a,b, respectively. After 1 cycle, CaTiO3 and Ca(OH)2 diffraction signals could be found in both sorbents. However, several CaCO3 diffraction signals (JCPDS 850849) clearly appeared in CaTi-5 after 20 cycles of repeated use. Given that the cycle experiment involved an initial carbonation part followed by regeneration, the CaCO3 species remaining after the cyclic experiment indicated that decarbonation efficiency decreased. The CaCO3 species could not capture CO2, so CaTi-5-capture capacity loss occurred after 20 cycles. Moreover, one CaCO3 diffraction signal with a very weak peak intensity was found in CaTi-3 after 20 cycles, indicating that cyclic stability started to decrease when the mole ratio of calcium to titanium was greater than 3. The SEM images of CaTi-3 and CaTi-5 after 20 cycles are shown in Figure 7. CaTi-5 presented a denser morphology than CaTi-3, indicating a lower carbonation performance of CaTi-5 compared to CaTi-3, which was consistent with the cyclic-experiment results.
CaTi-3 was selected for further evaluation of its multicycle stability at high temperatures under CO2 flowing regeneration conditions because of its superior cyclic performance (Figure 8). The results showed significant capacity loss at the third cycle, probably due to microstructure change, followed by a gradual increase in capacity, which may be related to structure optimization. After the seventh cycle, the capacity became unchanged and exhibited about a 32% decrease compared with the first cycle. The weight loss during regeneration was similar to weight gain during carbonation in each cycle. Thus, the capture capacity decreased after several times of repeated use, which may be related to low carbonation efficiency. Hence, CaO was not completely utilized at the end of carbonation. The XRD results of CaTi-3 after 1 and 10 cycles of repeated use showed diffraction signals attributed to CaTiO3, CaCO3, and CaO species (Figure 9). After one cycle, the diffraction signals due to CaTiO3 and CaO species should have appeared at the end of regeneration. The diffraction signals assigned to CaCO3 species was due to the reaction of some CaO with CO2 as the experiment finished. The same situation occurred after 10 cycles. The SEM images of CaTi-3 showed more aggregated and fused morphology with increasing repetition times (Figure 10).
The different types of behavior of the CaTi-3 sorbent cyclic experiment under N2 flowing at low temperatures (700 °C for 30 min, mild condition) or under CO2 flowing at high temperatures (920 °C for 15 min, severe condition) may be related to calcium formation on the support surface based on the XRD results of sorbents collected after cyclic experiment (Figure 11). Under mild regeneration conditions, the cyclic experiment process mainly involves Ca(OH)2 and CaCO3 species transformation. Ca(OH)2 is under the molten state (melting point ~512 °C), while CaCO3 is in the solid state (melting point ~825 °C) during carbonation and regeneration. During carbonation, the ions diffuse in molten Ca(OH)2 layers more efficiently; thus, almost all calcium species can be utilized. When CaCO3 forms, it will precipitate due to its larger density (2.71 cm3/g) than Ca(OH)2 (2.21 cm3/g). During regeneration, the molten Ca(OH)2 layer permits the easy regeneration of CaCO3 located in the interior position. Therefore, nearly no capture capacity decay occurs. Under severe regeneration conditions, the cyclic experiment process mainly involves CaO and CaCO3 species transformation. CaCO3 is in the solid state during carbonation and under the molten state during regeneration; furthermore, CaO is in the solid state, and sintering cannot occur during carbonation and regeneration (melting point ~2613 °C, Tammann temperature ~1154 °C) [29]. During carbonation, external CaO can react with CO2 to form CaCO3, whereas CO32− and O2− exchange occurs at the internal CaCO3/CaO interface [4]. During regeneration, CaO forms and precipitates in molten CaCO3 (density of CaO is 3.34 cm3/g). The microstructure stacked by these aggregated precipitates will affect subsequent cycle carbonation reaction performance. More precipitants with densely stacked structures reduce the carbonation efficiency.

4. Conclusions

We used a versatile aerosol-assisted self-assembly method to synthesize homogeneous CO2 adsorbent with calcium and titanium. The calcium titanium oxide species (CaTiO3) dispersed active species (CaO) well and mitigated the sintering effect during high-temperature multicycle carbonation and decarbonation. CO2-capture capacity increased with increased mole ratios of calcium and titanium. With a ratio equal to 3, the CO2-capture capacity became as large as 7 mmol/g with excellent reaction kinetics. Even after 20 cycles under mild regeneration conditions, the CO2-capture performance of CaTiO3-decorated CaO-based adsorbent was nearly retained with morphologies and structures unchanged. Moreover, after 10 cycles under severe regeneration conditions, the capture capacity still retained nearly 70% of the first cycle.

Supplementary Materials

The following is available online at https://www.mdpi.com/article/10.3390/nano11123188/s1, Figure S1: BJH pore size distribution result of calcium-based sorbents.

Author Contributions

Conceptualization, R.-W.C. and C.-J.L.; Data curation, R.-W.C. and C.-J.L.; Formal analysis, R.-W.C. and C.-J.L.; Investigation, R.-W.C.; Methodology, R.-W.C. and Y.-H.L.; Writing—original draft, R.-W.C.; Writing—review and editing, C.-J.L. and Y.-H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Ministry of Science and Technology, Taiwan, grant number 109-2116-M-002-035.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Krödel, M.; Landuyt, A.; Abdala, P.M.; Müller, C.R. Mechanistic understanding of CaO-based sorbents for high-temperature CO2 capture: Advanced characterization and prospects. ChemSusChem 2020, 13, 6259–6272. [Google Scholar] [PubMed]
  2. Sin, G.; Lee, J.; Karakoti, A.; Bahadur, R.; Yi, J.; Zhao, D.; AlBahily, K.; Vinu, A. Emerging trends in porous materials for CO2 capture and conversion. Chem. Soc. Rev. 2020, 49, 4360–4404. [Google Scholar]
  3. Yang, N.; Xue, R.; Huang, G.; Ma, Y.; Wang, J. CO2 adsorption performance and kinetics of ionic liquid-modified calcined magnesite. Nanomaterials 2021, 11, 2614. [Google Scholar] [CrossRef] [PubMed]
  4. Choi, S.; Drese, J.H.; Jones, C.W. Adsorbent materials for carbon dioxide capture from large anthropogenic point sources. ChemSusChem 2009, 2, 796–854. [Google Scholar] [CrossRef] [PubMed]
  5. Radfarnia, H.R.; Sayari, A. A highly efficient CaO-based CO2 sorbent prepared by a citrate-assisted sol–gel technique. Chem. Eng. J. 2015, 262, 913–920. [Google Scholar] [CrossRef]
  6. Hu, Y.; Jia, Q.; Shan, S.; Li, S.; Jiang, L.; Wang, Y. Development of CaO-based sorbent doped with mineral rejects–bauxite-tailings in cyclic CO2 capture. J. Taiwan Inst. Chem. Eng. 2015, 46, 155–159. [Google Scholar] [CrossRef]
  7. Wang, N.; Feng, Y.; Liu, L.; Guo, X. Effects of preparation methods on the structure and property of Al-stabilized CaO-based sorbents for CO2 capture. Fuel Process. Technol. 2018, 173, 276–284. [Google Scholar] [CrossRef]
  8. Liu, F.Q.; Li, W.H.; Liu, B.C.; Li, R.X. Synthesis, characterization, and high temperature CO2 capture of new CaO based hollow sphere sorbents. J. Mater. Chem. A 2013, 1, 8037–8044. [Google Scholar] [CrossRef]
  9. Sun, H.; Parlett, C.M.A.; Isaacs, M.A.; Liu, X.; Adwek, G.; Wang, J.; Shen, B.; Huang, J.; Wu, C. Development of Ca/KIT-6 adsorbents for high temperature CO2 capture. Fuel 2019, 235, 1070–1076. [Google Scholar] [CrossRef] [Green Version]
  10. Xu, Y.; Ding, H.; Luo, C.; Zheng, Y.; Zhang, Q.; Li, X.; Sun, J.; Zhang, L. Potential synergy of chlorine and potassium and sodium elements in carbonation enhancement of CaO-based sorbents. ACS Sustain. Chem. Eng. 2018, 6, 11677–11684. [Google Scholar] [CrossRef]
  11. Sun, P.; Grace, J.R.; Lim, C.J.; Anthony, E.J. The effect of CaO sintering on cyclic CO2 capture in energy systems. AIChE J. 2007, 53, 2432–2442. [Google Scholar] [CrossRef]
  12. Kierzkowska, A.M.; Pacciani, R.; Muller, C.R. CaO-based CO2 sorbents: From fundamentals to the development of new, highly effective materials. ChemSusChem 2013, 6, 1130–1148. [Google Scholar] [CrossRef] [PubMed]
  13. Naeem, M.A.; Armutlulu, A.; Imtiaz, Q.; Donat, F.; Schäublin, R.; Kierzkowska, A.; Müller, C.R. Optimization of the structural characteristics of CaO and its effective stabilization yield high-capacity CO2 sorbents. Nat. Commun. 2018, 9, 2408–2418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Nityashree, N.; Manohara, G.V.; Maroto-Valer, M.M.; Garcia, S. Advanced high-temperature CO2 sorbents with improved long-term cycling stability. ACS Appl. Mater. Interfaces 2020, 12, 33765–33774. [Google Scholar] [CrossRef]
  15. Sun, J.; Guo, Y.; Yang, Y.; Li, W.; Zhou, Y.; Zhang, J.; Liu, W.; Zhao, C. Mode investigation of CO2 sorption enhancement for titanium dioxide decorated CaO-based pellets. Fuel 2019, 256, 116009–116017. [Google Scholar] [CrossRef]
  16. Radfarnia, H.R.; Iliuta, M.C. Metal oxide-stabilized calcium oxide CO2 sorbent for multicycle operation. Chem. Eng. J. 2013, 232, 280–289. [Google Scholar] [CrossRef]
  17. Guo, H.; Kou, X.; Zhao, Y.; Wang, S.; Sun, Q.; Ma, X. Effect of synergistic interaction between Ce and Mn on the CO2 capture of calcium-based sorbent: Textural properties, electron donation, and oxygen vacancy. Chem. Eng. J. 2018, 334, 237–246. [Google Scholar] [CrossRef]
  18. Huang, C.H.; Chang, K.P.; Yu, C.T.; Chiang, P.C.; Wang, C.F. Development of high-temperature CO2 sorbents made of CaO-based mesoporous silica. Chem. Eng. J. 2010, 161, 129–135. [Google Scholar] [CrossRef]
  19. Valverde, J.M.; Pontiga, F.; Soria-Hoyo, C.; Quintanilla, M.A.S.; Moreno, H.; Duran, F.J.; Espin, M.J. Improving the gas–solids contact efficiency in a fluidized bed of CO2 adsorbent fine particles. Phys. Chem. Chem. Phys. 2011, 13, 14906–14909. [Google Scholar] [CrossRef]
  20. Perez-Vaquero, J.; Valverde, J.M.; Quintanilla, M.A.S. Flow properties of CO2 sorbent powders modified with nanosilica. Powder Technol. 2013, 249, 443–455. [Google Scholar] [CrossRef]
  21. Gunathilake, C.; Jaroniec, M. Mesoporous calcium oxide–silica and magnesium oxide–silica composites for CO2 capture at ambient and elevated temperatures. J. Mater. Chem. A 2016, 4, 10914–10924. [Google Scholar] [CrossRef]
  22. Ammendola, P.; Raganati, F.; Miccio, F.; Murri, A.N.; Landi, E. Insights into utilization of strontium carbonate for thermochemical energy storage. Renew. Energy 2020, 157, 769–781. [Google Scholar] [CrossRef]
  23. Raganati, F.; Chirone, R.; Ammendola, P. Calcium-looping for thermochemical energy storage in concentrating solar power applications: Evaluation of the effect of acoustic perturbation on the fluidized bed carbonation. Chem. Eng. J. 2020, 392, 123658–123668. [Google Scholar] [CrossRef]
  24. Yu, C.; Kuo, H.; Chen, Y. Carbon dioxide removal using calcium aluminate carbonates on titanic oxide under warm-gas conditions. Appl. Energy 2016, 162, 1122–1130. [Google Scholar] [CrossRef]
  25. Wu, S.F.; Zhu, Y.Q. Behavior of CaTiO3/nano-CaO as a CO2 reactive adsorbent. Ind. Eng. Chem. Res. 2010, 49, 2701–2706. [Google Scholar] [CrossRef]
  26. Lin, C.J.; Yang, W.T. Ordered mesostructured Cu-doped TiO2 spheres as active visible-light-driven photocatalysts for degradation of paracetamol. Chem. Eng. J. 2014, 237, 131–137. [Google Scholar] [CrossRef]
  27. Fan, J.; Boettcher, S.W.; Stucky, G.D. Nanoparticle assembly of ordered multicomponent mesostructured metal oxides via a versatile sol-gel process. Chem. Mater. 2006, 18, 6391–6396. [Google Scholar] [CrossRef]
  28. 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]
  29. Rodaev, V.V.; Razlivalova, S.S. Performance and durability of the Zr-doped CaO sorbent under cyclic carbonation–decarbonation at different operating parameters. Energies 2021, 14, 4822. [Google Scholar] [CrossRef]
Figure 1. SEM images of as-prepared calcium-based sorbents: (a) CaTi-1.2, (b) CaTi-2, (c) CaTi-3, (d) CaTi-5.
Figure 1. SEM images of as-prepared calcium-based sorbents: (a) CaTi-1.2, (b) CaTi-2, (c) CaTi-3, (d) CaTi-5.
Nanomaterials 11 03188 g001
Figure 2. XRD result of as-prepared calcium-based sorbents: (a) CaTi-1.2, (b) CaTi-2, (c) CaTi-3, (d) CaTi-5. The inverted triangle symbol, diamond symbol, and circle symbol represent CaTiO3, CaO, and Ca(OH)2, respectively.
Figure 2. XRD result of as-prepared calcium-based sorbents: (a) CaTi-1.2, (b) CaTi-2, (c) CaTi-3, (d) CaTi-5. The inverted triangle symbol, diamond symbol, and circle symbol represent CaTiO3, CaO, and Ca(OH)2, respectively.
Nanomaterials 11 03188 g002
Figure 3. Nitrogen adsorption/desorption isotherm of as-prepared calcium-based sorbents: (a) CaTi-1.2, (b) CaTi-2, (c) CaTi-3, (d) CaTi-5.
Figure 3. Nitrogen adsorption/desorption isotherm of as-prepared calcium-based sorbents: (a) CaTi-1.2, (b) CaTi-2, (c) CaTi-3, (d) CaTi-5.
Nanomaterials 11 03188 g003
Figure 4. CO2 capture performance of calcium-based sorbents: (a) CaTi-1.2, (b) CaTi-2, (c) CaTi-3, (d) CaTi-5.
Figure 4. CO2 capture performance of calcium-based sorbents: (a) CaTi-1.2, (b) CaTi-2, (c) CaTi-3, (d) CaTi-5.
Nanomaterials 11 03188 g004
Figure 5. Twenty cyclic repeating experiment results of CaTi-3 (a) and CaTi-5 (b); percentage of capture capacity of each cycle to the first cycle (c), and the carbonation performance of the first and the twentieth experiment of CaTi-3 (up) and CaTi-5 (down) (d).
Figure 5. Twenty cyclic repeating experiment results of CaTi-3 (a) and CaTi-5 (b); percentage of capture capacity of each cycle to the first cycle (c), and the carbonation performance of the first and the twentieth experiment of CaTi-3 (up) and CaTi-5 (down) (d).
Nanomaterials 11 03188 g005
Figure 6. XRD results of (a) CaTi-3 and (b) CaTi-5 collected from after 1 cycle and 20 cycle experiments, where the invert triangle symbol represents CaTiO3, the circle symbol represents Ca(OH)2, and the star symbol represents CaCO3.
Figure 6. XRD results of (a) CaTi-3 and (b) CaTi-5 collected from after 1 cycle and 20 cycle experiments, where the invert triangle symbol represents CaTiO3, the circle symbol represents Ca(OH)2, and the star symbol represents CaCO3.
Nanomaterials 11 03188 g006
Figure 7. SEM images of CaTi-3 (a) and CaTi-5 (b) after 20 cycle experiment.
Figure 7. SEM images of CaTi-3 (a) and CaTi-5 (b) after 20 cycle experiment.
Nanomaterials 11 03188 g007
Figure 8. Ten cyclic repeating experiment results of CaTi-3 (a), and the percentage of capture capacity of each cycle to the first cycle (b).
Figure 8. Ten cyclic repeating experiment results of CaTi-3 (a), and the percentage of capture capacity of each cycle to the first cycle (b).
Nanomaterials 11 03188 g008
Figure 9. XRD results of CaTi-3 collected from after 1 cycle and 10 cycles experiment. Where the invert triangle symbol represents for CaTiO3, the diamond symbol represents for CaO, the star symbol represents for CaCO3.
Figure 9. XRD results of CaTi-3 collected from after 1 cycle and 10 cycles experiment. Where the invert triangle symbol represents for CaTiO3, the diamond symbol represents for CaO, the star symbol represents for CaCO3.
Nanomaterials 11 03188 g009
Figure 10. SEM images of CaTi-3 after (a) 1 cycle, (b) 2 cycle, (c) 3 cycle, and (d) 20 cycle experiments.
Figure 10. SEM images of CaTi-3 after (a) 1 cycle, (b) 2 cycle, (c) 3 cycle, and (d) 20 cycle experiments.
Nanomaterials 11 03188 g010
Figure 11. Schematic diagram of CaTi-3 under (a) mild regeneration conditions, and (b) severe regeneration of multicycle experiment.
Figure 11. Schematic diagram of CaTi-3 under (a) mild regeneration conditions, and (b) severe regeneration of multicycle experiment.
Nanomaterials 11 03188 g011aNanomaterials 11 03188 g011b
Table 1. Surface area and CO2 capture capacity of calcium-based sorbents.
Table 1. Surface area and CO2 capture capacity of calcium-based sorbents.
SorbentSurface Area
(m2/g)
CO2 Capture Capacity
(mmole CO2/g Sorbent)
CaTi-1.250.2
CaTi-2102.6
CaTi-3177.0
CaTi-51210.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chang, R.-W.; Lin, C.-J.; Liou, Y.-H. Multicycle Performance of CaTiO3 Decorated CaO-Based CO2 Adsorbent Prepared by a Versatile Aerosol Assisted Self-Assembly Method. Nanomaterials 2021, 11, 3188. https://doi.org/10.3390/nano11123188

AMA Style

Chang R-W, Lin C-J, Liou Y-H. Multicycle Performance of CaTiO3 Decorated CaO-Based CO2 Adsorbent Prepared by a Versatile Aerosol Assisted Self-Assembly Method. Nanomaterials. 2021; 11(12):3188. https://doi.org/10.3390/nano11123188

Chicago/Turabian Style

Chang, Ren-Wei, Chin-Jung Lin, and Ya-Hsuan Liou. 2021. "Multicycle Performance of CaTiO3 Decorated CaO-Based CO2 Adsorbent Prepared by a Versatile Aerosol Assisted Self-Assembly Method" Nanomaterials 11, no. 12: 3188. https://doi.org/10.3390/nano11123188

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