Next Article in Journal
The Promise and Challenge of High Pressure Macromolecular Crystallography
Previous Article in Journal
Corrosion Resistance Performance of Epoxy Coatings Incorporated with Unmilled Micro Aluminium Pigments
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Decrease in the Adsorption Capacity of Adsorbents in the High-Temperature Carbonate Loop Process for CO2 Capture

Department of Gaseous and Solid Fuels and Air Protection, University of Chemistry and Technology Prague, 166 28 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
Crystals 2023, 13(4), 559; https://doi.org/10.3390/cryst13040559
Submission received: 6 February 2023 / Revised: 9 March 2023 / Accepted: 10 March 2023 / Published: 24 March 2023

Abstract

:
In this study, the sorption capacity of limestone samples for CO2 was investigated to determine the conditions under which they can be used in the high-temperature carbonate loop process. For the work, limestone samples from the Czech Republic were used, which contained a high proportion of CaO (more than 97 wt.%). A total of 20 cycles of calcination (950 °C) and subsequent CO2 sorption–carbonation (650 °C) were performed for each limestone sample tested. The sorption capacity towards CO2 in the 20th cycle was less than 10% of the value determined in the first carbonation cycle of the samples and the most significant decrease was observed between the first and second cycles. The highest sorption capacity was determined for the Branžovy sample, which captured 268 mL of CO2/per 1 g of sorbent by chemisorption. Only 15 mL of carbon dioxide per 1 g of sorbent was bound by physisorption. However, in repeated use, the Vitošov limestone had the highest sorption capacity for CO2. For all samples, the amount of carbon dioxide bound by physisorption was in the range of 4 to 10% of the amount bound by chemisorption. Due to sintering of the material, the BET specific surface area decreased by 95 to 96%.

1. Introduction

The carbonate loop is a post-combustion technology. Carbon dioxide, due to chemisorption, bonds to calcium oxide, forming calcium carbonate. Subsequently, calcium carbonate undergoes calcination, resulting in the release of carbon dioxide at elevated temperatures. The regenerated calcium oxide is reused in the carbonation process to bind carbon dioxide again. The process is repeated multiple times. During the carbonation process, heat is released. The heat released in the carbonator can be used for steam production. A similar amount of heat has to be supplied to the calcinator to provide the endothermic calcination reaction [1,2,3,4].
The main advantages of the high-temperature carbonate loop compared to other carbon capture technologies are lower operating cost, wide availability of natural limestone, and high carbon dioxide capture efficiency [5,6,7]. Limestone or dolomite can be used not only to capture carbon dioxide, but also other substances that are present in flue gases, such as SO2 [8,9]. Ströhle et al. tested a limestone carbonate looping process for a 1 MWth pilot plant and the total carbon dioxide capture was 92% [10]. Based on the study by Ströhle et al., Hilz et al. presented a publication on carbonate loop utilization in 2019 that showed a long-term project for a 20 MWth pilot unit that was intended for industrial use in an existing 600 MWel power plant [11]. An evaluation was carried out on a power plant burning pulverized coal with a capacity of 600 MW, which was equipped with a calcium carbonate loop. The simulation result showed that the net efficiency of the process was 33.8% with a carbon dioxide capture ratio of 94% [12].
In order for the process to be financially profitable, it is necessary for one batch of raw materials to realize, if possible, the largest number of calcination (CaCO3 decomposition) and carbonation (chemisorption of CO2 on CaO) work cycles. In the 1970s, Barker noted the thermal sintering problem, which limits the gradual reversibility of the reaction [13].
Wang et al. provided the carbonation reaction at the atomic level. The process is showed in Figure 1. Solid product molecules spread at the CO2–CaCO3 interface and connect in the existing C2O52− phase. Additionally, surface carbonates can combine with crystalline oxygen to form a metastable corner-bonded tetrahedron CO44− and diffuse into CaO particles. The increase in temperature accelerates the formation of external CO2-CO32− phases and the diffusion of internal CO32−, causing the adsorption of more CO2 molecules, and the speed of the carbonation reaction increases significantly [14].
The decrease in the sorption capacity of CaO used in the high-temperature carbonate loop process to capture CO2 at different partial pressures was also studied [14]. The authors state that the main contribution to the deactivation of the sorbent is the sintering of the material, which occurs under both normal and elevated pressures. The activity of the deactivated sorbent can be partially increased by its reaction with water vapor [15].
Authors have also closely studied the issue of CaO deactivation in the high-temperature carbonate loop process [16]. They described potential options to suppress the deactivation process, such as doping the material with alkali metals/alkaline earth metals, reducing the pressure of CO2, or injecting water vapor [16].
The carbonate loop generally consists of two interconnected fluidized bed reactors filled with limestone/calcium oxide [17]. The process diagram is shown in Figure 2.
Circulation in the fluidized bed equipment of limestone particles, which act as the sorption material, and mechanical stress, lead to a gradual size decrease. The decrease in particle size reduces the life time of the sorption material [18].
Fluidized bed technologies provide good contact between the gaseous phase and sorption material. In the case of the carbonate loop, limestone is the solid phase and the flue gas containing CO2 is the gaseous phase. The carbon dioxide content in flue gas is approximately 15% [2,19].
Limestones have a very low specific surface area. Limestone porosity is affected by temperature conditions and partial CO2 pressure. Natural limestone contains predominantly CaCO3. Equation (2) describes the endothermic decomposition of CaCO3 to porous CaO and gaseous CO2. Calcium oxide, which is one of the calcination products, has a more developed porous structure and a higher specific surface area, which allows good CO2 capture in the porous system, and subsequent reaction in the active sites of CaO is possible according to Equation (3).
The calcination is strongly influenced by the temperature in the range of 600–1000 °C, as shown in Figure 3. CaCO3 conversion increases rapidly with increasing temperature [20].
CaCO3 calcination (2) occurs at high temperature. The equilibrium pressure of CaCO3 must be higher than the partial pressure of carbon dioxide in the system [21]. The equilibrium pressure is calculated according to the following equation:
p C O 2 = e x p ( 17.74 0.00108 T + 0.332 ln T 22,020 T )
As the temperature increases, so does the equilibrium pressure. The concentrations of carbon dioxide in waste gases are around 15%. In order to achieve CO2 enrichment, calcination must take place below a concentration of 70% CO2. For this reason a higher calcination temperature than that of carbonation is needed. For example, to obtain a pure stream of CO2 with a partial pressure of 1 bar, a temperature greater than 900 °C is necessary [16]. As Hu and Scaroni (1996) described [22], as the temperature increases, the equilibrium pressure increases, so the partial pressure of carbon dioxide surrounding the material should be lower for complete and rapid decarbonation. The calcination reaction will not start when the pressure of CO2 is higher than the equilibrium pressure [22].
During calcination, calcium carbonate decomposition occurs at temperatures in the range of 850–950 °C, and thus carbon dioxide is released (Equation (2)). The regenerated calcium oxide is again utilized as a sorbent in the subsequent cycle. The released carbon dioxide is ready for further processing such as compression, transportation, and injection into the geological storage site [6,10,23,24]. During the CaO carbonation process, carbon dioxide chemisorption occurs (Equation (3)). Temperatures in the range of 550–750 °C are required for the reaction to proceed rapidly.
Calcination: CaCO3 (s) → CaO (s) + CO2 (g) ΔH298 = +179.2 kJ mol−1
Carbonation: CaO (s) + CO2 (g) → CaCO3 (s) ΔH298 = −179.2 kJ mol−1
Natural limestone also contains MgCO3. The reversible chemical reaction for removal of CO2 involving magnesium oxide is described by Equation (4). The thermal decomposition of magnesium carbonate produces magnesium oxide and carbon dioxide. This reaction occurs at a temperature greater than 650 °C [25].
MgO (s) + CO2 (g) → MgCO3 (s)
MgCO3 (s) → MgO (s) + CO2 (g)
Sulfur dioxide is a compound commonly present in flue gases. The presence of SO2 has a negative effect on the sorption of carbon dioxide onto CaO. Sulfur dioxide at a temperature of around 850 °C reacts with CaO to form calcium sulfate. Calcium sulfate is a very thermally stable compound that does not decompose even at calcination temperatures as high as 1300 °C. The reaction of sulfur dioxide with calcined limestone takes place as described in Equation (6).
CaO (s) + SO2 (g) + 1/2 O2 (g) → CaSO4 (s)
The reaction between SO2 and CaCO3, described in Equation (7), also reduces calcination efficiency.
CaCO3 (s) + SO2 (g) + 1/2 O2 (g) → CaSO4 (s) + CO2 (g)
The reaction between CaCO3 and gaseous SO2 is mainly influenced by temperature. The sulfation reaction rate is the fastest in the temperature range from 900 to 1100 °C. At a temperature above 1100 °C, the reaction rate decreases [26]. Sulfation under the reaction conditions commonly applied during the calcination and carbonation processes is a non-reversible process, which occurs at a much faster rate than carbon dioxide sorption [27]. Calcium sulfate is the crystal substance that, once formed, clogs primarily the surface pores of the sorption material. Therefore, the adsorption capacity of CO2 is reduced [28]. In combustion systems, it is possible to remove SO2 using direct limestone sulfation as described by Equation (7).
The purpose of this work was to investigate the possibility of using Czech Republic limestones in the high-temperature carbonate loop process, as the relevant properties of Czech limestones in connection with this technology have not yet been investigated. In the case of the application of a high-temperature carbonate loop in the territory of the Czech Republic, the use of these limestones should be advantageous due to the high content of CaO and, finally, due to the low procurement costs of the input material, which would not have to be imported from abroad.

2. Materials and Methods

Several types of limestones commonly available in the Czech Republic were used for this work to determine their suitability in the carbonate loop. It was therefore necessary to specify the changes in their properties before calcination, after calcination, and after exposure to carbon dioxide.
The suitable materials chosen to be tested as adsorbents in the high-temperature carbonate loop were four commercially available limestones from different deposits in the Czech Republic. The limestone samples were selected on the basis of previous high-temperature carbon dioxide adsorption capacity tests [29]. The adsorption capacity of CaO for CO2 depends primarily on its purity. Therefore, all limestone samples selected for testing contained a very high proportion of CaCO3. The samples were from the following deposits in the Czech Republic:
  • quarry Libotín
  • quarry Branžovy
  • quarry Vitošov
  • quarry Tetín
The limestone samples were prepared for testing in the high-temperature loop by crushing in a jaw crusher and sieving into fraction categories. The fraction size of 0.2–0.5 mm was selected to carry out the tests. This fraction size is often used in fluidized bed high-temperature carbonation loops. All samples were calcined in a muffle furnace at a temperature of 950 °C for 12 h in an inert atmosphere. Afterwards, in order to determine the structural and surface properties of the sample, the following analyses were performed: XRF, specific surface area, pore distribution, and total pore volume. The same analyses were performed using limestone samples after repeated carbon dioxide chemisorption tests at a temperature of 650 °C.

2.1. X-ray Fluorescence Analysis

X-ray fluorescence analysis (XRF) was performed using an ARL 9400 analyzer. XRF analysis enables the determination of the elemental composition of each limestone sample. The content of CaO and MgO was determined. On the basis of the content of calcium and magnesium oxides, it was possible to estimate the content of calcium and magnesium carbonate under the assumption that all oxides were converted into carbonates. The XRF analyzer is equipped with a special program that makes it possible to calculate the carbonate content in the sample [30]. The standard error of measurement for CaO is 0.06%, for MgO is 0.04%, for Al2O3 is 0.02%, for SiO2 is 0.03%, and for Fe2O3 is 0.02%. These analyses were performed to make it clear whether some of the tested samples do not contain significant amounts of impurities that could affect the deactivation of CaO in repeated calcination–carbonation cycles.

2.2. Specific Surface Area Measurement

The specific surface area was measured using a Coulter SA 3100 instrument (Beckman Coulter, Brea, CA, USA). The instrument works on the basis of the N2 physical adsorption from the gaseous phase at a temperature of −196 °C. Each sample was weighed into a special sample cell and subsequently degassed under deep vacuum at a temperature of 150 °C for 240 min. The sampling cell was again weighed and placed in the instrument’s measuring port to undergo sample evacuation under deep vacuum. After evacuation, the sample cell was immersed in liquid nitrogen and was dosed with precisely calculated volumes of gaseous nitrogen. Once the adsorption equilibrium stabilized, the equilibrium pressure was measured to calculate the amount of adsorbed nitrogen. The described procedure was first applied to determine the adsorption isotherm and then the desorption isotherm. Both isotherms were determined at a relative pressure in the range of 0–0.99 at a temperature of −196 °C. A pressure decrease in the sample cell containing the sample was carried out by gradually aspirating known nitrogen volumes from the sample cell. The measurements were evaluated on the basis of the shape of the adsorption and desorption isotherms. For the pressure range corresponding to relative pressure of 0–0.3, the quantity of adsorbed nitrogen was evaluated based on the BET equation and the BET specific surface area of each sample material determined from the coefficient obtained. The total pore volume was determined on the basis of the quantity of adsorbed nitrogen at a relative pressure close to 1. The pore size distribution was calculated from the desorption isotherms based on the Barret–Joyner–Halenda model, which uses the Kelvin equation.

2.3. Carbon Dioxide Sorption Determination Using the Quantachrome ASiQ Instrument

Cyclic CO2 sorption for each limestone sample was measured via the Quantachrome ASiQ sorption system. The sample weight was approximately 1 g. Initially, the sample cell containing the sample was heated to a temperature of 950 °C and the heating rate was 20 °C min−1. This led to the calcination of the sample according to Equation (2). The sample was left at a temperature of 950 °C for 4 h, during which the carrier gas (helium) flowed throughout the sampling. The time of calcination was selected based on previous laboratory tests, where the CO2 concentration was released by the CO2 analyzer. Afterward, evacuation was carried out, and the sample cell was left to cool at laboratory temperature. Once the sample cell had cooled, it was then weighed and the sample weight loss after calcination was determined. The following step was carbonation at a temperature of 650 °C in a pure carbon dioxide atmosphere. The reaction that takes place during this process is described by Equation (3). After carbonation, the sample was cooled to laboratory temperature and weighed to determine the weight gain. The calcination and carbonation cycles were repeated 20 times until the constant carbonation efficiency was confirmed.
The equipment recorded the total gas volume of carbon dioxide at a temperature of 650 °C and pressures in the range of 100–800 Torr. This volume represents the total volume used for both chemisorption and physisorption. After adsorption, the saturated material was evacuated for 120 min to remove the physically trapped amount from the sample (evacuation of the sample is sufficient to break the van der Waals bonds and release carbon dioxide) and the equipment measured the amount of gas released. The molecules of CO2 captured by chemisorption were firmly attached to the surface of the material, and to break the chemical bonds, more energy was needed in the form of an increase in temperature to 950 °C. The amount of CO2 released by heating the sample was also recorded by the equipment.

3. Results and Discussion

In Table 1 a comparison of the XRF elemental analysis is depicted for the four selected samples. The highest CaO content of 98.41 wt.% was found in the Branžovy sample. The lowest CaO content of 97.06 wt.% was found in the Tetín sample. Furthermore, the Tetín sample had the highest MgO content of 1.52 wt.%.
The samples were analyzed using a Coulter SA 3100 (i) before calcination, (ii) after calcination, and (iii) after undergoing the calcination-carbonation cycles in the Quantachrome ASiQ instrument. These analyses were performed to determine the changes in surface area and total pore volume as a result of the calcination process and repeated carbonation.
As can be seen from the results shown in Table 2, the specific surface area increased for all samples after calcination at a temperature of 950 °C. This phenomenon can most probably be attributed to the development of internal structures due to CO2 release. In contrast, after exposure to a carbon dioxide atmosphere, a decrease in specific surface area can be observed. The decrease in surface area is also partially caused by sample sintering. Note that sample sintering is a nonreversible process.
Pore size distribution using the Coulter SA 3100 was also determined before calcination, after calcination, and after twenty carbonation–calcination cycles. All acquired results are summarized in Figure 4, Figure 5, Figure 6 and Figure 7.
Comparing sample results in Figure 4, Figure 5, Figure 6 and Figure 7, it can be seen that the most prominent structural changes were in the mesopores and macropores. The only exception that demonstrated negligible changes was the Tetín sample, which was also the sample with the lowest and highest CaO and MgO content, respectively. The biggest structural changes were observed in the Libotín and Branžovy samples. In Figure 4 it can be observed that the pore size of the Libotín sample after calcination corresponded to the pore size after 20 cycles. In Figure 4, it can be seen that CaCO3 does not have micropores. After calcination of the sample, which contains CaO, the pores increased to 6 nm, and therefore the pore volume increased, as shown in Table 2. The thermal decomposition of CaCO3 increased the BET specific surface area due to pore formation in two size ranges of about 20–80 nm and below 6 nm. The same fact was published by Barker in 1973 [13]. He used mercury porosity to determine the pore volume distribution. During his study, calcium oxide showed a steep rise in pore volume in the 4 nm region and a large increase in the surface area due to the formation of pores in the two size distributions of 40 nm and smaller than 4 nm. This phenomenon was also noted by Han et al. 2022 [16], who found a decrease in the sorption capacity at a high calcination temperature. This was caused by an increase in CaO nanocrystals, and as soon as the conversion reached higher values, the sintering of the material also occurred [16].
The results agree with the research of Scaltsoyiannes and Lemonidou [31], who reported that CaO sintering during each calcination step affects porosity, and, conversely, CaCO3 sintering does not affect material deactivation. They experimentally verified that after each calcination step, the size of the CaO crystals increased, leading to a gradual loss in porosity and surface area with the number of cycles. They proposed a model that, based on measured data, accurately predicted limestone deactivation under different conditions [31]. The phenomenon of sintering during the carbonation process has also been investigated in other studies [32].
In Figure 5 it can be observed that the pore size of the Branžovy sample before calcination corresponded to the pore size after calcination; however, the pore size changed after twenty cycles. There was a significant increase in small pores and a decrease in macropores in this limestone, while at the same time the total pore volume was reduced. This may be due to sintering of the material. A study that examined the reduction in material sintering was described in a previous article [33]. Staf et al. examined the hydration of calcinates, which reduced the sintering of the material and suppressed the decrease in sorption capacity, which was stabilized after four cycles.
Repeated calcination–carbonation cycles led to the loss of small pores and an increase in the volume of macropores. These structural changes caused the loss of specific surface area, as mentioned in various research articles and confirmed by the results shown in Table 2. Like Basinas and his team [28], we found that (i) loss of specific surface area due to structural changes leads to a decrease in the calcium oxide sorption capacity for CO2, and (ii) sorption of carbon dioxide in CaO samples depends predominantly on material porosity and the volume of micropores.
The adsorption capacity of limestone was determined using Quantachrome ASiQ equipment under carbon dioxide atmosphere. All experimental results are depicted in Figure 8.
The theoretical sorption capacity for CO2 chemisorption was calculated from the CaCO3 content in the individual samples. For example, for the Libotín sample, the sorption capacity was 0.43 g CO2 per 1 g of adsorbent.
In the first adsorption cycle, all samples showed a higher sorption capacity for CO2 than their theoretical chemisorption capacity calculated from the CaO content in each sample. It is clear that a small part of CO2 was also sorbed by physical adsorption.
In all samples, repeated calcination–carbonation cycles caused a clear loss in sorption capacity, with the most prominent change being between the first and second cycles. Sorption capacity loss was probably caused by a decrease in pore volume during the first calcination–carbonation cycle. Antzara et al. observed the same phenomenon in their research [34]. In the literature, the most common reason for the decrease in the adsorption capacity of CaO for CO2 is the growth in the particle volume due to the formation of CaCO3 in the pores on the outer surface of the particles, which limits the diffusion of CO2 into the inner space of the particles. Another stated reason is the change in particle morphology associated with pore closure [35].
During the other cycles, the third to the fifteenth, a slight sorption capacity loss was also observed. Based on the results depicted in Figure 8, it can be stated that after the fifteenth cycle no additional significant sorption capacity loss was observed. Carbon dioxide sorption capacity loss can be primarily attributed to sorption material sintering as a result of long exposure to high temperatures during the calcination process. There was a significant decrease in the CO2 adsorption capacity after the first cycle, which may be due only to the sintering of the material.
According to data from the literature, partially deactivated CaO particles can be reactivated by the reaction of CaO with water vapor [33].
The adsorption isotherms for individual samples are shown in the following figures. The graphs show the amount of carbon dioxide used for physisorption and chemisorption that occurred in the limestones during the measurement. Figure 9 shows how much carbon dioxide volume was consumed for physisorption at a temperature of 650 °C.
Figure 10 shows the volume of carbon dioxide that was used for chemisorption at a temperature of 650 °C.

4. Conclusions

In this work, the CO2 adsorption capacity of CaO from various limestone samples was measured during 20 repeated cycles of calcination–carbonation. Experimental tests were carried out using four selected commercially available limestones from quarries in the Czech Republic. Carbon dioxide sorption capacity tests were performed using limestone particles with a fraction size of 0.2–0.5 mm. The properties and sorption capacities of the selected sorption materials were determined during repeated high-temperature carbon dioxide adsorption.
According to XRF analysis, Branžovy limestone had the highest CaO content (98.41 wt.%) and the Tetín limestone had the lowest (97.60 wt.%).
The assumption is that the adsorption of carbon dioxide in limestone takes place preferably as chemisorption at a temperature of 650 °C. After the first adsorption, all limestones showed a higher sorption capacity for CO2 than their theoretical chemisorption capacity, corresponding to the fact that part of the carbon dioxide was bound to the material by physisorption. During the measurement of individual limestones, the physisorption measurement was also evaluated using a Quantachrome ASiQ instrument. The highest amount of carbon dioxide sorbed by physisorption 23 mL g1 was consumed by the Vitošov sample, which contained the highest amount of Al2O3 and SiO2. The amount of CO2 consumed by physisorption ranged from 4 to 10% of the amount consumed by chemisorption for all samples.
The most noticeable loss in sorption capacity was observed between the first and second cycles. Sorption capacity loss was caused by material sintering. From the fifteenth cycle onwards, no significant sorption capacity loss was observed. The loss in carbon dioxide sorption capacity for CaO during repeated calcination–carbonation cycles was caused primarily by two factors. The first is a high carbon dioxide content, which causes fast saturation of the active layers in the CaO particles. The second is material sintering due to long exposure to high temperatures during the calcination process. The loss in sorption capacity was confirmed by the significant changes in the specific surface area and total pore volume in all sorption materials tested.
After 20 carbonation–calcination cycles, the pore volumes of the sorbents changed on average by 70%. This was probably caused by the closing of parts of the pore due to sintering, which limited the access of CO2 to the inner space of the adsorbent particles, and therefore its sorption capacity for CO2.
Even more significant was the drop in the BET specific surface of the tested samples, which was 95–96% after 20 carbonation–calcination cycles. This can also be attributed to the surface sintering of the adsorbent particles.
It should be noted that even a low sulfur dioxide content has a negative effect on the carbon dioxide sorption capacity of limestone. Therefore, it is recommended to perform flue gas desulfurization when limestone materials are used in a high-temperature–carbonate loop.
After the first adsorption, the highest CO2 sorption capacity was found in the Branžovy limestone, but when using the limestones in the carbonation–calcination cycles, Vitošov limestone was found to be the limestone with the highest sorption capacity. Therefore, due to its higher sorption capacity in repeated use, Vitošov limestone would be the best of the samples tested for application in the high-temperature carbonate loop process.

Author Contributions

Conceptualization, V.K., L.J. and K.C.; methodology, V.K., L.J. and K.C.; validation, V.K. and L.J.; formal analysis, V.K. and L.J.; investigation, V.K., L.J. and K.C.; data curation, V.K.; writing—original draft preparation, V.K.; writing—review and editing, V.K., L.J. and K.C.; supervision, K.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by Norway Grants (project no.: NF-CZ08-OV-1-005-2015 “Research of high temperature CO2 sorption from flue gas using carbonate loop”).

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shimizu, T.; Hirama, T.; Hosoda, H.; Inagaki, M.; Tejima, K. A Twin Fluid-Bed Reactor for Removal of CO2 from Combustion Processes. Chem. Eng. Res. Des. 1999, 77, 62–68. [Google Scholar] [CrossRef]
  2. Junk, M.; Kremer, J.; Ströhle, J.; Eimer, N.; Priesmeier, U.; Weingärtner, C.; Stallmann, O.; Epple, B. Design of a 20 MWth carbonate looping pilot plant for CO2-capture of coal fired power plants by means of limestone. Energy Procedia 2014, 63, 2178–2189. [Google Scholar] [CrossRef] [Green Version]
  3. Reitz, M.; Junk, M.; Ströhle, J.; Epple, B. Design and operation of a 300 kWth indirectly heated carbonate looping pilot plant. Int. J. Greenh. Gas Control 2016, 54, 272–281. [Google Scholar] [CrossRef]
  4. Fennell, P.S.; Davis, S.J.; Mohammed, A. Decarbonizing cement production. Joule 2021, 5, 1305–1311. [Google Scholar] [CrossRef]
  5. Ebrahimi, H.; Rahmani, M. A new design for CO2 capture and usage in a syngas production unit by carbonate chemical looping. J. Nat. Gas Sci. Eng. 2016, 36, 241–251. [Google Scholar] [CrossRef]
  6. Dean, C.C.; Blamey, J.; Florin, N.H.; Al-Jeboori, M.J.; Fennell, P.S. The calcium looping cycle for CO2 capture from power generation, cement manufacture and hydrogen production. Chem. Eng. Res. Des. 2011, 89, 836–855. [Google Scholar] [CrossRef] [Green Version]
  7. Hanak, D.P.; Anthony, E.J.; Manovic, V. A review of developments in pilot-plant testing and modelling of calcium looping process for CO2 capture from power generation systems. Energy Environ. Sci. J. 2015, 8, 2199–2249. [Google Scholar] [CrossRef] [Green Version]
  8. Huang, C.; Chang, K.; Yu, C.; Chiang, P.; Wang, C. Development of high-temperature CO2 sorbents made of CaO-based mesoporous silica. Chem. Eng. J. 2010, 161, 129–135. [Google Scholar] [CrossRef]
  9. Lee, M.S.; Goswami, D.Y.; Kothurkar, N.; Stefanakos, E.K. Development and evaluation of calcium oxide absorbent immobilized on fibrous ceramic fabrics for high temperature carbon dioxide capture. Powder Technol. 2015, 274, 313–318. [Google Scholar] [CrossRef] [Green Version]
  10. Ströhle, J.; Junk, M.; Kremer, J.; Galloy, A.; Epple, B. Carbonate looping experiments in a 1 MWth pilot plant and model validation. Fuel 2014, 127, 13–22. [Google Scholar] [CrossRef]
  11. Hilz, J.; Haaf, M.; Helbig, M.; Lindqvist, N.; Ströhle, J.; Epple, B. Scale-up of the carbonate looping process to a 20 MWth pilot plant based on long-term pilot tests. Int. J. Greenh. Gas Control 2019, 88, 221–341. [Google Scholar] [CrossRef]
  12. Rolfe, A.; Huang, Y.; Haaf, M.; Rezvani, S.; McIlveen-Wright, D.; Hewitt, N.J. Integration of the calcium carbonate looping process into an existing pulverized coal-fired power plant for CO2 capture: Techno-economic and environmental evaluation. Appl. Energy 2018, 222, 169–179. [Google Scholar] [CrossRef]
  13. Barker, R. The reversibility of the Reaction CaCO3=CaO+CO2. J. Appl. Chem. Biotechnol. 1973, 23, 733–743. [Google Scholar] [CrossRef]
  14. Wang, N.; Feng, Y.; Guo, X. Atomistic mechanisms study of the carbonation reaction of CaO for high temperature CO2 capture. Appl. Surf. Sci. 2020, 532, 147425. [Google Scholar] [CrossRef]
  15. Kuramoto, K.; Fujimoto, S.; Morita, A.; Shibano, S.; Suzuki, Y.; Hatano, H.; Shi-Ying, L.; Harada, M.; Takarada, T. Repetitive Carbonation-Calcination Reactions of Ca-Based Sorbents for Efficient CO2 Sorption at Elevated Temperatures and Pressures. Ind. Eng. Chem. Res. 2003, 42, 975–981. [Google Scholar] [CrossRef]
  16. Han, R.; Wang, Y.; Xing, S.; Pang, C.; Hao, Y.; Song, C.; Liu, Q. Progress in reducing calcination reaction temperature of Calcium-Looping CO2 capture technology: A critical review. Chem. Eng. J. 2022, 450, 137952. [Google Scholar] [CrossRef]
  17. Mehrpooya, M.; Sharifzadeh, M.M.M.; Rajabi, M.; Aghbashlo, M.; Tabatabai, M.; Hosseinpour, S.; Ramakrishna, S. Design of an integrated process for simultaneous chemical looping hydrogen production and electricity generation with CO2 capture. Int. J. Hydrog. Energy 2017, 42, 8486–8496. [Google Scholar] [CrossRef]
  18. Materic, V.; Holt, R.; Hyland, M.; Jones, M.I. An internally circulating fluid bed for attrition testing of Ca looping sorbents. Fuel 2014, 127, 116–123. [Google Scholar] [CrossRef]
  19. Valverde, J.M.; Sanchez-Jimenez, P.E.; Perez-Maqueda, L.A.; Quintanilla, M.A.S.; Perez-Vaquero, J. Role of crystal structure on CO2 capture by limestone derived CaO subjected to carbonation/recarbonation/calcination cycles at Ca–looping conditions. Appl. Energy 2014, 125, 264–275. [Google Scholar] [CrossRef] [Green Version]
  20. Zhong, Q.; Bjerle, I. Calcination kinetics of limestone and the microstructure of nascent CaO. Termochim. Acta 1993, 223, 109–120. [Google Scholar] [CrossRef]
  21. Silcox, G.D.; Kramlich, J.C.; Pershing, D.W. A mathematical model for the flash calcination of dispersed calcium carbonate and calcium hydroxide particles. Ind. Eng. Chem. Res. 1989, 28, 155–160. [Google Scholar] [CrossRef]
  22. Hu, N.; Scaroni, A.W. Calcination of pulverized limestone particles under furnace injection conditions. Fuel 1996, 75, 177–186. [Google Scholar] [CrossRef]
  23. Blamey, J.; Anthony, E.J.; Wang, J.; Fennell, P.S. The calcium looping cycle for large-scale CO2 capture. Prog. Energy Combust. Sci. 2010, 36, 260–279. [Google Scholar] [CrossRef]
  24. 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]
  25. Valverde, J.M.; Sanchez-Jimenez, P.E.; Perez-Maqueda, L.A. Ca-looping for postcombustion CO2 capture: A comparative analysis on the performances of dolomite and limestone. Appl. Energy 2015, 138, 202–215. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, H.; Xu, H.; Zheng, C.-G.; Qiu, J.-R. Temperature dependence on reaction of CaCO3 and SO2 in O2/CO2 coal combustion. J. Cent. South Univ. Technol. 2009, 16, 845–850. [Google Scholar] [CrossRef]
  27. Laursen, K.; Duo, W.; Grace, J.R.; Lim, J. Sulfation and reactivation characteristics of nine limestones. Fuel 2000, 79, 153–163. [Google Scholar] [CrossRef]
  28. Basinas, P.; YingHai, W.; Grammelis, P.; Anthony, E.J.; Grace, J.R.; Lim, C.J. Effect of pressure and gas concentration on CO2 and SO2 capture performance of limestones. Fuel 2014, 122, 236–246. [Google Scholar] [CrossRef]
  29. Jílková, L.; Vrbová, V. Testing of Selected Czech Limestones for Chemisorption of CO2. Paliva 2016, 4, 132–137. [Google Scholar] [CrossRef]
  30. Ciahotný, K.; Staf, M.; Hlinčík, T.; Vrbová, V.; Jílková, L.; Randáková, S. Removing carbon dioxide from flue gas using high temperature carbonate looping. Paliva 2015, 7, 84–90. [Google Scholar] [CrossRef]
  31. Scaltsoyiannes, A.A.; Lemonidou, A.A. On the factors affecting the deactivation of limestone under calcium looping conditions: A new comprehensive model. Chem. Eng. Sci. 2021, 243, 116797. [Google Scholar] [CrossRef]
  32. Labus, K. Comparison of the Properties of Natural Sorbents for the Calcium Looping Process. Materials 2021, 14, 548. [Google Scholar] [CrossRef]
  33. Staf, M.; Vrbová, V.; Jílková, L.; Miklová, B. Regeneration of sorption capacity of limestones for CO2 capture by introduction of water vapour. Paliva 2016, 8, 67–77. [Google Scholar] [CrossRef]
  34. Antzara, A.; Heracleous, E.; Lemonidou, A. Development of CaO-based Mixed Oxides as Stable Sorbents for Post-Combustion CO2 capture via Carbonate Looping. Energy Procedia 2014, 63, 2160–2169. [Google Scholar] [CrossRef] [Green Version]
  35. Alvarez, D.; Abanades, J.C. Pore-Size and Shape Effects on the Recarbonation Performance of Calcium Oxide Submitted to Repeated Calcination/Recarbonation Cycles. Energy Fuels 2005, 19, 270–278. [Google Scholar] [CrossRef]
Figure 1. Snapshots of CaCO3 shell and CaO core during carbonation reaction [14].
Figure 1. Snapshots of CaCO3 shell and CaO core during carbonation reaction [14].
Crystals 13 00559 g001
Figure 2. Schematic diagram of the calcium looping CO2 capture process [14].
Figure 2. Schematic diagram of the calcium looping CO2 capture process [14].
Crystals 13 00559 g002
Figure 3. Effect of reaction time and temperature on the calcination efficiency at 1 mbar and 600–1000 °C [20].
Figure 3. Effect of reaction time and temperature on the calcination efficiency at 1 mbar and 600–1000 °C [20].
Crystals 13 00559 g003
Figure 4. A comparison of Libotín limestone pore size distribution before calcination, after calcination, and after 20 cycles.
Figure 4. A comparison of Libotín limestone pore size distribution before calcination, after calcination, and after 20 cycles.
Crystals 13 00559 g004
Figure 5. A comparison of Branžovy limestone pore size distribution before calcination, after calcination, and after 20 cycles.
Figure 5. A comparison of Branžovy limestone pore size distribution before calcination, after calcination, and after 20 cycles.
Crystals 13 00559 g005
Figure 6. A comparison of Vitošov limestone pore size distribution before calcination, after calcination, and after 20 cycles.
Figure 6. A comparison of Vitošov limestone pore size distribution before calcination, after calcination, and after 20 cycles.
Crystals 13 00559 g006
Figure 7. A comparison of Tetín limestone pore size distribution before calcination, after calcination, and after 20 cycles.
Figure 7. A comparison of Tetín limestone pore size distribution before calcination, after calcination, and after 20 cycles.
Crystals 13 00559 g007
Figure 8. CO2 adsorption capacity of each CaO sample during each calcination-carbonation cycle.
Figure 8. CO2 adsorption capacity of each CaO sample during each calcination-carbonation cycle.
Crystals 13 00559 g008
Figure 9. Volume of carbon dioxide used for physisorption.
Figure 9. Volume of carbon dioxide used for physisorption.
Crystals 13 00559 g009
Figure 10. Volume of carbon dioxide used for chemisorption.
Figure 10. Volume of carbon dioxide used for chemisorption.
Crystals 13 00559 g010
Table 1. Selected oxide content in limestone samples via XRF analysis.
Table 1. Selected oxide content in limestone samples via XRF analysis.
Oxide Content [wt.%]
Limestone Sample IdentificationCaOMgOAl2O3SiO2Fe2O3
Libotín97.940.880.100.610.06
Branžovy98.410.700.130.250.21
Vitošov97.930.590.370.720.21
Tetín97.061.520.220.470.20
Table 2. Comparison of specific surface area and pore volume.
Table 2. Comparison of specific surface area and pore volume.
Limestone Sample IdentificationBET Specific Surface Area [m2 g−1]Total Pore Volume [mL g−1]
BeforeAfter
Calcination
After
20 Cycles
BeforeAfter
Calcination
After 20 Cycles
Libotín0.6773.5160.1570.0110.0150.005
Branžovy0.3804.0750.1700.0040.0260.001
Vitošov0.4557.7690.4920.0040.0450.013
Tetín0.5652.8040.1000.0060.0100.003
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

Kyselová, V.; Jílková, L.; Ciahotný, K. Decrease in the Adsorption Capacity of Adsorbents in the High-Temperature Carbonate Loop Process for CO2 Capture. Crystals 2023, 13, 559. https://doi.org/10.3390/cryst13040559

AMA Style

Kyselová V, Jílková L, Ciahotný K. Decrease in the Adsorption Capacity of Adsorbents in the High-Temperature Carbonate Loop Process for CO2 Capture. Crystals. 2023; 13(4):559. https://doi.org/10.3390/cryst13040559

Chicago/Turabian Style

Kyselová, Veronika, Lenka Jílková, and Karel Ciahotný. 2023. "Decrease in the Adsorption Capacity of Adsorbents in the High-Temperature Carbonate Loop Process for CO2 Capture" Crystals 13, no. 4: 559. https://doi.org/10.3390/cryst13040559

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