Next Article in Journal
Utilizing Limestone Alone for Integrated CO2 Capture and Reverse Water-Gas Reaction in a Fixed Bed Reactor: Employing Mass and Gas Signal Analysis
Previous Article in Journal
Experimental Study on Gas Production Capacity of Composite Reservoir Depletion in Deep Carbonate Gas Reservoirs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Competitive Adsorption of Moisture and SO2 for Carbon Dioxide Capture by Zeolites FAU 13X and LTA 5A

1
School of Minerals Processing and Bioengineering, Central South University, Changsha 410083, China
2
Hunan Valin Lianyuan Iron and Steel Co., Ltd., Loudi 417000, China
*
Author to whom correspondence should be addressed.
Processes 2024, 12(8), 1547; https://doi.org/10.3390/pr12081547
Submission received: 13 June 2024 / Revised: 17 July 2024 / Accepted: 22 July 2024 / Published: 24 July 2024
(This article belongs to the Section Separation Processes)

Abstract

:
Zeolites exhibit significant potential as porous materials for selective carbon dioxide (CO2) capture, leveraging their distinctive adsorption properties. However, the presence of moisture (H2O) and sulfur dioxide (SO2) in flue gas streams can significantly affect the efficiency of CO2 capture. This study investigates the CO2 adsorption characteristics of zeolites FAU 5A and LTA 13X, revealing the competitive adsorption mechanism between H2O(g), SO2, and CO2. The zeolites exhibit CO2 adsorption capacities of 93.19 mg/g and 95.80 mg/g for 5A and 13X, respectively, and demonstrate good regeneration potential. Metal cations correlated positively with CO2 adsorption. H2O(g), SO2, and CO2 exhibit a competitive adsorption relationship, with H2O(g) having the highest adsorption capacity, followed by SO2 and CO2. Additionally, the synergistic effect of SO2 and H2O(g) on CO2 adsorption is elucidated. These findings provide valuable insights into the competitive adsorption behavior of moisture and SO2 for CO2 capture using zeolites LTA 5A and FAU 13X, contributing to the development of more efficient CO2 capture processes and the design of tailored adsorbents for industrial applications.

1. Introduction

Global warming is closely linked to the rise in atmospheric greenhouse gas (GHG) concentrations, attributable to human emissions since the onset of the industrial era [1]. Carbon dioxide emissions (CO2) are one of the primary reasons for global warming [2,3]. Technologies for carbon capture, utilization, and storage (CCUS) are anticipated to be crucial in moving the world’s energy system toward net-zero emissions [4]. A wide range of technologies, including adsorption, membrane separation, low-temperature distillation, and absorption, have been proposed and applied to capture CO2 produced after burning [5,6,7,8,9]. Adsorption via solid sorbents is deemed an appealing alternative separation technology due to its simplicity and lack of liquid waste production [10,11]. Moreover, the regeneration of solid adsorbents is more efficient than that of liquid absorbents, primarily because CO2 is physiosorbed on solids, whereas it is chemisorbed on liquids. Consequently, solid adsorbents generally require less energy for regeneration [12].
Zeolite is a low-cost regeneration adsorbent utilized extensively in adsorption and gas separation due to its distinctive channel structure, outstanding thermal stability, and enormous adsorption capacity [4,13,14,15,16]. Siriwardane et al. compared the CO2 adsorption properties of 4A, 5A, 13X, and APG-II zeolites and discovered that zeolite 13X has a higher capacity for CO2 adsorption [17]. Cavenati et al. investigated the adsorption of CO2 and N2 on 13X zeolite at 4 MPa pressure [18] and demonstrated that at 298 K and 4 MPa pressure, 13X zeolite can adsorb CO2 at a rate of up to 7.5 mol/kg. The 13X and Y zeolites have a significant capability for CO2 adsorption [19]. Marcu and Sandulescu [20] used Y-type molecular sieves to adsorb SO2 and found that Y-type molecular sieves have good adsorption performance for SO2 at 25–150 °C, with an adsorption capacity of up to 20–175 mg/g. It can be followed that zeolites may have good potential to treat CO2 and SO2 simultaneously via co-adsorption.
One of the major obstacles preventing the large-scale growth of adsorption carbon capture is the presence of moisture in the industrial flue gas [21]. Microscale or even remnants of water can be successfully removed from the gas using molecular sieves to separate the mixture. Li et al. investigated the cyclic adsorption process of 12% CO2 + 95% RH mixtures at 30 °C and the penetration curve of CO2 and H2O(g) mixes on 13X zeolites [22,23]. According to the test findings, CO2 and moisture were competitive when absorbed by 13X zeolite, even in a vacuum. Stern et al. investigated the impact of 5A zeolite on the ability of N2 and CO2 to adsorb after water absorption [24] and demonstrated that zeolite molecular sieves lost their ability after absorbing water. They propose that the shift is caused by hydrate produced by water absorption clogging molecular sieve holes. Shen et al. conducted heat treatments on commercial NaX zeolites to alter the composition of the O-T-O (T = Al, Si) framework. By redistributing the Na+ ions within the O-T-O (T = Al, Si) framework, they effectively reduced the CO2 adsorption energy and inhibited the formation of water agglomerates [25]. The flue gas, apart from water, typically contains additional contaminants. Despite pretreatment processes aimed at eliminating compounds such as siloxanes, sulfides, and halides, trace impurities persist in the gas streams entering the separation column. The impact of these trace impurities on the functional capacity and longevity of adsorbents remains insufficiently explored.
The adsorption capacity of alkaline metal oxides to CO2 is often high, and the formed products have high strength and good wear resistance. Walton et al. investigated how modifying alkali metal cations affected the ability of X and Y zeolites to adsorb CO2 [26]. In the ion exchange-modified X-type zeolite, the carbon dioxide adsorption capacity declines with an increase in the metal cation radius; for the modified Y-type zeolite, the order of carbon dioxide adsorption capacity is Cs < Rb ≈ K < Li ≈ Na.
This study investigated the CO2 adsorption characteristics of the large-adsorption-capacity zeolite 13X with an S/A of 1.77 and zeolite 5A with an S/A of 1.51 using different molecular sieve particle sizes, adsorption temperatures, adsorption pressures, CO2 concentrations, regeneration times, and H2O(g) contents. Moreover, the impact of alkali metal cation alteration on the CO2 adsorption capabilities of X and A zeolites and the mechanism of competitive adsorption between SO2/H2O(g)/CO2 were primarily explored.

2. Materials and Methods

2.1. Materials

Zeolites were provided by Chinese National Pharmaceutical Group Chemical Rea-gent Co., Ltd. (Shanghai, China). In Table 1, the chemical compositions are displayed. Similar S/A values were found in the zeolite: 1.51 for 5A and 1.77 for 13X. Their cations differ significantly from one another. Zeolite 5A had a greater CaO concentration (11.84 wt%), but zeolite 13X had a higher Na2O content (13.39 wt%).
The specific surface areas and pore size distributions of zeolites 5A and 13X were analyzed using the Brunner–Emmet–Teller (BET) and Barrett–Joyner–Halenda (BJH) methods. Zeolites 5A and 13X exhibited specific surface areas of 411 and 472 m2/g, respectively [27]. Both demonstrated an abundant microporous structure, with micropore volumes of 0.21 and 0.15 cm3g−1 and an average pore diameter of 2.82 and 2.31 nm, and SEM images in Figure 1 also support the results.

2.2. Methods

2.2.1. Experimental Procedure

The test process is consistent with our previous report [24]. A schematic diagram of the experimental apparatus used in this study is shown in Figure 2. Two layers of nickel foam were used to hold the bed in place for the adsorption test, which was conducted in a U-tube with a 10 mm diameter. The measurements of the CO2 concentration and the temperature and relative humidity make up the two primary parts of the testing system. An enhanced flue gas analyzer (MGA-5, YORK Instrument Ltd., Heilbronn, Germany) with a measurement accuracy of 1 ppm was used to detect CO2 gas. A flue gas analyzer partly extracted the fixed bed’s exhaust gas for analysis, and the remaining gas was absorbed using a scrubber bottle filled with a sodium hydroxide solution. In addition, relative humidity and temperature were detected using temperature and humidity loggers with an accuracy of 2% RH and 0.1 °C, respectively.

2.2.2. Molecular Sieve Activation and Modification Methods

The zeolite molecular sieves were initially activated before usage to eliminate the adsorbed gas and moisture. The procedure involves placing the molecular sieve in a porcelain boat, transferring it into a muffle furnace set at a specific baking temperature, and subjecting it to a predetermined baking duration. The activated molecular sieve is rapidly removed and placed in a drying dish with water-absorbing silica gel for testing and subsequent detection to prevent it from absorbing moisture and other gases in the air.
Zeolite was altered via acid modification to evaluate the impact of cations and SiO2/Al2O3 on adsorption capacity. The following were the steps: The two molecular sieves were uniformly combined with 1:10 (g/mL) solid-to-liquid ratios of hydrochloric acid and hydrofluoric acid solutions of various strengths, and they were then left to react at 30 °C for 4 h. After modification, the molecular sieve was filtered using a vacuum filter and washed with anhydrous ethanol. Subsequently, the molecular sieve was dried in a constant temperature blast oven at 100 °C and then activated at 300 °C for 2 h in a muffle furnace.

2.2.3. Distributed Water Distribution Method

By adjusting the various aeration times, the distributed water distribution system fulfills the goal of controlling gas moisture. The nitrogen valve that carries water should be closed once the required preparation for water distribution is finished. The CO2 mixture, having the desired concentration, is introduced into the molecular sieve fixed bed to examine the adsorption characteristics when the temperature of the solid bed falls to 30 °C, which eliminates the thermal effect of the moisture adsorption process.

2.2.4. Evaluation of CO2 Adsorption Capacity

The adsorption property of CO2 was determined using the breakthrough curve method, and the adsorption capacity (A) and the breakthrough time (t) were used to characterize the adsorption characteristics of zeolites [28,29]. The adsorption capacity (A) was defined as in Equation (1):
A = K M Q t 0 t i ( C 0 C ) d t / ( m × V m )
where A is the unit zeolite mass of the adsorption gas mass (mg/g); Q is the total gas flow rate (m3/s); m is the mass of the adsorbent (g); M is the molar mass of the adsorbed gases (g/mol); Vm is the standard molar volume (m3/mol); t0 is the start time of adsorption (s); ti is the arbitrary time of adsorption process that is used as a variable in the integration process (s); C0 is the inlet CO2 concentration (ppm); and C is the outlet CO2 concentration at an arbitrary time during the adsorption process (ppm).
Breakthrough time (t) was defined as in Equation (2):
t = t 1 t 0
where t0 is the start time of the experiment and t1 is the time when the exhaust gas concentration is equal to 1% of the inlet concentration.

3. Results and Discussion

3.1. Influencing Parameters on CO2 Adsorption Capacity

3.1.1. Effect of Molecular Sieve Particle Size

The impact of different molecular sieve particle sizes on the dynamic adsorption capacity of CO2 was studied under the conditions of a molecular sieve of 1 g, an activation temperature of 300 °C, an activation time of 2 h, a CO2 concentration of 10%, a mixed gas flow rate of 1 L/min, an adsorption temperature of 30 °C, an adsorption pressure of 1 atm, and a relative humidity of 24%. The initial rising slope of the integration curve in Figure 3 demonstrates that the CO2 adsorption amount on the molecular sieves increased progressively as the particle size decreased, enhancing the CO2 adsorption rate. When the particle size exceeds 2.8 mm, both the adsorption capacity and rate are lower compared to other sizes. This effect is attributed to the larger molecular sieves having a smaller specific surface area, which limits the adsorption rate, and to the slower diffusion rate of the adsorbate within the adsorbent, further constraining the adsorption process. Nonetheless, the maximum CO2 adsorption capacities for the 5A and 13X molecular sieves reach 93.19 mg/g and 95.80 mg/g, respectively, at a particle size of 0.26–0.43 mm. Beyond this range, the CO2 adsorption capacity experiences minimal variation with further decreases in particle size. Additionally, a reduction in particle size increases the resistance within the bed, thereby rendering the properties of the fixed bed more susceptible to alteration. The subsequent experiments employed a molecular sieve with a particle size ranging from 0.26 to 0.43 mm.

3.1.2. Effect of Activation Temperature and Time

The effect of activation temperature and time on the dynamic adsorption capacity of CO2 was studied under the following conditions: a molecular sieve of 1 g, a CO2 concentration of 10%, a mixed gas flow rate of 1 L/min, an adsorption temperature of 30 °C, an adsorption pressure of 1 atm, and a relative humidity of 24%. As depicted in Figure 4, the adsorption capacity of CO2 initially escalates sharply with increases in activation temperature, subsequently stabilizing. Notably, at an activation temperature of 300 °C, the maximal adsorption capacities for zeolites 5A and 13X were recorded at 93.19 mg/g and 95.80 mg/g, respectively. As the activation time increased, the saturation adsorption capacity of CO2 for both the 5A and 13X molecular sieves initially rose and then stabilized. Specifically, the capacity in 5A sieves stabilized after 1 h of activation, whereas in 13X sieves, it did so after 2 h. Consequently, to facilitate conditions for subsequent comparative analyses, an activation duration of 2 h was chosen for both types of sieves.

3.1.3. Effect of Adsorption Temperature

One of the most significant factors affecting zeolite molecular sieve adsorption is adsorption temperature [24]. The impact of fixed-bed adsorption temperature on the CO2 adsorption capabilities of two zeolite molecular sieves was examined under the conditions of a molecular sieve dosage of 1 g, an activation temperature of 300 °C, an activation time of 2 h, a mixed gas flow rate of 1 L/min, an adsorption pressure of 1 atm, a CO2 gas distribution concentration of 10%, and a relative humidity of 24%. Figure 5 demonstrates that as the adsorption temperature was raised from 30 °C to 100 °C, the CO2 adsorption amount of the 5A molecular sieves decreased from 93.19 mg/g to 19.55 mg/g, a decrease of 79.01%; and the CO2 adsorption amount of the 13X molecular sieve decreased from 95.80 mg/g to 25.78 mg/g, a reduction of 72.91%. The adsorption of CO2 by both zeolite molecular sieves is temperature-sensitive and the molecular sieves of the 13X and 5A types are less effective at dynamic CO2 adsorption at higher temperatures. The CO2 gas adsorption process suffers from a rise in temperature since the adsorption process is exothermic [30]. It might be related to the nonpolar properties of CO2, a molecule with a weak physical adsorption force and rapid desorption at high temperatures [31].

3.1.4. Effect of CO2 Concentration

The CO2 concentration of exhaust gases from sophisticated industrial operations is unstable [32,33]. Figure 6 illustrates the effect of varying CO2 gas concentrations on the CO2 adsorption of two zeolite molecular sieves under the conditions of a molecular sieve dosage of 1 g, an activation temperature of 300 °C, an activation time of 2 h, a mixed gas flow rate of 1 L/min, an adsorption temperature of 30 °C, an adsorption pressure of 1 atm, and a relative humidity of 24%. The adsorption capacities of the two molecular sieves grew as the CO2 concentration in the mixed gas rose from 1% to 10%. The CO2 adsorption capacities of the 5A type molecular sieve increased from 23.93 mg/g to 93.19 mg/g, and the CO2 adsorption capacities of the 13X type molecular sieve increased from 38.63 mg/g to 95.80 mg/g. The adsorption capacities of both molecular sieves were enhanced at elevated CO2 concentrations. On the one hand, an increase in concentration increases the likelihood that CO2 molecules will bind to the molecular sieve’s active sites. On the other hand, an increase in concentration results in a greater concentration difference between CO2 on the molecular sieve’s surface and the gas phase, which increases the adsorption driving force and shifts the adsorption equilibrium.

3.1.5. Effect of Acid Cations and SiO2/Al2O3

Zeolite was altered via acid modification to evaluate the impact of cations and SiO2/Al2O3 on adsorption capacity. Figure 7 demonstrates that HCl and HF modification induces similar trends in the SiO2/Al2O3 ratios and CaO and Na2O contents in 5A and 13X zeolites under the conditions of a molecular sieve dosage of 1 g, an activation temperature of 300 °C, an activation time of 2 h, a mixed gas flow rate of 1 L/min, an adsorption temperature of 30 °C, an adsorption pressure of 1 atm, and a CO2 gas distribution concentration of 10%. Specifically, HF treatment reduces the SiO2/Al2O3 ratios in both zeolites, with minimal changes in CaO and Na2O levels (Figure 8). This reduction negatively affects the CO2 penetration time, as evidenced by dynamic breakthrough curve analyses. Conversely, HCl modification increases the SiO2/Al2O3 ratios while decreasing CaO and Na2O contents in these zeolites, which, according to the dynamic breakthrough curve of CO2, adversely impacts the penetration time due to the significant decrease in metal cation (Ca2+, Na+) levels.

3.1.6. Regeneration Potential of the Zeolites

The effect of cycle regeneration times on the CO2 adsorption capacities in two zeolite molecular sieves was investigated using the following conditions: a molecular sieve dosage of 1 g, an adsorption temperature of 30 °C, a mixed gas flow rate of 1 L/min, a CO2 gas concentration of 10%, an adsorption pressure of 1 atm, and a relative humidity of 24%. The molecular sieve samples underwent a cycle procedure of ‘adsorption saturation–activation desorption–adsorption saturation’. The activation desorption phase was conducted at 300 °C for 2 h. The 5A and 13X molecular sieves each completed seven adsorptive cycles during testing. Figure 9 demonstrates that after seven cycles, the saturation adsorption capacities of both molecular sieves varied slightly, maintaining their original performance levels. After activation at 300 °C for 2 h, the molecular sieve releases the adsorbed CO2 without compromising its microstructure or pore system. This preservation of structural integrity ensures continued good adsorption performance and enables effective regeneration of the molecular sieve.

3.2. Competitive Adsorption between H2O(g), SO2, and CO2

3.2.1. Effects of Moisture Content

Under the conditions of a molecular sieve dosage of 1 g, an activation temperature of 300 °C, an activation time of 2 h, a mixed gas flow rate of 1 L/min, an adsorption temperature of 30 °C, an adsorption pressure of 1 atm, and a CO2 gas distribution concentration of 10%, the effect of moisture on the dynamic adsorption characteristics of CO2 was studied (Figure 10). The variation in moisture content was reflected by fluctuations in relative humidity during the experiment.
The lowest value of CO2 concentration gradually increases during the adsorption process as relative humidity increases, and it is easier to reach adsorption saturation quickly after bed penetration. During the adsorption, H2O(g) fulfilled dual roles: It engaged in competitive adsorption with CO2, thereby diminishing the adsorption capacity of CO2. Concurrently, the adsorption of H2O(g) on zeolite generated substantial heat, which elevated the temperature of the fixed bed. This temperature increase, in turn, led to a reduction in CO2 adsorption capacity.
To verify competitive adsorption, a stepwise water distribution method was used to eliminate the influence of H2O(g) adsorption heat. According to Figure 11, the CO2 penetration curves on the two molecular sieves follow similar laws as the water distribution time increases. The minimum value of carbon dioxide in the adsorption process gradually increases as the moisture content increases, and carbon dioxide penetrates the fixed bed in a shorter time. It demonstrates that as the mixed water content increases, the adsorption capacity of CO2 on the two molecular sieves gradually decreases. When the water distribution time is 60 min for the two molecular sieves, the adsorption capacity of CO2 molecules is lost. When the relative humidity of the dynamic water distribution is much lower than that of the stepwise water distribution, the CO2 adsorption capacity obtained via the dynamic real-time water distribution method is lower than that of the stepwise water distribution. When the relative humidity of the dynamic water distribution is 60% RH, the lowest C/C0 value of CO2 in the 5A molecular sieve adsorption process is 0.59, and the lowest C/C0 value of CO2 in the 13X molecular sieve adsorption process is 0.55. However, when the step-by-step water distribution process is used, the lowest C/C0 value of CO2 during the 5A molecular sieve adsorption process is 0.43 and the lowest C/C0 value of CO2 during the 13X molecular sieve adsorption process is 0.42. This phenomenon occurs because the step-by-step water distribution process eliminates the influence of moisture adsorption heat on CO2 adsorption.
To demonstrate the heat effect of water vapor adsorption, water vapor with a relative humidity of 90% RH at 30 °C was introduced into a fixed bed with a molecular sieve using the nitrogen loading method. Figure 12 shows that water vapor releases a significant amount of adsorption heat during the adsorption process. The molecular sieve fixed bed reaches a maximum temperature of 69 °C at 320 s. When the flow rate is 1000 mL/min, the maximum temperature of the layer reaches 80.5 °C at 280 s, which is 50.5 °C higher than without steam. The maximum temperature of the molecular sieve fixed bed reaches 97 °C at 80 s when the flow rate is set to 2000 mL/min, which is 67 °C higher than the temperature without steam. The maximum temperature of the bed gradually increases with the increase in water vapor per unit of time, and the time when the maximum temperature appears gradually decreases. The effect of adsorption temperature on CO2 adsorption capacity has already been discussed.

3.2.2. Synergistic Effects of SO2 and H2O(g)

Varying levels of SO2 are frequently present in industrial exhaust gases [34,35]. The effect of SO2 on the CO2 breakthrough curve was studied under the conditions of a molecular sieve dosage of 1 g, an activation temperature of 300 °C, an activation time of 2 h, a mixed gas flow rate of 1 L/min, an adsorption temperature of 30 °C, an adsorption pressure of 1 atm, and a 24% relative humidity, and the change in relative humidity was observed during the process (Figure 13). The SO2 penetration times were 440 s for the 5A molecular sieve at a concentration of 4436 ppm and 510 s for the 13X molecular sieve at 3942 ppm. Both zeolite molecular sieves exhibited robust desulfurization capabilities within the mixed flue gas containing SO2, H2O(g), and CO2. However, the CO2 adsorption was hampered in the presence of moisture and SO2, and CO2 only remained adsorbent for a short time. CO2 penetration times of 80 s and 90 s were calculated for molecular sieves of types 5A and 13X, respectively, and their adsorption amounts were 15% and 18% of those without SO2 dosing.
This phenomenon is caused by the polarity of the SO2, H2O(g), and CO2 molecules. Because the zeolite molecular sieve is composed of aluminum–oxygen tetrahedra, a metal cation is required to balance the excess negative charge of the aluminum ions. The adsorption of adsorbent molecules on zeolite molecular sieves is subject to strong adsorption forces due to the presence of metal cations. Under certain conditions, the polarity of the adsorbent molecules is related to the strength of the adsorption force. SO2 and H2O(g) molecules are polar molecules while CO2 is nonpolar. Although both SO2 and H2O(g) are polar molecules, their molecular polarity is different; the SO2 molecule’s sulfur–oxygen bond’s electronegativity difference (0.86) is smaller than the H2O(g) molecule’s oxygen–hydrogen bond’s (1.24), and the larger the difference in electronegativity, the more polar the chemical bond [36,37]. The SO2 molecule’s bond angle is also higher than that of H2O(g). Thus, the order of affinity during adsorption between the metal cation and the adsorbent molecule is H2O(g) > SO2 > CO2, which is consistent with the findings of the experiments. This suggests that if the adsorbent molecules’ kinetic diameters are identical, the adsorbent molecules’ polarity significantly impacts the adsorption process.

4. Conclusions

This study examined the elements that affect the CO2 adsorption capacity of the 5A and 13X zeolite molecular sieves under diverse conditions. The adsorption capacity of the 5A and 13X zeolite molecular sieves, under optimized conditions, was found to be 93.19 mg/g and 95.80 mg/g, respectively, both showcasing impressive cyclic adsorption stability during CO2 adsorption.
During the co-adsorption process involving multiple components, a competitive adsorption relationship was identified among the three adsorbents: SO2, H2O(g), and CO2. The affinity between the adsorbent molecules and the molecular sieve was found to be directly related to the polarity of the adsorbent molecules—the higher the polarity, the stronger the affinity with the molecular sieve. The affinity on the zeolites follows the following sequence: H2O(g) > SO2 > CO2.
The application of a dynamic real-time water distribution method, coupled with a stepwise water distribution method, revealed that H2O(g) plays a dual role in the adsorption process. Firstly, there is competitive adsorption between CO2 and water molecules, leading to a decrease in CO2 adsorption capacity with an increase in water vapor content. Secondly, the adsorption of water vapor on the molecular sieve generates substantial adsorption heat, causing a rise in bed temperature and, consequently, a further decrease in CO2 adsorption capacity.

Author Contributions

Conceptualization, X.Y. and J.C.; methodology, X.Y.; software, Q.W.; validation, L.X., H.L. and M.R.; formal analysis, J.C.; investigation, Q.W.; resources, X.Y.; data curation, Q.W.; writing—original draft preparation, X.Y.; writing—review and editing, J.C.; visualization, J.C.; supervision, M.R.; project administration, M.R.; funding acquisition, X.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

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

Conflicts of Interest

Author Xiduan Yang was employed by the company Hunan Valin Lianyuan Iron and Steel Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. The company had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Houghton, J.T.; Ding, Y.; Griggs, D.J.; Noguer, M.; Linden, P.; Dai, X.G.; Maskell, K.; Johnson, C. Climate Change 2001: The Scientific Basis; Cambridge University Press: Cambridge, UK, 2001; Volume 81, p. 208. [Google Scholar]
  2. Kim, E.J.; Siegelman, R.L.; Jiang, H.Z.H.; Forse, A.C.; Lee, J.H.; Martell, J.D.; Milner, P.J.; Falkowski, J.M.; Neaton, J.B.; Reimer, J.A.; et al. Cooperative carbon capture and steam regeneration with tetraamine-appended metal-organic frameworks. Science 2020, 369, 392–396. [Google Scholar] [CrossRef] [PubMed]
  3. Singh, 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] [CrossRef]
  4. IEA. Legal and Regulatory Frameworks for CCUS, IEA, Paris. 2022. Available online: https://www.iea.org/reports/legal-and-regulatory-frameworks-for-ccus (accessed on 19 January 2023).
  5. Mazari, S.A.; Si Ali, B.; Jan, B.M.; Saeed, I.M.; Nizamuddin, S. An overview of solvent management and emissions of amine-based CO2 capture technology. Int. J. Greenh. Gas Control 2015, 34, 129–140. [Google Scholar] [CrossRef]
  6. Pałka, K.; Pokrowiecki, R. Porous titanium implants: A review. Adv. Eng. Mater. 2018, 20, 15–25. [Google Scholar] [CrossRef]
  7. Drage, T.C.; Snape, C.E.; Stevens, L.A.; Wood, J.; Wang, J.; Cooper, A.I.; Dawson, R.; Guo, X.; Satterley, C.; Irons, R. Materials challenges for the development of solid sorbents for post-combustion carbon capture. J. Mater. Chem. 2012, 22, 2815–2823. [Google Scholar] [CrossRef]
  8. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341, 1230444. [Google Scholar] [CrossRef]
  9. Giordano, L.; Roizard, D.; Favre, E. Life cycle assessment of post-combustion CO2 capture: A comparison between membrane separation and chemical absorption processes. Int. J. Greenh. Gas Control 2018, 68, 146–163. [Google Scholar] [CrossRef]
  10. Ho, M.T.; Allinson, G.W.; Wiley, D.E. Reducing the Cost of CO2 Capture from Flue Gases Using Pressure Swing Adsorption. Ind. Eng. Chem. Res. 2008, 47, 4883–4890. [Google Scholar] [CrossRef]
  11. Harrison, D.P. The role of solids in CO2 capture: A mini review. In Greenhouse Gas Control Technologies 7; Rubin, E.S., Keith, D.W., Gilboy, C.F., Wilson, M., Morris, T., Gale, J., Thambimuthu, K., Eds.; Elsevier Science Ltd.: Oxford, UK, 2005; pp. 1101–1106. [Google Scholar]
  12. Samanta, A.; Zhao, A.; Shimizu, G.K.H.; Sarkar, P.; Gupta, R. Post-Combustion CO2 Capture Using Solid Sorbents: A Review. Ind. Eng. Chem. Res. 2012, 51, 1438–1463. [Google Scholar] [CrossRef]
  13. Yousef, A.M.; El-Maghlany, W.M.; Eldrainy, Y.A.; Attia, A. New approach for biogas purification using cryogenic separation and distillation process for CO2 capture. Energy 2018, 156, 328–351. [Google Scholar] [CrossRef]
  14. Ho, T.M.; Howes, T.; Bhandari, B.R. Encapsulation of gases in powder solid matrices and their applications: A review. Powder Technol. 2014, 259, 87–108. [Google Scholar] [CrossRef]
  15. Hefti, M.; Marx, D.; Joss, L.; Mazzotti, M. Adsorption equilibrium of binary mixtures of carbon dioxide and nitrogen on zeolites ZSM-5 and 13X. Microporous Mesoporous Mater. 2015, 215, 215–228. [Google Scholar] [CrossRef]
  16. Zhang, C.; Xu, H.; Geng, X.; Wang, J.; Xiao, J.; Zhu, P. Effect of spray distance on microstructure and tribological performance of suspension plasma-sprayed hydroxyapatite–titania composite coatings. J. Therm. Spray Tech. 2016, 25, 1255–1263. [Google Scholar] [CrossRef]
  17. Siriwardane, M.S.; Shen, E.P. Fisher Adsorption of CO2 on Zeolites at Moderate Temperatures. Energy Fuels 2005, 19, 1153–1159. [Google Scholar] [CrossRef]
  18. Cavenati, S.; Grande, C.A.; Rodrigues, A.E. Adsorption Equilibrium of Methane, Carbon Dioxide, and Nitrogen on Zeolite 13X at High Pressures. J. Chem. Eng. Data 2004, 49, 1095–1101. [Google Scholar] [CrossRef]
  19. Shao, W.; Zhang, L.; Li, L.; Lee, R.L. Adsorption of CO2 and N2 on synthesized NaY zeolite at high temperatures. Adsorption 2009, 15, 497. [Google Scholar] [CrossRef]
  20. Marcu, I.C.; Sandulescu, I. Study of sulfur dioxide adsorption on Y zeolite. J. Serbian Chem. Soc. 2004, 69, 563–569. [Google Scholar] [CrossRef]
  21. Boer, D.G.; Langerak, J.; Pescarmona, P.P. Zeolites as Selective Adsorbents for CO2 Separation. ACS Appl. Energy Mater. 2023, 6, 2634–2656. [Google Scholar] [CrossRef]
  22. Li, G.; Xiao, P.; Webley, P.; Zhang, J.; Singh, R.; Marshall, M. Capture of CO2 from high humidity flue gas by vacuum swing adsorption with zeolite 13X. Adsorption 2008, 14, 415–422. [Google Scholar] [CrossRef]
  23. Li, G.; Xiao, P.; Webley, P.A.; Zhang, J.; Singh, R. Competition of CO2/H2O in adsorption based CO2 capture. Energy Procedia 2009, 1, 1123–1130. [Google Scholar] [CrossRef]
  24. Stern, S.A.; DiPaolo, F.S. The Adsorption of Atmospheric Gases on Molecular Sieves at Low Pressures and Temperatures. The Effect of Preadsorbed Water. J. Vac. Sci. Technol. 1967, 4, 347–355. [Google Scholar] [CrossRef]
  25. Shen, M.; Kong, F.; Guo, W.; Zuo, Z.; Gao, T.; Chen, S.; Tong, L.; Zhang, P.; Wang, L.; Chu, P.K.; et al. Impact of H2O on CO2 adsorption and co-adsorption: Mechanism and high-performance adsorbents for efficient H2O-CO2 capture. Chem. Eng. J. 2024, 479, 147923. [Google Scholar] [CrossRef]
  26. Walton, K.S.; Abney, M.B.; Douglas LeVan, M. CO2 adsorption in Y and X zeolites modified by alkali metal cation exchange. Microporous Mesoporous Mater. 2006, 91, 78–84. [Google Scholar] [CrossRef]
  27. Li, G.H.; Wang, Q.S.; Jiang, T.; Luo, J.; Rao, M.J.; Peng, Z.W. Roll-up effect of sulfur dioxide adsorption on zeolites FAU 13X and LTA 5A. Adsorption 2017, 23, 699–710. [Google Scholar] [CrossRef]
  28. Shah, H.R.; Ahmad, K.; Bashir, M.S.; Shah, S.S.A.; Najam, T.; Ashfaq, M. Metal organic frameworks for efficient catalytic conversion of CO2 and CO into applied products. Mol. Catal. 2022, 517, 112055. [Google Scholar] [CrossRef]
  29. Ahmad, K.; Nazir, M.A.; Qureshi, A.K.; Hussain, E.; Najam, T.; Javed, M.S.; Shah, S.S.A.; Tufail, M.K.; Hussain, S.; Khan, N.A.; et al. Engineering of Zirconium based metal-organic frameworks (Zr-MOFs) as efficient adsorbents. Mater. Sci. Eng. B 2020, 262, 114766. [Google Scholar] [CrossRef]
  30. Liu, B.; Lian, Y.; Li, S.; Deng, S.; Zhao, L.; Chen, B.; Wang, D. Experimental investigation on separation and energy-efficiency performance of temperature swing adsorption system for CO2 capture. Sep. Purif. Technol. 2019, 227, 115670. [Google Scholar] [CrossRef]
  31. Liu, H.; Meng, X.; Dao, T.D.; Zhang, H.; Li, P.; Chang, K.; Wang, T.; Li, M.; Nagao, T.; Ye, J. Conversion of Carbon Dioxide by Methane Reforming under Visible-Light Irradiation: Surface-Plasmon-Mediated Nonpolar Molecule Activation. Angew. Chem. 2015, 54, 11545–11549. [Google Scholar] [CrossRef] [PubMed]
  32. Fan, X.; Wong, G.; Gan, M.; Chen, X.; Yu, Z.; Ji, Z. Establishment of refined sintering flue gas recirculation patterns for gas pollutant reduction and waste heat recycling. J. Clean. Prod. 2019, 235, 1549–1558. [Google Scholar] [CrossRef]
  33. Gao, C.; Ge, H.; Lu, Y.; Wang, W.; Zhang, Y. Decoupling of provincial energy-related CO2 emissions from economic growth in China and its convergence from 1995 to 2017. J. Clean. Prod. 2021, 297, 126627. [Google Scholar] [CrossRef]
  34. Wang, S.; Zhang, Q.; Zhang, G.; Wang, Z.; Zhu, P. Effects of sintering flue gas properties on simultaneous removal of SO2 and NO by ammonia-Fe(II)EDTA absorption. J. Energy Inst. 2017, 90, 522–527. [Google Scholar] [CrossRef]
  35. Liu, J.; Wang, S.; Yi, H.; Tang, X.; Li, Z.; Yu, Q.; Zhao, S.; Gao, F.; Zhou, Y.; Wang, Y. Air pollutant emission and reduction potentials from the sintering process of the iron and steel industry in China in 2017. Environ. Pollut. 2022, 307, 119512. [Google Scholar] [CrossRef] [PubMed]
  36. Li, C.Q.; Liu, G.R.; Qin, S.; Zhu, T.Y.; Song, J.F.; Xu, W.Q. Emission reduction of PCDD/Fs by flue gas recirculation and activated carbon in the iron ore sintering. Environ. Pollut. 2023, 327, 121520. [Google Scholar] [CrossRef]
  37. Shen, C.M.; Worek, W.M. Cosorption characteristics of solid adsorbents. Int. J. Heat Mass Transf. 1994, 37, 2123–2129. [Google Scholar] [CrossRef]
Figure 1. Morphology of the zeolites (a) 13X and (b) 5A.
Figure 1. Morphology of the zeolites (a) 13X and (b) 5A.
Processes 12 01547 g001
Figure 2. Experimental setup for adsorption in a fixed-bed column.
Figure 2. Experimental setup for adsorption in a fixed-bed column.
Processes 12 01547 g002
Figure 3. Influence of zeolite particle size on CO2 dynamic adsorption capacity (a) zeolite 5A and (b) zeolite 13X.
Figure 3. Influence of zeolite particle size on CO2 dynamic adsorption capacity (a) zeolite 5A and (b) zeolite 13X.
Processes 12 01547 g003
Figure 4. Effect of activation temperature (a) and time (b) on CO2 adsorption capacity.
Figure 4. Effect of activation temperature (a) and time (b) on CO2 adsorption capacity.
Processes 12 01547 g004
Figure 5. Effect of adsorption temperature on CO2 adsorption capacity.
Figure 5. Effect of adsorption temperature on CO2 adsorption capacity.
Processes 12 01547 g005
Figure 6. Effect of CO2 concentration on adsorption capacity.
Figure 6. Effect of CO2 concentration on adsorption capacity.
Processes 12 01547 g006
Figure 7. Effect of acid modification on SiO2/Al2O3 and content of CaO and Na2O in zeolites (a) 5A-HCl, (b) 13X-HCl, (c) 5A-HF, and (d) 13X-HF.
Figure 7. Effect of acid modification on SiO2/Al2O3 and content of CaO and Na2O in zeolites (a) 5A-HCl, (b) 13X-HCl, (c) 5A-HF, and (d) 13X-HF.
Processes 12 01547 g007
Figure 8. Effect of acid modification on dynamic breakthrough curves of CO2 (a) 5A-HCl, (b) 13X-HCl, (c) 5A-HF, and (d) 13X-HF.
Figure 8. Effect of acid modification on dynamic breakthrough curves of CO2 (a) 5A-HCl, (b) 13X-HCl, (c) 5A-HF, and (d) 13X-HF.
Processes 12 01547 g008
Figure 9. Effect of regeneration times on CO2 adsorption capacity.
Figure 9. Effect of regeneration times on CO2 adsorption capacity.
Processes 12 01547 g009
Figure 10. Effect of relative humidity on adsorption of CO2 (a) zeolite 5A and (b) zeolite 13X.
Figure 10. Effect of relative humidity on adsorption of CO2 (a) zeolite 5A and (b) zeolite 13X.
Processes 12 01547 g010
Figure 11. Effect of water vapor on adsorption of CO2 (a) zeolite 5A and (b) zeolite 13X.
Figure 11. Effect of water vapor on adsorption of CO2 (a) zeolite 5A and (b) zeolite 13X.
Processes 12 01547 g011
Figure 12. Effect of water vapor on fixed bed temperature.
Figure 12. Effect of water vapor on fixed bed temperature.
Processes 12 01547 g012
Figure 13. Effect of SO2 on CO2 breakthrough curves (a) zeolite 5A and (b) zeolite 13X.
Figure 13. Effect of SO2 on CO2 breakthrough curves (a) zeolite 5A and (b) zeolite 13X.
Processes 12 01547 g013
Table 1. Chemical composition of zeolites.
Table 1. Chemical composition of zeolites.
Zeolite (wt. %)Fe2O3Al2O3SiO2CaOMgONa2OK2O
5A0.9723.7535.7511.842.082.950.30
13X1.2221.3337.790.942.0513.390.26
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

Yang, X.; Wang, Q.; Chen, J.; Liu, H.; Xu, L.; Rao, M. Competitive Adsorption of Moisture and SO2 for Carbon Dioxide Capture by Zeolites FAU 13X and LTA 5A. Processes 2024, 12, 1547. https://doi.org/10.3390/pr12081547

AMA Style

Yang X, Wang Q, Chen J, Liu H, Xu L, Rao M. Competitive Adsorption of Moisture and SO2 for Carbon Dioxide Capture by Zeolites FAU 13X and LTA 5A. Processes. 2024; 12(8):1547. https://doi.org/10.3390/pr12081547

Chicago/Turabian Style

Yang, Xiduan, Qishuai Wang, Jing Chen, Huibo Liu, Liangping Xu, and Mingjun Rao. 2024. "Competitive Adsorption of Moisture and SO2 for Carbon Dioxide Capture by Zeolites FAU 13X and LTA 5A" Processes 12, no. 8: 1547. https://doi.org/10.3390/pr12081547

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