Next Article in Journal
Analytical Continuation within the Freundlich Adsorption Model. Comment on Edet, U.A.; Ifelebuegu, A.O. Kinetics, Isotherms, and Thermodynamic Modeling of the Adsorption of Phosphates from Model Wastewater Using Recycled Brick Waste. Processes 2020, 8, 665
Previous Article in Journal
Bi-Functional Catalyst/Sorbent for a H2-Rich Gas from Biomass Gasification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Static and Dynamic Simulation of Single and Binary Component Adsorption of CO2 and CH4 on Fixed Bed Using Molecular Sieve of Zeolite 4A

by
Supatsorn Parinyakit
1 and
Patcharin Worathanakul
1,2,*
1
Department of Chemical Engineering, Faculty of Engineering, King Mongkut’s University of Technology North Bangkok, 1518 Pracharat 1, Wongsawang, Bangsue, Bangkok 10800, Thailand
2
Center of Eco-Materials and Cleaner Technology (CECT), Science and Technology Research Institute, King Mongkut’s University of Technology North Bangkok, 1518 Pracharat 1, Wongsawang, Bangsue, Bangkok 10800, Thailand
*
Author to whom correspondence should be addressed.
Processes 2021, 9(7), 1250; https://doi.org/10.3390/pr9071250
Submission received: 16 May 2021 / Revised: 9 July 2021 / Accepted: 16 July 2021 / Published: 20 July 2021
(This article belongs to the Special Issue Materials and Processes for Carbon Capture by Means of Adsorption)

Abstract

:
The simulation of carbon dioxide (CO2)-methane (CH4) mixed gas adsorption and the selectivity on zeolite 4A using Aspen Adsorption were studied. The influence of temperature ranging from 273 to 343 K, pressure up to 10 bar and various compositions of CO2 in the binary system were simulated. The findings of the study demonstrate that the models are accurate. In addition, the effects of various key parameters such as temperature, pressure, and various compositions of binary gases were investigated. The highest CO2 and CH4 adsorption are found at 273 K and 10 bar in the Langmuir isotherm model with 5.86 and 2.88 mmol/g, respectively. The amount of CO2 adsorbed and the selectivity of the binary mixture gas depends on the composition of CO2. The kinetics of adsorption for pure components of CO2 at high temperatures can reach saturation faster than CH4. The influence of the physical properties of zeolite 4A on kinetic adsorption were also studied, and it was observed that small adsorbent particles, large pore diameter, and large pore volume would enter saturation quickly. The prediction of CO2-CH4 mixed gas adsorption and selectivity on zeolite 4A were developed for further use for commercial gas separation.

Graphical Abstract

1. Introduction

Carbon dioxide (CO2) and methane (CH4) are the main components of greenhouse gases that affect global warming. CO2 emissions primarily involve the burning of fossil fuels [1]. Therefore, carbon dioxide trapping and storage also reduces CO2 emissions into the atmosphere, and the stored CO2 can be used for various benefits. The purity of CO2 is often used directly in the food industry and enhanced oil recovery. There are new chemical and biological transformations of CO2 into a feedstock for the manufacturing of chemicals and materials such as organic chemistry, minerals, and polymers. Conversion of CO2 into polymers is one of the added value methods for CO2 applications, CO2 is used to copolymerize with various monomers [2]. Therefore, improving the purity of CO2 from the burning of fossil fuels through various processes is important. Improving of CH4 purity produced from natural gas, fermenting organic matter, and coal and natural gas refining by removing contaminant gases such as CO2, O2, N2, H2S, and H2O is also important. Pure CH4 is important for industries such as pulp and paper manufacturing, food processes, and petroleum refineries. In addition, CH4 is an ingredient in various materials such as, fabric, antifreeze, and fertilizer [3]. The purity of CO2 and CH4 can be accomplished effectively through pressure swing adsorption (PSA) or temperature swing adsorption (TSA).
The CO2 and CH4 gas separation process can be performed through various methods, including distillation, extraction, membrane separation, and adsorption [4,5,6,7]. The chemical industry realized the sustainable development of new innovative processes that use energy and materials more efficiently, since separation processes are of great economic importance accounting for 40–60% of operating costs in the industry. Therefore, a separation process must be developed to save energy consumption and costs. Gas separation by adsorption technology with effective adsorbents is the most common method in the chemical industry [8]. The adsorption process of the components of a fluid mixture flowing through the packed bed of an adsorbent porous material with a large surface area must be considered for proper separation. Different characteristics of the adsorbent influence their applications and these characteristics are influenced by the preparation methods. The most common adsorbents used for the purification of CO2/CH4 purposes are activated carbon, alumina, silica gel, and zeolites [9,10,11,12]. This research chose the adsorbent zeolite because of its high surface area and high adsorption capacity.
Zeolites are crystalline aluminosilicates of alkaline metals, alkaline earth metals, or other cations with various porous characteristics [13]. Zeolites can be used for a variety of applications depending on the pore structure and the properties of each type of zeolite. They can be used for separation or purification gas processes because of their molecular sieve property. Zeolite has a specific cation for the ion exchange process. Therefore, zeolite is an important adsorbent with specific properties for CO2 adsorption because of the high adsorbent surface area and the medium and small pores such as synthesized zeolite A, zeolite X, and zeolite Y [14,15].
Moreover, the development in the simulation process is important to help save cost, energy, and time. In addition, it can help to reduce environmental pollution due to the suitable conditions found in the simulation models. Aspen Adsorption is a simulator developed for the design of the adsorption process simulation, covering a wide range of adsorption conditions. Therefore, this research chose Aspen Adsorption for the simulation of the adsorption processes. The molecular simulation of the adsorption processes can also be simulated using GCMC simulation. Chong and Myshakin [16] studied the molecular simulations of the competitive adsorption of the CO2-CH4 mixture on illitic clay surfaces under dry conditions, which showed that CO2 was specially adsorbed on illitic surfaces and possessed the ability to promote methane desorption. The gas mixture adsorption isotherms of CO2:CH4 indicated that the CO2 concentration increase, caused consistent suppression of CH4 adsorption [16]. However, the adsorption simulation can use mathematical models for prediction breakthrough curves for adsorption. Al Mesfer [17] studied the simulation for CO2 adsorption from CO2-N2 mixture using a packed column on activated carbon, which illustrated that breakthrough time decreased with increasing temperature, feed rate, and increased CO2 concentration [17]. The experimental results were compared to simulate mathematical models of breakthrough curves with high accuracy [18]. Until now, there were no studies for CO2 adsorption from CO2-CH4 mixed gas using zeolite 4A, which were reviewed.
This computer simulation research through Aspen Adsorption aimed for CO2 adsorption from CO2-CH4 mixed gas using zeolite 4A and the optimum conditions for adsorption. Therefore, the main objectives of this research were to study the factors affecting CO2 and CH4 adsorption and the selectivity of binary gas mixtures on zeolite 4A at temperatures ranging from 273 to 343 K, a pressure range of 0–10 bar, and the compositions of CO2 in the binary system were 10:90, 30:70, 40:60, 50:50, 60:40, and 80:20.

2. Materials and Methods

2.1. Simulation Setting

Data for program modeling included the physical properties of the adsorbent tower, the adsorbent and the conditions of operation. For this research, the adsorbent material was zeolite 4A and CO2 and CH4 were adsorbates. The physical properties of the adsorbents and the column for the packing tower are shown in Table 1. The characterization of zeolite 4A and experimental data are from Seabra and coworkers [19]. This research used the results from the experiment of adsorption of pure CO2 and CH4 on zeolite 4A at 303 and 343 K to predict the adsorption at 273, 283, and 323 K with the program Aspen Adsorption V11.
The zeolite 4A diameter was in the range of 0.16–0.25 cm in granular form as shown in Table 1. The pellet density and total intrusion volume were obtained by mercury intrusion and the solid density was obtained by helium picnometry. The BET surface area was measured from the N2 adsorption at 77 K in granule form [19].

Dynamic Simulation of Adsorption Experiments

Dynamic adsorption experiments were simulated by the Aspen Adsorption V11 program. The flow sheet including adsorbent characteristics (gas bed), feed gas, and product streams is shown in Figure 1. The internal diameter bed was 9.1 mm and the packing length was 97.5 mm in the column. The gas bed model configuration allowed for specifying the number of layers, labels for each component and setting the geometry of the bed. There were a set of assumptions for all layers, constant variables, and initial conditions. General assumptions of Aspen Adsorption including partial differential equation (PDE) discretization method were used to approximate spatial derivatives and number PDE nodes with 20 nodes. Material balance was assumed with convection only, momentum balance was laminar, and turbulent flow conditions during operation used the Ergun equation. The kinetic model was set as linear lumped resistance (LDF) and the mass transfer coefficient was constant. The isotherm for the pure components was determined using the Langmuir isotherm. The simulation model was calculated in the multicomponent with an extended Langmuir model based on partial pressure. Energy balance for this research was defined for non-isothermal conditions with no conduction. The heat transfer to the environment was set as adiabatic (no external heat transfer) [20,21].

2.2. Adsorption Isotherm Model

2.2.1. Pure Component Adsorption Isotherm

Isotherms of pure CO2 and CH4 at 273, 283, 303, 323 and 343 K and pressure ranging from 1 to 10 bar were studied using the Langmuir isotherm model. This simulation was compared to the results of Seabra and coworkers [19].
The Langmuir isotherm model is a replica of the easiest and most popular isotherm model as shown in Equation (1). This model was used for monolayer adsorption and physical adsorption [22].
q = q m b P 1 + b P
where q (mmol/g) is the amount adsorbed, qm (mmol/g) is the maximum adsorption capacity of the adsorbent, P (bar) is the pressure and b (bar−1) is the Langmuir constant. The Langmuir constant depends on the temperature of the system, represented by Equation (2).
b = b 0 e Δ H R T
where b0 (bar−1) is the adsorption constant at an infinite temperature, −ΔH (J/mmol) is the heat of adsorption, R (J⋅K−1⋅mol−1) is the universal gas constant, and T (K) is the temperature of the system.

2.2.2. Binary Component Adsorption Isotherm

Adsorption equilibrium for the binary gas mixture between CO2 and CH4 was predicted. The temperatures were at 273, 283, 303, 323, and 343 K with pressure ranging from 1 to 10 bars. The compositions of CO2 to CH4 were 10:90, 30:70, 40:60, 50:50, 60:40, and 80:20 using an extended Langmuir (EL) isotherm model and the ideal adsorbed solution theory (IAST). The extended Langmuir isotherm model for a binary component or multicomponent adsorption was developed from the Langmuir model. In pure components, EL used the information of adsorption from Equation (3) [23].
q i = q m , i b i P i 1 + Σ b i P i
where qi (mmol/g) is the amount adsorbed of component i, qm (mmol/g) is the maximum adsorption capacity of the adsorbent of component i, Pi (bar)is the partial pressure of component i and b (bar−1) is the Langmuir constant of component i. Therefore, the EL model for binary components is shown in Equations (4) and (5).
For binary component:
q 1 = q m , 1 b 1 P 1 1 + b 1 P 1 + b 2 P 2
q 2 = q m , 2 b 2 P 2 1 + b 1 P 1 + b 2 P 2
IAST is used to predict the adsorption capacity of binary mixed gas using pure component data. IAST is a thermodynamic method based on the adsorption equilibrium with Raoult’s law for vapor-liquid equilibrium. The equilibrium between adsorbed phase and ideal gas phase can be specified by Equation (6) [24,25,26].
P y i = P i 0 ( π * ) x i
where yi and xi are the molar fractions of component i in the gas phase and adsorbed phase, respectively. P (bar) is the total pressure of the mixture, and P i 0 ( π * ) (bar) is the equilibrium gas phase pressure of pure component i corresponding to solution temperature and solution spreading pressure, π*.
For a pure component i, the spreading pressure using Equations (7) and (8) was followed:
π * = π i A R T = 0 P i 0 q i P i d P i
where πi* is the reduced spreading pressure of component i in the adsorbed phase, πi is the spreading pressure of component i in the adsorbed phase, A is the specific surface area of the adsorbent, qi is the pure component adsorption isotherm equation, and P i 0 is the standard state pressure of pure component i corresponding to spreading pressure of the mixture.
At the standard state, reduced spreading pressure of the mixture (π*) is the same as the reduced spreading pressure of a single component according to Equation (8).
π 1 * = π 2 * = = π *
For binary mixtures, Equations (6)–(8) were solved numerically and the total adsorbed amount was calculated by Equations (9)–(11).
1 q T = x 1 q 1 ( p 1 0 ) + x 2 q 2 ( p 2 0 )
x 1 + x 2 = 1
q i = x i q T
where qT (mmol/g) is the total amount adsorbed.

2.3. Modeling of Mass Transfer Coefficient

Mass transfer coefficients (MTC) were shown in Equation (12) which was assumed to be constant. MTC included the effects of micropore, macropore and film resistances [27]. The effect of macropore was considered using the following equation.
k i = 15 ε p D p i R p 2
where ki (s−1) is the overall mass transfer coefficient of species i, Dpi (cm2/s) is the macropore diffusivity of species i, and Rp (cm) is the particle radius and εp is the porosity of adsorbent particle or intraparticle.
The effective macropore diffusivity can be determined using the Bosanquet equation [28]:
1 D p i = τ ( 1 D k i + 1 D m i )
where τ is the pore tortuosity factor, Dki (cm2/s) is the Knudsen diffusivity and Dmi (cm2/s) is the molecular diffusivity.
Estimation of the molecular diffusivity of binary gas mixtures with the best method calculated from the Lennard-Jones equation represented using Equation (14) [28]. The molecular diffusivity was determined using the following equation.
D m = 0.001858 T 3 2 P σ 12 2 Ω D { 1 M 1 + 1 M 2 }
where Mi (g/mol) is the molecular weight of species i, ΩD is the collision integral and σ12 (Å) is the collision diameter of the binary pair of species A and B.
Knudsen diffusion is gas diffusion through small pores, which can be calculated using Equation (15).
D k i = 9700 R P T m i
where RP (cm) is pore radius.

2.4. Selectivity of CO2 over CH4 in Binary Mixture Gas

Selectivity represents the ratio of the amount of adsorption of the two gases. It can also be called separation coefficient. If the selectivity of CO2/CH4 is high, it means that the adsorption amount of CO2 is greater than CH4 [10].
The adsorption selectivity of CO2 over CH4 in binary mixtures was defined in Equation (16).
S C O 2 / C H 4 = q C O 2 q C H 4 P C O 2 P C H 4
where q C O 2 and q C H 4 (mmol/g) are the amount adsorbed of CO2 and CH4, and P C O 2 and P C H 4 (bar) are the partial pressures of CO2 and CH4, respectively.

2.5. Breakthrough Curves Modeling

To assess the performance of the fixed-bed adsorption column and measure the breakthrough curves, it is necessary to design and utilize the lab-scale experimental setup. By optimizing the mathematical models to the measured experimental data, the useful information can be used to design large-scale industrial columns and to predict the practical conditions. The model used in this work are the Thomas model and the Yoon–Nelson model.

2.5.1. Thomas Model

The Thomas model [29] is one of the most commonly used in the prediction of the breakthrough curve and the describing of the column performance. This model was developed based on the Langmuir kinetics of adsorption that assumed negligible axial dispersion in the column adsorption. The rate of the driving force carries out the second-order reversible reaction kinetics [30]. The Thomas model is shown in Equation (17):
C C 0 = 1 1 + exp ( k T h q 0 M Q k T h C 0 t )
where q0 is the equilibrium adsorbate uptake in adsorbent (mg/g), Q is the flow rate (mL/min), M is the mass of the adsorbent (g), C is effluent concentration (mg/L), C0 is influent concentration (mg/L), t is time (min), and kTh is the Thomas model constant (mL/min·mg).

2.5.2. Yoon-Nelson Model

The Yoon-Nelson model [31] is based on the assumption that the rate of decrease in the probability of adsorption for each adsorbate is proportional to the probability of adsorbate breakthrough on the adsorbent [32]. The Yoon-Nelson model was used for the following equation.
C C 0 = exp ( k Y N t τ k Y N ) 1 + exp ( k Y N t τ k Y N )
where kYN is the Yoon-Nelson constant (min−1) and τ is the time required to reach the effluent concentration to 50% of the influent concentration (min).

3. Results and Discussion

3.1. Pure Component Adsorption Isotherm

Adsorption isotherm of pure CO2 and pure CH4 at different temperatures and pressures are shown in Figure 2. The experimental data for pure CO2 and CH4 adsorption at 303 and 343 K were from Seabra [19]. The experimental results are fitted with the Langmuir isotherm model in Equations (1) and (2).
The adsorption capacity of CO2 and CH4 decreased with the rising temperature, indicating the adsorption of CO2 and CH4 are exothermic physical. The type of physical and chemical adsorption depends on the amount of heat in the adsorption. For heat of adsorption with 80 kJ·mol−1 or more, the adsorption process indicates chemisorption, while lower values represent a physical adsorption [33]. Values of the isosteric heat of adsorption in zeolite 4A of 47.8 kJ mol−1 for the adsorption of CO2 [34] and 16.72 kJ·mol−1 for CH4 [35] were found. Thus, the adsorption of CO2 and CH4 on zeolite 4A was physical adsorption; it could adsorb well when the temperature is decreased. In fact, as temperature decreases, gas molecules have less kinetic energy because the bond between gas and adsorbent is increased [36,37]. Moreover, the effect of pressure on the adsorption capacity is shown in Figure 2. The adsorption capacity increased rapidly as the pressure increased due to an increase in the gas molecules hitting the surface. Therefore, the increase in pressure caused the adsorption rate to increase linearly. However, when the pressure became high and almost the entire surface of the adsorbent received saturated gas, the pressure had little effect on the adsorption capacity. Ultimately, it could reach a point where the pressure did not affect the adsorption capacity because the number of adsorption sites was fixed, and no more adsorption occurred in those sites. At the same pressure and temperature conditions, the adsorption capacity of CO2 is much higher than CH4 because CO2 has a quadrupole moment and polarizability greater than CH4, and it also has a high critical temperature as shown in Table 2 [38,39].
Table 3 shows the parameters of the Langmuir isotherm model for CO2 and CH4 adsorption on zeolite 4A. Where qm,0 is maximum adsorption capacity, b0 is Henry law constant, Q/R is adsorption heat and X is empirical constant. The constant values in Table 3 were used in the Langmuir isotherm model in Equations (1) and (2).
Parameters of the previously mentioned equations were determined by minimizing the root-mean-square (RMS) in Equation (19):
RMS = ( i = 1 N ( q i c a l q i e x p ) 2 N ) 0.5
where N is the number of data points and q i c a l and q i e x p are calculated and experimental adsorbed amounts, respectively. Low RMS value indicates that the Langmuir isotherm model is suitable.
The Langmuir isotherm model depends on pressure, maximum adsorption capacity of the adsorbent, and the Langmuir constant. Maximum adsorption capacity and the Langmuir constant depend on temperature. Therefore, the amount of adsorption for each pressure and temperature could be determined [22].
The crystal structure of zeolite 4A with sodium cation distribution as shown in Figure 3. There are three sites for sodium cation distribution including site I (S1) at the center of the 6-rings of sodalite cages, site II (S2) at the center of the 8-ring window of α cages and site III (S3) at opposite the 4-rings on the interior of α cages. Sodium ions in zeolite 4A contains 12 ions per unit cell (S1:S2:S3 = 8:3:1) [40,41].
The adsorption mechanism of the CO2 molecule on zeolite 4A showed that CO2 interacts with the sodium cations in the adsorption site. Interaction between the CO2 quadrupole and sodium cations was electrostatic interaction. Sites of CO2 adsorption interaction with the sodium cation were shown in a single cation site. CO2 interacted with two sodium cations at dual cation sites. Moreover, CO2 could interact with more than two sodium cations and was denoted as multiple cation sites [42,43]. For instance, interaction between CO2 and zeolite 4A was shown in Figure 4. The CO2 molecule interacted with the sodium cation in S1 perpendicular to the plane of the 6-rings of sodalite cages along the body diagonal. If the distance between the CO2 molecule and sodium cation is long distance, it indicates a weak interaction.
Likewise, CH4 interacted with sodium cation the same as CO2. However, the distance or molecular arrangement may differ due to the different properties of CO2 and CH4 as shown in Table 2. The CH4 has zero quadrupole moment and polarizability less than CO2. Therefore, the electrostatic affinity and adsorption capacity for CH4 were less than CO2. For the distance between the gas molecule and the cation, the longer distance indicated a weak interaction. In addition, CO2 and CH4 molecules could interact with oxygen atoms of the framework as well.

3.2. Binary Component Adsorption Isotherm

The simulation was predicted for a binary gas mixture using extended Langmuir Equations (4) and (5) and the ideal adsorbed solution theory models used in Equations (9)–(11). The composition of CO2 and CH4 in the gas mixture affected the amount of adsorption. The predictions of the adsorption of the binary gas mixture between CO2 and CH4 on zeolite 4A with the EL isotherm model were shown in Figure 5a,b.
The results showed that the total amount of adsorbed gas mixture increased with the amount of CO2 adsorption. In other words, the higher the composition of CO2 received the greater the amount of total gas adsorption. Thus, the adsorption capacity of the mixed gas was between the two pure gases due to gas mixture had competition and was a hinderance between CO2 and CH4 molecules in the adsorption. It was indicated that the quadrupole moment of CO2 could result in strong interactions between CO2 molecules and the surface of zeolite 4A. The effect of pressure and temperature on the adsorption of the binary gas mixture showed the same tendency as the pure component [44,45,46]. Therefore, the adsorption of the mixed gas was also a physical adsorption because the total amount adsorbed decreased with the rising temperature.
Zeolite 4A has a cubic structure as shown in Figure 3. The effective size of its windows depends on the sodium cation of zeolite 4A, which has a pore window size of approximately 0.38 nm. The kinetic diameter affected the separation of CO2 from CH4 as observed in theory. These pores of zeolite have dimensions very close to the kinetic diameters of CO2 and CH4, allowing CO2 to diffuse through the adsorbent faster than CH4. Therefore, CO2 can be separated from CH4 while CO2 diffuses more quickly in narrow pores than CH4 with a kinetic diameter effect. For the molecular sieve effect, both CO2 and CH4 have different kinetic diameters inside a zeolite as shown in Table 2. CO2 has the smallest kinetic diameter at 0.33 nm, and CH4 at 0.38 nm. Zeolite 4A showed pore window apertures that are similar to the kinetic diameter of CH4. Therefore, CO2 could enter the zeolite 4A freely, but CH4 was blocked [47].
Figure 6 showed the comparison between the EL and IAST models of the CO2-CH4 mixture in different CO2 and CH4 ratios. The total adsorption of IAST is a little higher than the EL model. The total adsorption between IAST and EL models was approximately the same. The IAST model showed better adsorption of the CO2-CH4 gas mixture than the EL model compared to the experimental results according to the study of Rios [48]. In Wu’s research [39], IAST could be used to predict the behavior of a binary mixture with very high accuracy. The IAST model was able to work very well when the adsorbates were similar sizes. On the other hand, the EL model was able to predict sorption behavior with acceptable precision.

3.3. Selectivity for Separating of CO2-CH4 Mixture

The selectivity of binary gas mixtures can be calculated from Equation (16). The selectivity at different compositions of CO2:CH4 and pressures shown in Figure 7a at 303 K. If the composition of CO2 increased, the selectivity also increased because the interaction of CO2 molecules with the atoms of the zeolite 4A structure was stronger than that with CH4 molecules. In addition, CO2 molecules particularly adsorbed well in the pores of zeolite 4A and hindered the diffusion of the weaker adsorbing CH4 molecules. The obtained selectivity from IAST and EL models for different compositions are shown in Figure 7a. It was indicated that the EL selectivity of CO2/CH4 was constant for all gas compositions and pressures. On the other hand, the IAST selectivity showed various results for both total pressure and composition. From this prediction of IAST, the composition of CO2 would affect selectivity when the pressure was increased. Figure 7b shows the different temperatures on selectivity of 50:50 of CO2:CH4 ratio, at 1 bar. The selectivity increased with rising temperature [49] which is the same in both models. For this reason, the pressure and temperature had a positive impact on the adsorption selectivity of CO2 over CH4. Therefore, to predict the selectivity of CO2-CH4, the IAST calculation was based on Langmuir and EL calculations.
As mentioned above, the use of an adsorbent must be considered for high efficiency separation. Therefore, the comparison of the adsorption capacity in each adsorbent can be used as analytical data for improving the adsorbent.
Table 4 shows the adsorption capacity for the CO2-CH4 binary gas mixture with zeolite compared to others. It can be seen that the adsorption values of both CO2 and CH4 from zeolite 4A showed similar trends to those of 13 X and 5A zeolites at the same conditions. The BET specific surface area of zeolite 4A was close to that of zeolite 13X. Therefore, the adsorption capacities of CO2 were similar to CH4. There were no different adsorption effects in each zeolite in Table 4. The adsorption capacity of CO2 showed greater than CH4. The selectivity of the binary gas mixture has an important parameter because it can form zeolite into an ideal material adsorbent for CO2 and CH4 mixture gas separation.

3.4. Dynamic Simulation of Adsorption Experiment

Figure 8 shows the kinetic adsorption of CO2 and CH4 on zeolite 4A at different temperatures (273, 303, 343 K) and pressures (1, 5, 10 bar).
It was observed that at the beginning of the adsorption, the amount of CO2 adsorbed on the adsorbent was slightly fast and then slowly decreased until it reached equilibrium. In the initial stages, CO2 molecules directly contacted with the adsorbent, resulting in great interaction between adsorbate and adsorbent [50]. After total pores were adsorbed without any further adsorption of CO2 molecules, the process of adsorption went to saturation. In addition, when the temperature increased, faster saturation was observed due to the exothermic process for CO2 adsorption. In the exothermic process, increasing temperature caused adsorption to decrease because of the decreased of the attraction between the adsorbate and the adsorbent [37,39,44].
The mass transfer coefficient was increased with increased temperature as shown in Table 5. This is caused by CO2 and CH4 molecules moving faster with higher temperature from the increased kinetic energy [28,51].
From the simulation adsorption model, it was found that pressure affected the adsorption capacity and mass transfer coefficient. The mass transfer coefficient decreased with the rising pressure because of effective diffusivity decreased from decreased molecular diffusion [28,51]. The relation of mass transfer coefficient, effective diffusivity and molecular diffusivity are shown in Equations (12)–(14), respectively. Equation (15) shows Knudsen diffusion which depends on temperature. Figure 8b shows that CH4 adsorption took a long time to adsorb, due to the decrease in the polarizability and kinetic diameter of CH4, which was larger than CO2 as shown in Table 2.

3.4.1. The Effect of the Physical Properties of Zeolite 4A on Kinetic Adsorption

The physical properties of various zeolite 4A with different particle size, pore volume and pore diameter are shown in Table 6. Three types of zeolites were compared in this study: zeolite 4A-0.2 cm, zeolite 4A (HSD, high bulk density)-0.2 cm and zeolite 4A-0.4 cm.
To study the physical properties of the zeolite 4A adsorbents, the model developed to predict the breakthrough curves of CO2 and CH4 adsorptions on zeolite 4A was shown in Figure 9. It was found that small particle sizes of zeolite went into saturation and balance more quickly than large ones because of long diffusion inside the pores [52]. For different pore volumes with the same particle size, at large pore volumes could go into saturation quickly because the adsorption capacity was greater for larger pore volumes. Moreover, after all the pores were occupied by the adsorbate, the adsorbent could no longer adsorb CO2 molecules. The efficiency was high if the amount of porosity was large. Less pore diameter affects the kinetic adsorption on fast diffusion. The effect of pore diameter on the adsorption is mentioned in Section 3.2.
The mass transfer coefficient of large particles was smaller than the small ones as shown in Table 7. From Equation (12), the mass transfer coefficient was inverse to the particle radius. Therefore, the particle radius was large; the mass transfer coefficient was reduced.

3.4.2. The Effect of Binary CO2-CH4 Mixed Gas

Figure 10a,b showed the effect of the composition ratios of CO2 and CH4 of 80:20, 50:50 and 30:70 at 303 K and 10 bar pressure.
When CO2 composition increased, the saturation of the breakthrough curve was decreased, representing the faster kinetics of the adsorption process with a high CO2 content [53]. It was predicted that CO2 had a higher adsorption with zeolite 4A than CH4. As a result, when CO2 composition was high, competition of CO2 and CH4 for adsorption was less than with less composition of CO2 [54,55].

3.4.3. Modeling of Breakthrough Curves

Modeling of the breakthrough curves were obtained from experiments using the Thomas and Yoon-Nelson model. The experimental data of the adsorption of CO2 on zeolite 4A at 573.15 K and flow rate of 5 L/h were obtained from Tobarameekul [56].
Figure 11 shows the ability of the Thomas and Yoon-Nelson model to predict the experimental breakthrough curves. The prediction of the Yoon-Nelson model is better than Thomas’s model. Both models were formed according to the experimental data. The parameters from fitting the different models to experimental breakthrough curves were shown in Table 8. The correlation coefficient (R2) of the Yoon-Nelson model is greater than that of Thomas.

4. Conclusions

Adsorption isotherms of pure CO2 and pure CH4 on zeolite 4A with 273 to 343 K, and pressure up to 10 bar using the Langmuir model were performed. Adsorption generally depended on the temperature. Adsorption decreased with increasing temperature because adsorption processes were exothermic reaction. On the other hand, at a constant temperature, the adsorption capacity increased with pressure. Therefore, the highest CO2 and CH4 adsorption from this study was found at 273 K and 10 bar. The properties of CO2 and CH4 affected on the adsorption capacity so that the adsorption capacity of CO2 was much higher than CH4. The effect of temperature and pressure on the binary gas mixture had the same effect on the pure component of adsorption. However, the adsorption of the mixed gas increased with the amount of CO2 entered. The effect of pore size on adsorption showed that CO2 with a smaller kinetic diameter could be separated from CH4 with a larger kinetic diameter as CO2 diffuses more quickly in narrow pores than CH4. In addition, Zeolite 4A has pore window apertures that are similar to the kinetic diameter of CH4, then CO2 could enter the zeolite 4A freely, but CH4 was blocked. Simulation models for gas mixtures were IAST and EL models. The amount of adsorption of the IAST model was greater than the EL model and the selectivity also increased with the amount of CO2 entered, and the results showed that selectivity rose with the temperature. Moreover, the amount of CO2 adsorbed from dynamic adsorption simulation increased with increasing pressure because of effective diffusivity decreased from decreased molecular diffusion. At the same time, CH4 showed the same trend as CO2, but its adsorption capacity was less than CO2. In addition, the rising temperature could reach the equilibrium faster than the low temperature. Small adsorbent particles and large pore volume could enter the saturation fast. For the kinetic adsorption simulation of the CO2-CH4 binary mixture gas adsorption, the increased composition of CO2 would greatly benefit the efficiency of CO2 adsorption in the mixed gas system. Moreover, the prediction of the breakthrough curves from the Yoon-Nelson model was better than Thomas’s model.

Author Contributions

Conceptualization, S.P. and P.W.; methodology, S.P. and P.W.; software, S.P.; validation, S.P. and P.W.; formal analysis, S.P. and P.W.; investigation, P.W.; data curation, S.P. and P.W.; writing—original draft preparation, S.P.; writing—review and editing, P.W.; visualization, S.P.; supervision, P.W.; project administration, P.W.; funding acquisition, P.W. Both authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by King Mongkut’s University of Technology North Bangkok. Contract no. KMUTNB-63-KNOW-043.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jawaharraj, K.; Shrestha, N.; Chilkoor, G.; Dhiman, S.S.; Islam, J.; Gadhamshetty, V. Valorization of methane from environmental engineering applications: A critical review. Water Res. 2020, 187, 116400. [Google Scholar] [CrossRef]
  2. Paraschiv, S.; Paraschiv, L.S. Trends of carbon dioxide (CO2) emissions from fossil fuels combustion (coal, gas and oil) in the EU member states from 1960 to 2018. Energy Rep. 2020, 6, 237–242. [Google Scholar] [CrossRef]
  3. Qin, Y.; Wang, X. Conversion of CO2 into Polymers. In Green Chemistry and Chemical Engineering; Han, B., Wu, T., Eds.; Springer: New York, NY, USA, 2019; pp. 323–347. [Google Scholar]
  4. Yousef, A.M.; El-Maghlany, W.M.; Eldrainy, Y.A.; Attia, A. Low-Temperature Distillation Process for CO2/CH4 Separation: A Study for Avoiding CO2 Freeze-Out. J. Heat Transf. 2017, 140. [Google Scholar] [CrossRef]
  5. Qin, C.; Jiang, Y.; Zhou, J.; Song, X.; Liu, Z.; Li, D.; Zhou, F.; Xie, Y.; Xie, C. Effect of supercritical CO2 extraction on CO2/CH4 competitive adsorption in Yanchang shale. Chem. Eng. J. 2021, 412, 128701. [Google Scholar] [CrossRef]
  6. Zhang, Y.; Sunarso, J.; Liu, S.; Wang, R. Current status and development of membranes for CO2/CH4 separation: A review. Int. J. Greenh. Gas. Control. 2013, 12, 84–107. [Google Scholar] [CrossRef]
  7. Klewiah, I.; Berawala, D.S.; Alexander Walker, H.C.; Andersen, P.Ø.; Nadeau, P.H. Review of experimental sorption studies of CO2 and CH4 in shales. J. Nat. Gas. Sci. Eng. 2020, 73, 103045. [Google Scholar] [CrossRef]
  8. Wankat, P.C. A review of: “Gas Separation Technology”. Sep. Purif. Methods 1991, 20, 199–200. [Google Scholar] [CrossRef]
  9. Álvarez-Gutiérrez, N.; Gil, M.V.; Rubiera, F.; Pevida, C. Adsorption performance indicators for the CO2/CH4 separation: Application to biomass-based activated carbons. Fuel Process. Technol. 2016, 142, 361–369. [Google Scholar] [CrossRef] [Green Version]
  10. Tamnanloo, J.; Fatemi, S.; Golmakani, A. Binary Equilibrium Adsorption Data and Comparison of Zeolites with Activated Carbon for Selective Adsorption of CO2 from CH4. Adsorpt. Sci. Technol. 2014, 32, 707–716. [Google Scholar] [CrossRef]
  11. Zhang, Y.; Chen, K.; Lv, C.; Wu, T.; Wen, Y.; He, H.; Yu, S.; Wang, L. Adsorption Separation of CO2/CH4 from Landfill Gas by Ethanolamine-Modified Silica Gel. WaterAir Soil Pollut. 2021, 232. [Google Scholar] [CrossRef]
  12. Liang, D.; Hu, Y.; Bao, Q.; Zhang, J.; Feng, J.; Sun, P.; Ma, Y.; Zhang, H. A suitable zeolite Rho for separating CO2/CH4 in pressure swing adsorption (PSA) process. Inorg. Chem. Commun. 2021, 127, 108547. [Google Scholar] [CrossRef]
  13. Verma, M.; Kaur, G. A mini review on zeolite. Mater. Today Proc. 2021. [Google Scholar] [CrossRef]
  14. Dusselier, M.; Davis, M.E. Small-Pore Zeolites: Synthesis and Catalysis. Chem. Rev. 2018, 118, 5265–5329. [Google Scholar] [CrossRef]
  15. Li, Y.; Li, L.; Yu, J. Applications of Zeolites in Sustainable Chemistry. Chem 2017, 3, 928–949. [Google Scholar] [CrossRef] [Green Version]
  16. Chong, L.; Myshakin, E.M. Molecular simulations of competitive adsorption of carbon dioxide—Methane mixture on illitic clay surfaces. Fluid Phase Equilibria 2018, 472, 185–195. [Google Scholar] [CrossRef]
  17. Al Mesfer, M.K.; Amari, A.; Danish, M.; Al Alwan, B.A.; Shah, M. Simulation study of fixed-bed CO2 adsorption from CO2/N2 mixture using activated carbon. Chem. Eng. Commun. 2020, 1–10. [Google Scholar] [CrossRef]
  18. Brea, P.; Delgado, J.A.; Águeda, V.I.; Uguina, M.A. Modeling of breakthrough curves of N2, CH4, CO, CO2 and a SMR type off-gas mixture on a fixed bed of BPL activated carbon. Sep. Purif. Technol. 2017, 179, 61–71. [Google Scholar] [CrossRef]
  19. Seabra, R.; Ribeiro, A.; Gleichmann, K.; Ferreira, A.; Rodrigues, A. Adsorption equilibrium and kinetics of carbon dioxide, methane and nitrogen on binderless zeolite 4A adsorbents. Microporous Mesoporous Mater. 2018, 277. [Google Scholar] [CrossRef]
  20. Wood, K.; Liu, Y.; Yu, Y. Simulation of Adsorption Processes; Wiley: Hoboken, NJ, USA, 2018; pp. 1–153. [Google Scholar]
  21. Plaza, M.G.; Durán, I.; Querejeta, N.; Rubiera, F.; Pevida, C. Experimental and Simulation Study of Adsorption in Postcombustion Conditions Using a Microporous Biochar. 1. CO2 and N2 Adsorption. Ind. Eng. Chem. Res. 2016, 55, 3097–3112. [Google Scholar] [CrossRef] [Green Version]
  22. Azizian, S.; Eris, S. Chapter 6—Adsorption isotherms and kinetics. In Interface Science and Technology; Ghaedi, M., Ed.; Elsevier: Amsterdam, The Netherlands, 2021; Volume 33, pp. 445–509. [Google Scholar]
  23. Kapoor, A.; Ritter, J.A.; Yang, R.T. An extended Langmuir model for adsorption of gas mixtures on heterogeneous surfaces. Langmuir 1990, 6, 660–664. [Google Scholar] [CrossRef]
  24. Goel, C.; Bhunia, H.; Bajpai, P. Prediction of Binary Gas Adsorption of CO2 /N2 and Thermodynamic Studies on Nitrogen Enriched Nanostructured Carbon Adsorbents. J. Chem. Eng. Data 2017, 62, 214–225. [Google Scholar] [CrossRef]
  25. LeVan, M.D.; Vermeulen, T. Binary Langmuir and Freundlich isotherms for ideal adsorbed solutions. J. Phys. Chem. 1981, 85, 3247–3250. [Google Scholar] [CrossRef]
  26. McIntyre, S.M.; Shan, B.; Wang, R.; Zhong, C.; Liu, J.; Mu, B. Monte Carlo Simulations to Examine the Role of Pore Structure on Ambient Air Separation in Metal–Organic Frameworks. Ind. Eng. Chem. Res. 2018, 57, 9240–9253. [Google Scholar] [CrossRef]
  27. Hauchhum, S.; Mahanta, P.; De Wilde, J. Capture of CO2 from Flue Gas onto Coconut Fibre-Based Activated Carbon and Zeolites in a Fixed Bed. Transp. Porous Media 2015, 110. [Google Scholar] [CrossRef]
  28. Krishna, R.; van Baten, J.M. Investigating the validity of the Bosanquet formula for estimation of diffusivities in mesopores. Chem. Eng. Sci. 2012, 69, 684–688. [Google Scholar] [CrossRef]
  29. Thomas, H.C. Heterogeneous Ion Exchange in a Flowing System. J. Am. Chem. Soc. 1944, 66, 1664–1666. [Google Scholar] [CrossRef]
  30. Afroze, S.; Sen, T.K.; Ang, H.M. Adsorption performance of continuous fixed bed column for the removal of methylene blue (MB) dye using Eucalyptus sheathiana bark biomass. Res. Chem. Intermed. 2016, 42, 2343–2364. [Google Scholar] [CrossRef]
  31. Yoon, Y.H.; Nelson, J.H. Application of Gas Adsorption Kinetics I. A Theoretical Model for Respirator Cartridge Service Life. Am. Ind. Hyg. Assoc. J. 1984, 45, 509–516. [Google Scholar] [CrossRef] [PubMed]
  32. El-Naas, M.H.; Alhaija, M.A.; Al-Zuhair, S. Evaluation of an activated carbon packed bed for the adsorption of phenols from petroleum refinery wastewater. Environ. Sci. Pollut. Res. 2017, 24, 7511–7520. [Google Scholar] [CrossRef]
  33. Khalili, S.; Khoshandam, B.; Jahanshahi, M. A comparative study of CO2 and CH4 adsorption using activated carbon prepared from pine cone by phosphoric acid activation. Korean J. Chem. Eng. 2016, 33, 2943–2952. [Google Scholar] [CrossRef]
  34. Romero-Pérez, A.; Aguilar-Armenta, G. Adsorption Kinetics and Equilibria of Carbon Dioxide, Ethylene, and Ethane on 4A(CECA) Zeolite. J. Chem. Eng. Data 2010, 55, 3625–3630. [Google Scholar] [CrossRef]
  35. Harper, R.J.; Stifel, G.R.; Anderson, R.B. Adsorption of gases on 4A synthetic zeolite. Can. J. Chem. 1969, 47, 4661–4670. [Google Scholar] [CrossRef]
  36. Wei, M.; Yu, Q.; Duan, W.; Hou, L.; Liu, K.; Qin, Q.; Liu, S.; Dai, J. Equilibrium and Kinetics Analysis of CO2 Adsorption on Waste Ion-exchange Resin-based Activated Carbon. J. Taiwan Inst. Chem. Eng. 2017, 77, 161–167. [Google Scholar] [CrossRef]
  37. Golchoobi, A.; Pahlavanzadeh, H. Molecular simulation, experiments and modelling of single adsorption capacity of 4A molecular sieve for CO2–CH4 separation. Sep. Sci. Technol. 2016, 51, 2318–2325. [Google Scholar] [CrossRef]
  38. Mofarahi, M.; Gholipour, F. Gas adsorption separation of CO2/CH4 system using zeolite 5A. Microporous Mesoporous Mater. 2014, 200, 1–10. [Google Scholar] [CrossRef]
  39. Wu, Y.-J.; Yang, Y.; Kong, X.-M.; Li, P.; Yu, J.-G.; Ribeiro, A.M.; Rodrigues, A.E. Adsorption of Pure and Binary CO2, CH4, and N2 Gas Components on Activated Carbon Beads. J. Chem. Eng. Data 2015, 60, 2684–2693. [Google Scholar] [CrossRef]
  40. Cornelius, M.-L.U.; Price, L.; Wells, S.A.; Petrik, L.F.; Sartbaeva, A. The steric influence of extra-framework cations on framework flexibility: An LTA case study. Z. Krist. Cryst. Mater. 2019, 234, 461–468. [Google Scholar] [CrossRef]
  41. Rzepka, P.; Bacsik, Z.; Smeets, S.; Hansen, T.C.; Hedin, N.; Wardecki, D. Site-Specific Adsorption of CO2 in Zeolite NaK-A. J. Phys. Chem. C 2018, 122, 27005–27015. [Google Scholar] [CrossRef]
  42. Grajciar, L.; Čejka, J.; Zukal, A.; Areán, C.O.; Palomino, G.T.; Nachtigall, P. Controlling the Adsorption Enthalpy of CO2 in Zeolites by Framework Topology and Composition. ChemSusChem 2012, 5, 2011–2022. [Google Scholar] [CrossRef] [PubMed]
  43. Zukal, A.; Arean, C.O.; Delgado, M.R.; Nachtigall, P.; Pulido, A.; Mayerová, J.; Čejka, J. Combined volumetric, infrared spectroscopic and theoretical investigation of CO2 adsorption on Na-A zeolite. Microporous Mesoporous Mater. 2011, 146, 97–105. [Google Scholar] [CrossRef]
  44. Abdul Kareem, F.A.; Shariff, A.M.; Ullah, S.; Mellon, N.; Keong, L.K. Adsorption of pure and predicted binary (CO2:CH4) mixtures on 13X-Zeolite: Equilibrium and kinetic properties at offshore conditions. Microporous Mesoporous Mater. 2018, 267, 221–234. [Google Scholar] [CrossRef]
  45. Pino, D.; Bessières, D. CH4 /CO2 Mixture Adsorption on a Characterized Activated Carbon. J. Chem. Eng. Data 2017, 62. [Google Scholar] [CrossRef]
  46. Zhou, F.; Le-Hussain, F.; Guo, Z.; Yanici, S.; Cinar, Y. Adsorption/desorption characteristics for methane, nitrogen and carbon dioxide of coal samples from Southeast Qinshui Basin, China. Energy Explor. Exploit. 2013, 31, 645–666. [Google Scholar] [CrossRef] [Green Version]
  47. Cheung, O.; Hedin, N. ChemInform Abstract: Zeolites and Related Sorbents with Narrow Pores for CO2 Separation from Flue Gas. ChemInform 2014, 45. [Google Scholar] [CrossRef]
  48. Rios, R.B.; Stragliotto, F.M.; Peixoto, H.R.; Torres, A.E.B.; Bastos-Neto, M.; Azevedo, D.C.S.; Cavalcante, C.L., Jr. Studies on the adsorption behavior of CO2-CH4 mixtures using activated carbon. Braz. J. Chem. Eng. 2013, 30, 939–951. [Google Scholar] [CrossRef]
  49. Khalili, S.; Ghoreyshi, A.A.; Jahanshahi, M.; Khoshandam, B. Predictions of the adsorption equilibrium of CO2/O2 mixture on multi-walled carbon nanotube using Ideal Adsorbed Solution Theory. J. Water Environ. Nanotechnol. 2016, 1, 9–17. [Google Scholar] [CrossRef]
  50. Singh, V.K.; Kumar, E.A. Comparative Studies on CO2 Adsorption Kinetics by Solid Adsorbents. Energy Procedia 2016, 90, 316–325. [Google Scholar] [CrossRef]
  51. Malekbala, M.; Soltani, S.; Abdul Rashid, S.; Luqman Chuah, A.; Rashid, U.; Imededdine, N.; Choong, T.; Teo, S.H. Optimization the Process of Chemically Modified Carbon Nanofiber Coated Monolith via Response Surface Methodology for CO2 Capture. Materials 2020, 13, 1775. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Gomez, L.; Zacharia, R.; Bénard, P.; Chahine, R. Simulation of binary CO2/CH4 mixture breakthrough profiles in MIL-53(Al). J. Nanomater. 2015. [Google Scholar] [CrossRef] [Green Version]
  53. Pino, L.; Italiano, C.; Vita, A.; Fabiano, C.; Recupero, V. Sorbents with high efficiency for CO2 capture based on amines-supported carbon for biogas upgrading. J. Environ. Sci. 2016, 48, 138–150. [Google Scholar] [CrossRef]
  54. Ahn, H.; Moon, J.-H.; Hyun, S.-H.; Lee, C.-H. Diffusion Mechanism of Carbon Dioxide in Zeolite 4A and CaX Pellets. Adsorption 2004, 10, 111–128. [Google Scholar] [CrossRef]
  55. Arefi Pour, A.; Sharifnia, S.; Neishabori Salehi, R.; Ghodrati, M. Adsorption separation of CO2/CH4 on the synthesized NaA zeolite shaped with montmorillonite clay in natural gas purification process. J. Nat. Gas. Sci. Eng. 2016, 36, 630–643. [Google Scholar] [CrossRef]
  56. Tobarameekul, P.; Sangsuradet, S.; Chat, N.; Worathanakul, P. Enhancement of CO2 Adsorption Containing Zinc-ion-exchanged Zeolite NaA Synthesized from Rice Husk Ash. Appl. Sci. Eng. Prog. 2020. [Google Scholar] [CrossRef]
Figure 1. Flowsheet configuration used to run the simulation of adsorption using Aspen Adsorption V11 simulation.
Figure 1. Flowsheet configuration used to run the simulation of adsorption using Aspen Adsorption V11 simulation.
Processes 09 01250 g001
Figure 2. Adsorption isotherm on zeolite 4A at different temperatures and pressures: (a) pure CO2; (b) pure CH4.
Figure 2. Adsorption isotherm on zeolite 4A at different temperatures and pressures: (a) pure CO2; (b) pure CH4.
Processes 09 01250 g002
Figure 3. The crystal structure of zeolite 4A.
Figure 3. The crystal structure of zeolite 4A.
Processes 09 01250 g003
Figure 4. Interaction between CO2 and zeolite 4A.
Figure 4. Interaction between CO2 and zeolite 4A.
Processes 09 01250 g004
Figure 5. Adsorption isotherm of CO2-CH4 mixture on zeolite 4A with the EL isotherm model at different temperatures: (a) 273 K and (b) 303 K.
Figure 5. Adsorption isotherm of CO2-CH4 mixture on zeolite 4A with the EL isotherm model at different temperatures: (a) 273 K and (b) 303 K.
Processes 09 01250 g005
Figure 6. Adsorption isotherm of CO2-CH4 mixture on zeolite 4A at 303 K with different models for EL (dash line) and IAST (solid line).
Figure 6. Adsorption isotherm of CO2-CH4 mixture on zeolite 4A at 303 K with different models for EL (dash line) and IAST (solid line).
Processes 09 01250 g006
Figure 7. The adsorption selectivity of CO2-CH4 mixture on zeolite 4A: (a) at different pressures and compositions and 303 K (solid line is IAST based on a Langmuir isotherm and dash line is EL isotherm); (b) at difference temperature, 1 bar and 50:50 of CO2:CH4 ratio.
Figure 7. The adsorption selectivity of CO2-CH4 mixture on zeolite 4A: (a) at different pressures and compositions and 303 K (solid line is IAST based on a Langmuir isotherm and dash line is EL isotherm); (b) at difference temperature, 1 bar and 50:50 of CO2:CH4 ratio.
Processes 09 01250 g007
Figure 8. Cumulative adsorbed amount versus time at different temperatures (a) CO2 (b) CH4 with different pressures for 1 bar (round dot), 5 bar (dash line), and 10 bar (solid line).
Figure 8. Cumulative adsorbed amount versus time at different temperatures (a) CO2 (b) CH4 with different pressures for 1 bar (round dot), 5 bar (dash line), and 10 bar (solid line).
Processes 09 01250 g008
Figure 9. Breakthrough curves on the effect of physical properties of zeolite 4A at 303 K and 10 bar (a) CO2 and (b) CH4.
Figure 9. Breakthrough curves on the effect of physical properties of zeolite 4A at 303 K and 10 bar (a) CO2 and (b) CH4.
Processes 09 01250 g009
Figure 10. Breakthrough curves on the effect of binary gas mixture on zeolite 4A at 303 K and 10 bar (a) CO2 and (b) CH4.
Figure 10. Breakthrough curves on the effect of binary gas mixture on zeolite 4A at 303 K and 10 bar (a) CO2 and (b) CH4.
Processes 09 01250 g010
Figure 11. Comparison of the prediction of the Thomas and Yoon-Nelson model with the experimental breakthrough curve.
Figure 11. Comparison of the prediction of the Thomas and Yoon-Nelson model with the experimental breakthrough curve.
Processes 09 01250 g011
Table 1. Physical properties of the zeolite 4A adsorbent and column used in simulation of adsorption.
Table 1. Physical properties of the zeolite 4A adsorbent and column used in simulation of adsorption.
ParametersValue
Packing length (mm)97.5
Internal bed diameter (mm)9.1
Particle size (cm)0.2 (average diameter)
Pellet density (kg/m3)1109
Solid density of adsorbent (kg/m3)2429
Total intrusion volume (cm3/g)0.3147
BET surface area (m2/g)501.4
Table 2. Properties of carbon dioxide and methane.
Table 2. Properties of carbon dioxide and methane.
PropertiesCO2CH4
Kinetic diameter (Å)3.33.8
Critical temperature (K)304.12190.56
Quadrupole moment (×1026 esu cm2)4.300
Polarizability (×10−25 cm3)29.1125.93
Table 3. Parameters of the Langmuir isotherm model for CO2 and CH4 adsorption on zeolite 4A.
Table 3. Parameters of the Langmuir isotherm model for CO2 and CH4 adsorption on zeolite 4A.
EquationParametersCO2CH4
q = q m b P 1 + b P qm,0 (mmol/g)3.78813.4698
b = b 0 e Q R T b0 (bar−1)0.70785.88 × 10−5
q m = q m , 0 exp ( X ( 1 T T 0 ) ) Q/R (K)470.912465.8
X (dimensionless)1.71880.003
RMS (dimensionless)0.09150.0236
Table 4. Comparison of the adsorption capacity for CO2-CH4 binary gas mixture with different adsorbents at 323 K.
Table 4. Comparison of the adsorption capacity for CO2-CH4 binary gas mixture with different adsorbents at 323 K.
AdsorbentP
(Bar)
CO2:CH4 q C O 2
(mmol/g)
q C H 4
mmol/g
Ref.
Zeolite 4A640:604.0760.159This work
1050:504.4000.102
Zeolite 5A5.840:602.8520.150[38]
Zeolite 13X1050:504.2770.252[44]
Table 5. Mass transfer coefficient CO2 and CH4 adsorption on zeolite 4A (0.2 cm) at different conditions of pressure and temperature.
Table 5. Mass transfer coefficient CO2 and CH4 adsorption on zeolite 4A (0.2 cm) at different conditions of pressure and temperature.
T
(K)
P
(Bar)
DP (cm2/s)k (s−1)
CO2CH4CO2CH4
27310.01960.03860.32220.6357
50.00460.00930.07490.1524
100.00230.00470.03820.0781
30310.02320.04500.38200.7400
50.00550.01020.09060.1684
100.00300.00570.04890.0942
34310.02830.05460.46620.8985
50.00690.01270.11370.2089
100.00360.00710.05840.1174
Table 6. Physical properties of zeolite 4A with different types.
Table 6. Physical properties of zeolite 4A with different types.
Type of
Zeolite
Particle
Size
(cm)
Pore
Volume
(cm3/g)
Pore
Diameter
(nm)
BET
Surface
(m2/g)
Zeolite 4A-0.2 cm0.20.3147361501.4
Zeolite 4A (HSD)-0.2 cm0.20.1606320509.8
Zeolite 4A-0.4 cm0.40.3012314510.4
Table 7. Mass transfer coefficients of CO2 and CH4 adsorption on different types of zeolites 4A at 303 K and 10 bar.
Table 7. Mass transfer coefficients of CO2 and CH4 adsorption on different types of zeolites 4A at 303 K and 10 bar.
Type of
Zeolite
DP (cm2/s)k(s−1)
CO2CH4CO2CH4
Zeolite 4A-0.2 cm0.00300.00570.04890.0942
Zeolite 4A (HSD)-0.2 cm0.00280.00560.02690.0546
Zeolite 4A-0.4 cm0.00280.00570.0120.0244
Table 8. Thomas and Yoon-Nelson model parameters.
Table 8. Thomas and Yoon-Nelson model parameters.
ModelParameterValue
ThomaskTh (L/mg min)1.232 × 10−3
q0 (mg/g)654,490.9
R20.948
Yoon-NelsonkYN (min−1)0.567
τ (min)3.347
R20.974
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Parinyakit, S.; Worathanakul, P. Static and Dynamic Simulation of Single and Binary Component Adsorption of CO2 and CH4 on Fixed Bed Using Molecular Sieve of Zeolite 4A. Processes 2021, 9, 1250. https://doi.org/10.3390/pr9071250

AMA Style

Parinyakit S, Worathanakul P. Static and Dynamic Simulation of Single and Binary Component Adsorption of CO2 and CH4 on Fixed Bed Using Molecular Sieve of Zeolite 4A. Processes. 2021; 9(7):1250. https://doi.org/10.3390/pr9071250

Chicago/Turabian Style

Parinyakit, Supatsorn, and Patcharin Worathanakul. 2021. "Static and Dynamic Simulation of Single and Binary Component Adsorption of CO2 and CH4 on Fixed Bed Using Molecular Sieve of Zeolite 4A" Processes 9, no. 7: 1250. https://doi.org/10.3390/pr9071250

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