Next Article in Journal
Application and Comparison of a Modified Protection Scheme Utilizing a Proportional–Integral Controller with a Conventional Design to Enhance Doubly Fed Induction Generator Wind Farm Operations during a Balanced Voltage Dip
Previous Article in Journal
Physical Properties of Hydrocarbon Source Reservoir in the Lucaogou Formation in Junggar Basin (China) and Its Influence on Adsorption Ability and Surface Free Energy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Modeling of CO2 Adsorption on Surface-Functionalized Rubber-Seed Shell Activated Carbon: Isotherm and Kinetic Analysis

1
Department of Chemical Engineering, Universiti Teknologi PETRONAS, Bandar Seri Iskandar 32610, Perak, Malaysia
2
HICoE, Centre of Biofuel and Biochemical Research, Institute of Self-Sustainable Building, Department of Chemical Engineering, Universiti Teknologi PETRONAS, Bandar Seri Iskandar 32610, Perak, Malaysia
3
Department of Fundamental and Applied Sciences, Universiti Teknologi PETRONAS, Bandar Seri Iskandar 32610, Perak, Malaysia
*
Author to whom correspondence should be addressed.
Processes 2023, 11(10), 2833; https://doi.org/10.3390/pr11102833
Submission received: 11 May 2023 / Revised: 23 September 2023 / Accepted: 23 September 2023 / Published: 26 September 2023

Abstract

:
Currently, adsorption is considered a promising technology for CO2 separation with a wide range of adsorbents. A detailed study of equilibrium and kinetics plays a crucial role in the design and operation of industrial adsorption units. In this study, isotherm and kinetics of CO2 adsorption on two RSS-derived AC samples previously prepared in our laboratory were evaluated using equilibrium experiments for pure CO2 at 25 °C and 40 °C and 1 bar. Blank and IL-functionalized AC showed CO2 adsorption capacity of 2.16 mmol/g, 1.96 mmol/g, 1.12 mmol/g and 1.71 mmol/g at 25 °C and 40 °C, respectively. Langmuir, Freundlich, and Temkin equations were used to model adsorption isotherm in low-pressure regions. The obtained results revealed that the Freundlich model provides an accurate fitting to the experimental findings, which indicate that the adsorption process occurs in a heterogeneous phase. Additionally, kinetic analysis was performed by using four empirical models, namely pseudo-first order, pseudo-second order, Elovich, and Avrami’s fractional models. Among the considered kinetic models, the pseudo-second order model fits best for both blank and IL-functionalized AC. Intra-particle and Boyd’s film diffusion models were evaluated for the adsorption mechanism.

Graphical Abstract

1. Introduction

The rapid growth of modern industry has steadily increased the consumption of fossil fuels. The overreliance on fossil fuels has resulted in resource depletion and accelerated destruction of the environment. The huge release of CO2 caused by the burning of fossil fuels has led to serious environmental issues such as melting ice caps, acid rain, rising sea levels, ozone layer depletion, and urban smog. In 2013, the use of fossil fuels caused an enormous global CO2 emission of 36 billion tonnes [1]. The need to control CO2 emissions has urged the development of new processes/materials, some of which are still under research for optimization and enhancement. Absorption is a popular separation technique that is widely used on a commercial scale for post-combustion CO2 capture. The use of amine solvents, including monoethanolamine (MEA), diethanolamine (DEA), and methyldiethanolamine (MDEA), is very common. Despite its popularity, this process has shortcomings (such as equipment corrosion, harmful by-products due to amine leaching, and high regeneration cost) that require an imperative need for a viable alternative [2,3,4,5]. One suitable alternative technique for CO2 separation, especially from point sources, is adsorption [6,7]. Many porous materials have been developed, including zeolites [8,9,10,11], molecular sieves [12], metal-organic frameworks (MOFs) [13,14], carbon nanotubes [15,16,17], and activated carbons [18,19,20,21]. Zeolites and MOFs are widely used for CO2 separation because of their high selectivity for CO2 [22]. Activated carbon (AC) has attracted much research attention due to its advantageous features, including good CO2 adsorption capacity at low concentrations, high selectivity, tunable morphology, and surface functionalization [23,24]. Moreover, a wide range of biomass and agricultural waste are available as precursors. Rubber seed shell is a suitable precursor because of its high carbon content, low cost, and easy availability. Malaysia is the fifth largest producer of natural rubber. As per the Department of Statistics, Malaysia (DOSM), stocks of rubber production were recorded more than 40,000 tons each month between 2018 and 2021 [25]. A statistical overview of rubber production during this period is presented in Figure 1.
Recently, a new branch of hybrid adsorbents called supported ionic liquid phases (SILPs) became popular as they modify the surface and improve the inherent adsorption capacity of carbonaceous adsorbents by surface functionalization with ionic liquids (ILs). Researchers have employed many combinations of ILs and AC for surface functionalization studies [26,27,28]. In our previous study, we successfully transformed rubber seed shells (RSS) into AC using two different methods of chemical activation using K2CO3 [29]. Furthermore, a surface functionalization study was performed on prepared AC using IL via the wet impregnation technique. The prepared adsorbents were characterized and tested for CO2 adsorption [30].
For large-scale applications, a suitable adsorbent should demonstrate high CO2 adsorption capacity and selectivity along with a rapid rate of adsorption. This requires an acceptable knowledge of the adsorption kinetics and mechanism. However, extensive research studies have been conducted to study the CO2 adsorption performance of porous carbon. There is still fragmented literature available on isotherm, and kinetic studies focused on the mechanism of CO2 adsorption, especially for IL-functionalized AC. Previous studies suggest many kinetic models for gas–solid adsorption. The accurate estimation of the kinetic parameters Of CO2 adsorption is difficult due to the many mechanisms involved. Therefore, the simplest approach entails the fitting of experimental adsorption data by several proven kinetic models and choosing the best-fitted model. Generally, a typical kinetic model lumps all the mass transfer resistances into a single overall mass transfer coefficient. This makes the calculations of kinetic parameters simpler but does not provide information regarding the actual rate-controlling step [31].
This study aims to test the suitability of blank and surface-functionalized RSS-derived ACs by performing isotherm and kinetic analyses. The details of the synthesis and characterization of the prepared adsorbents are discussed in our previous study [30]. Batch adsorption experiments were conducted at 25 °C and 40 °C using pure CO2 (99.98%). Experimental equilibrium data were fitted using three isotherm models, namely Freundlich, Langmuir, and Temkin isotherm models. For kinetic analysis, four kinetic models were used to estimate the rate parameters of CO2 adsorption. The details of the isotherm and kinetic equations employed are presented in later sections. Finally, rate-limiting kinetic models, intra-particle diffusion, and Boyd’s film diffusion models were used for estimating the mass transfer mechanism controlling the CO2 adsorption process on these adsorbents.

2. Materials and Methods

2.1. Preparation of Adsorbent

The RSS AC used in this study was synthesized from RSS via the chemical activation method using K2CO3. An impregnation ratio of 1:1 was used, and the activated RSS was carbonized at 800 °C for 120 min under a uniform flow of N2 in a tube furnace. Further details on the synthesis of RSS AC are mentioned elsewhere [30]. The prepared AC sample had a BET surface area and total pore volume of 683.45 m2/g and 0.37 cm3/g with an average pore diameter of 2.16 nm. RSS AC sample is referred to as AC-blank in later sections.
Surface functionalization of AC was performed by [bmpy][Tf2N] IL using the wet impregnation technique. The impregnation solution was made by mixing the desired amount (30 wt% AC) of IL with methanol. AC was mixed with the prepared impregnation solution and stirred at 350 rpm for 30 min. The mixture was left overnight and later dried in an oven at 60 °C for 8 h to remove the solvent. The prepared AC was stored in an airtight bottle and labeled as AC-functionalized.

2.2. CO2 Adsorption Study

The prepared AC samples were tested for CO2 (99.98% pure, provided by Linde Malaysia Sdn. Bhd.) adsorption using high-pressure volumetric analyzer (HPVA II) batch experiments at 25 °C and 40 °C. Each sample was weighed (0.2 g) and filled into the sample holder. Before the adsorption test, the sample was degassed at 150 °C for 4 h to remove impurities. Upon completion of degassing, the sample was cooled to room temperature and transferred to the analysis port. Helium was dosed for free space volume measurement at both ambient and analysis temperatures. Afterward, the sample cell was evacuated to remove helium gas completely, and CO2 was dosed for the adsorption test. After the adsorption test was complete, desorption was performed by reducing the pressure in the sample cell. The temperature of the sample cell was maintained using a Julabo recirculating water bath. Pressure and temperature data were recorded throughout the experiment in a comprehensive data analysis package using Microsoft Excel macro (v.22.0.6) software. A schematic diagram of the HPVA system is shown in Figure 2 [30].

2.3. Adsorption Isotherm

The isotherm study of gas–solid adsorption is significant as it helps to understand the interactions between adsorbate molecules and solid adsorbent. Adsorption isotherm studies are also a useful tool to understand the adsorption performance and to estimate the equilibrium adsorption capacity in some cases. There are numerous adsorption isotherm models available in the literature. A suitable isotherm model is selected, which provides a precise fit of experimental data over the entire pressure range [13]. In this study, Langmuir, Freundlich, and Temkin adsorption isotherms were used. These empirical models describe the process of adsorption of gases on microporous solids, and their applicability for CO2 adsorption has often been reported in the literature [19,32,33].

2.3.1. Langmuir Isotherm Model

The Langmuir isotherm is the simplest theoretical model for explaining the monolayer adsorption on a homogeneous surface. The basic assumptions of the model are as follows [34]:
  • The number of localized active sites is fixed.
  • Each adsorption site can accommodate only one adsorbate molecule, and once adsorbed, adsorbate molecules lose the freedom to migrate to other active sites.
  • All adsorption sites are energetically equal, indicating a homogeneous surface.
  • Adsorbate molecules adsorbed on neighboring active sites do not interact with each other.
  • The adsorption process is in dynamic equilibrium, and the rate of adsorption and desorption are equal at equilibrium.
The mathematical expression of the model is expressed in Equation (1):
q e = q m k L P e 1 + k L P e ,
where, qe and qm are the equilibrium and maximum monolayer adsorption capacity expressed as (mmol/g). Pe (bar) indicates the equilibrium pressure, and kL is the Langmuir isotherm constant, and it represents the affinity between adsorbent and adsorbate molecules. A larger kL value suggests maximum surface coverage and stronger affinity among the adsorbate molecules and the adsorbent. The value of kL is obtained from the linearized form of the model presented in Equation (2).
P e q e = 1 q m P e + 1 k L q m ,

2.3.2. Freundlich Isotherm Model

The Freundlich model is another empirical model which explains a non-ideal and reversible adsorption process. Unlike the Langmuir model, adsorption occurs by multilayer formation. Moreover, the adsorption energy exponentially increased with increasing surface coverage because of reduced active sites [35]. Therefore, as more adsorbate molecules are adsorbed, a higher energy will be required for the other adsorbate molecules to attach to the surface of the adsorbent. The mathematical expression of the model is expressed in Equation (3)
q e = k F ( P e ) 1 n ,
where, kF (mmol·g−1·bar−1/n) is the Freundlich isotherm constant. The value of kF increases when the adsorption capacity is increased. n is the heterogeneity factor, also called the Freundlich coefficient, and represents the deviation in adsorption behavior from linearity. It also indicates the type of adsorption, whether it is physical adsorption (n > 1) or chemical adsorption (n < 1) [35]. 1/n is the Freundlich intensity parameter with a value between 0 and 10 and indicates the magnitude of the driving force for adsorption or surface heterogeneity [36]. Additionally, it also informs about the favorability of the adsorption process. When 1/n < 1, it indicates a spontaneous adsorption process [37]. Generally, the Freundlich isotherm model provides a reasonable fit at a low to mid-pressure range but fails at high pressure [38]. The reason for this behavior is that this model, unlike the Langmuir model, does not approach a limiting (fixed) adsorption capacity as the pressure approaches ∞. This means that, theoretically, an infinite surface coverage is estimated due to the absence of saturation limitation and shows multilayer adsorption on heterogeneous surfaces [38].
Freundlich isotherm constants are estimated using the linearized form expressed in Equation (4):
log q e = log k F + 1 n log P e ,

2.3.3. Temkin Isotherm Model

The Temkin model explains how an adsorption process is affected by adsorbate/adsorbate interactions. Its major assumption is that the heat of adsorption of adsorbate molecules in a single layer decreases linearly with the surface coverage. Temkin model is expressed by Equation (5):
q e = R T b l n P e + R T b l n k T ,
where, kT is the Temkin isotherm constant expressed as (cm3/g·bar), and b is the Temkin constant related to the heat of adsorption expressed as (J/mol).

2.4. Adsorption Kinetics

2.4.1. Rate of Adsorption

The kinetic study is important given that the residence time for completion of the adsorption process, adsorption bed size, and the installation cost are all inherently linked with the rate of adsorption [39]. The ability of an adsorbent to withstand large feed flow rates and its efficiency in dynamic processes depend on the rate of adsorption. A slow adsorption rate makes an adsorbent unfavorable for industrial applications despite a high adsorption capacity and good selectivity. Among the available empirical models, two of the most extensively applied models, Lagergen’s pseudo-first order and pseudo-second order models, are used to study the kinetic behavior of CO2 adsorption on prepared RSS AC. Additionally, the Elovich equation is also fitted to study the chemisorption behavior, and a fractional order kinetic model based on Avrami’s equation of particle nucleation is also used. Avrami’s model has been evaluated for amine-functionalized solid adsorbents [40]; therefore, it is relevant to our research focus.

2.4.2. Pseudo-First Order Model

In the nineteenth century, Lagergren presented an empirical model which states that the rate of adsorption is proportional to the number of available active sites. The mathematical expression of the model is presented in Equation (6):
d q t d t = k 1 ( q e q t ) ,
where, qe and qt (mmol g−1) represent the quantity of adsorbate adsorbed at equilibrium and at any time (min), respectively. k1 (min−1) is the first-order rate constant. Integrating Equation (6) under the boundary conditions t = 0, qt = 0, and t = ∞, qt = qe yields Equation (7) and is used to calculate k1.
q t = q e [ 1 e x p ( k 1 t ) ]
The pseudo-first order model represents the reversible interactions among adsorbent and adsorbate molecules and better explains the CO2 adsorption on physical adsorbents such as activated carbon and zeolites [41,42].

2.4.3. Pseudo-Second Order Model

The pseudo-second order equation was presented by Blanchard et al. to study the adsorption of heavy metals on zeolites [43]. This model assumes that the rate of adsorption is proportional to the square of unoccupied active sites and is mathematically expressed by Equation (8).
d q t d t = k 2 ( q e q t ) 2 ,
where, k2 (mmol g min−1) is the second-order rate constant. Integrating Equation (8) under the boundary conditions t = 0, qt = 0 and t = ∞, qt = qe, yields Equation (9).
q t = q e 2 k 2 t 1 + q e k 2 t ,
Under the assumption of this model, the adsorption process involves strong binding of gas molecules to the surface of the adsorbent. Therefore, it provides a better fit when the adsorption involves chemical interactions or a combination of both physisorption and chemisorption, such as functionalized adsorbents [44]. It also explains solid diffusion rate-controlled processes not accurately explained by the pseudo-first order model. Based on Equation (9), four linearized forms of the model are obtained and used for the calculation of kinetic parameters.
t q t = 1 q e t + 1 k 2 q e 2 ,
1 q t = 1 k 2 q e 2 t + 1 q e ,
q t = 1 k 2 q e q t t + q e ,
q t t = k 2 q e q t + k 2 q e 2 ,

2.4.4. Elovich Kinetic Model

For gas–solid adsorption processes involving chemical interactions, the rate of adsorption decreases over time because of increasing surface coverage. Once equilibrium is reached, desorption starts, and the net rate of adsorption is reduced to zero. One of the suitable kinetic models to describe such a process is the Elovich equation presented in Equation (14):
d q t d t = α e x p ( β q t ) ,
The model parameters are evaluated from the linear form expressed in Equation (15).
q t = 1 β ln α β + 1 β l n ( t ) ,

2.4.5. Avrami’s Kinetic Model

This kinetic model was initially proposed to simulate phase transitions and crystal growth [45]. The mathematical expression of the model is presented in Equation (16):
d q t d t = k A n A t n A 1 q e q t ,
where, kA is Avrami’s kinetic constant and nA is Avrami’s exponent that reflects changes in the adsorption mechanism during the process and the dimensional growth of adsorption sites. nA is equal to 1, 2, or 3 for one-dimensional, two-dimensional, and three-dimensional growth [46]. The integrated form of Equation (16) is used for the calculation of kinetic parameters and is expressed by Equation (17):
q t = q e ( 1 exp k A t n A ) ,

2.5. Validity of Kinetic Model

The reliability of kinetic models is evaluated by the non-linear coefficient of determination (R2) and a normalized standard deviation error function (Δq). The value of R2 shows how well the experimental data points are fitted by the model and is calculated using Equation (18). The error function (Δq) determines the deviation between experimental data and those predicted by a kinetic model equation. The error for each kinetic model is calculated using Equation (19) [40].
R 2 = 1 i = 1 m q i , e x p q i , c a l 2 i = 1 m q i , e x p q e x p 2 × m 1 m p ,
q = [ ( q i , e x p q i , c a l ) / q i , e x p ] 2 N 1 × 100 ,
where, qi,exp, and qi,cal are the experimental and calculated values of adsorbate uptake per mass of the adsorbent expressed as (mmol/g), respectively. qexp is the average of the experimental adsorption capacity. Furthermore, m represents the number of experimental data points until equilibrium is reached (i.e., qt/qe = 1), and p is the number of parameters of the kinetic model [40]. Fitting of all four kinetic models was performed using Excel Solver Tool (Microsoft Excel (Version 2308 Build 16.0.16731.20182)). The value of R2 closer to 1 and the lower value of the error function suggest that the CO2 adsorption process is successfully described by the given kinetic model.

2.6. Rate-Limiting Kinetic Models

The generic kinetic models described in the previous section are convenient for predicting the rate of adsorption by estimating the kinetic parameters. However, the identification of the rate-limiting step is difficult with these models because all the resistances to mass transfer are lumped together in a single step. For porous adsorbents such as AC and silica, the physical meaning of the evaluated rate constant helps to know the mass transfer mechanisms involved. A typical CO2 adsorption process includes the following steps (i) transport of CO2 molecules from bulk gas phase towards the gas–solid interface, (ii) film or boundary layer diffusion where CO2 molecules diffuse through the gas film to reach the surface of the adsorbent, (iii) inter-particle diffusion involving the diffusion of CO2 molecules among the s voids of the agglomerated adsorbent particles, (iv) intra-particle diffusion where CO2 molecules are transported into inner pores of the adsorbent particle, and finally (v) surface adsorption where adsorbed CO2 molecules interact with affinity groups in internal active sites. Generally, the rate of CO2 adsorption is controlled by either film diffusion, intra-particle diffusion, or both in some cases [47,48,49,50,51]. In this study, intra-particle diffusion and Boyd’s film diffusion models are evaluated to investigate the CO2 adsorption mechanism.

2.6.1. Intra-Particle Diffusion Model

This model was originally presented by Weber and Morris to describe the pore diffusion effect and the adsorption behavior for spherical adsorbent particles and is expressed by Equation (20):
q t = k i p d t 1 2 + C ,
where, kipd is the rate constant expressed as (mmol g−1 min−0.5) and C shows the boundary layer thickness. Intra-particle diffusion is the sole rate-controlling mechanism if qt vs. t1/2, called the Weber–Morris plot, is a straight line passing through the origin. A non-linear plot with an intercept indicates that other mechanisms also contribute as a rate-limiting step of adsorption [41].

2.6.2. Boyd’s Film Diffusion Model

This model assumes that gas film adjacent to the surface of the adsorbent offers the main resistance for diffusion of adsorbate molecules, commonly referred to as film or boundary layer resistance. The mathematical expression of the model is represented by Equation (21).
F = 1 6 π 2 n = 1 1 n 2 e x p ( n 2 B t ) ,
For F < 0.85:
B t = π π π 2 F 3 2 ,
For F > 0.85:
B t = 0.4977 ln 1 F ,
where F is the fractional adsorption capacity expressed as:
F = q t q e
Equations (21)–(24) are used to identify the mechanism steps controlling the adsorption process. Whether the adsorption process is controlled by pore diffusion or film diffusion is decided by the graph of Bt vs. time. Pore diffusion is the only rate-controlling step if the Bt vs. time graph is a straight line passing through the origin. In the case of a linear or non-linear plot that does not pass through the origin, film diffusion also significantly affects the rate of adsorption in addition to pore diffusion [50,52,53]. In Boyd’s film diffusion model, the overall resistance to mass transfer is divided into two main components: the film or boundary layer resistance and the internal pore diffusion resistance within the adsorbent material. Bt is called the Biot number and is defined as the ratio of the film resistance to the pore resistance. The magnitude of Bt indicates which resistance dominates in determining the rate-limiting step of the adsorption process.

3. Results and Discussion

3.1. Adsorption Isotherm Analysis

Experimental data of CO2 adsorption obtained from HPVA were fitted to three empirical models to study the adsorption isotherm. The applicability of isotherm models is estimated by the magnitude of the R2 value. The calculated isotherm model parameters are summarized in Table 1. A comparison between experimental isotherm and fitted models is also presented in Figure 3. Langmuir and Freundlich isotherm constants kL and kF represent the affinity among adsorbent and adsorbate molecules and show an increasing trend with higher CO2 adsorption capacity. The values of both constants are reduced at high temperatures for AC-blank, indicating a physisorption behavior. On the other hand, AC-functionalized shows higher values of both kL and kF at high temperatures and does not follow the usual trend. This is due to the higher CO2 adsorption capacity of AC-functionalized at 40 °C than at 25 °C. An increase in the value of isotherm constants indicates that AC-functionalized does not adsorb CO2 entirely by physisorption and has some chemical interactions involved due to IL functionalities.
The Langmuir model parameters for AC-functionalized do not fit the experimental findings accurately as qm representing the maximum monolayer adsorption capacity increases with temperature despite a reduced physisorption behavior at 40 °C. The value of the n parameter of the Freundlich model is greater than 1 for all samples representing physisorption as the main process of adsorption [54]. The values of the Freundlich heterogeneity parameter (1/n) are less than 1 and confirm that the adsorption process is spontaneous. Moreover, these values do not vary much after functionalization, which indicates that the surface of the adsorbent is energetically homogenous after IL impregnation.
The Temkin model represents the adsorbate–adsorbent interactions. The value of the Temkin constant (kT) decreases with temperature for the AC-blank, indicating that the CO2 molecules interact with AC physically (likely weak van der Waals interaction). On the other hand, kT values for AC-functionalized indicate stronger interaction between CO2 and IL functionalities and follow an increasing trend with temperature. Overall, based on the value of R2, the Freundlich model fits better with experimental data of CO2 adsorption for both blank and IL-functionalized AC. Moreover, the Freundlich model parameters agree with the experimental findings, implying that the adsorption process occurs in a heterogeneous phase and is not restricted to monolayers.

3.2. Adsorption Kinetic Analysis

The CO2 adsorption kinetics study was performed by fitting four kinetic model equations. Theoretically, either one of the considered kinetic models could explain the CO2 adsorption process on given adsorbents. Nevertheless, all models might not be a proper fit for both adsorbents depending on the nature of the adsorption process. All four kinetic models are fitted for AC-blank and AC-functionalized. The best-fitted model is used to calculate kinetic constants. The calculated values of kinetic parameters and the corresponding R2 values and errors are recorded in Table 2. It has been established that both pseudo-first order and pseudo-second order models have shortcomings in predicting CO2 adsorption on porous adsorbents such as AC. Various findings in the literature confirm that the pseudo-first order model overestimates the CO2 uptake in the early stage of adsorption, whereas the calculated CO2 uptake is underestimated in the later stages close to equilibrium [31,55]. The fitted kinetic models for blank AC at 25 °C are presented in Figure 4. Both Elovich and the pseudo-second order model provide a good fit for experimental data. The pseudo-second order model provides a better fit to experimental data with an R2 close to 1. This shows that adsorption capacity increases exponentially with the unoccupied surface active sites [42]. The obtained results agree with earlier findings of CO2 adsorption on biomass-derived physical adsorbents [19,56]. Similarly, CO2 adsorption is better fitted by a pseudo-second order model for AC-functionalized, as evident from the fitted kinetic model shown in Figure 5 at both temperatures. Additionally, the R2 values for the pseudo-second order model are much closer to unity compared with other kinetic models. As stated earlier, the pseudo-second order model better suits the adsorption processes involving chemical interactions [57]. This model fits better for AC-functionalized due to chemical interactions among CO2 molecules and IL functionalities.
Temperature has a positive effect on the rate of mass transfer. The values of k1 and k2 increase with temperature due to an increase in the rate of diffusion of CO2 molecules and their mobility. Due to higher mobility, CO2 molecules move faster inside the pores, which, in turn, increases the rate of adsorption. However, in some cases, k1 and k2 can decrease with temperature. It is important to mention that these observations are specific to a given system and may not always follow the same trend for different systems and under different conditions. The behavior of adsorption kinetics can be influenced by various factors, such as the nature of adsorbate–adsorbent interactions, surface properties, and the presence of chemical interactions. As shown in Table 2, calculated values of k1 and k2 increase with temperature for both blank and functionalized AC. High temperature improves the adsorption rate by minimizing the film layer resistance, but its effect on desorption is more pronounced [58]. At high temperatures, CO2 molecules have a greater tendency to escape from the surface of the adsorbent by overcoming the weak van der Waals forces for physisorption. As desorption starts, the net rate of adsorption is reduced at high temperatures. Similar observations were made in another study for CO2 adsorption, which concluded that the time to reach equilibrium at 75 °C is less than that required at low temperatures (25–40 °C) [41]. The shorter equilibrium time is due to faster movement of gas molecules, higher rate of collisions, and lower adsorbed volume. The values of k2 observed for AC-functionalized do not follow this trend because of a higher CO2 adsorption capacity at 40 °C, possibly due to the involvement of strong chemical interactions and high reaction rate. For functionalized AC, higher temperature results in better distribution of IL molecules due to lower viscosity and surface tension. This makes active sites more accessible, and CO2 molecules are also more mobile at high temperatures, which results in an increase in the adsorption capacity. Similar findings were reported by Erto et al. [26] for commercial AC functionalized with two different ILs. Experimental findings showed a higher CO2 adsorption capacity of functionalized AC at 80 °C compared with bare AC despite a poor performance at room temperature due to better IL dispersion at elevated temperatures. Overall, AC samples show promising results in terms of CO2 adsorption when compared with previous studies. A detailed comparison of the adsorption performance of prepared adsorbents reported in the literature is provided in our previous publication [30].
Avrami’s model provides a reasonable fit for only the AC-blank. The unique quality of the fit of this model at both low and high surface coverage is linked with its ability to accommodate complex reaction pathways. Both samples are best represented by a pseudo-second order model with the lowest error function compared with others.

3.3. CO2 Adsorption Mechanism

The study of adsorption kinetics is useful for estimating the rate of adsorption. However, conventional kinetic models do not provide much information regarding the actual mechanism that governs gas–solid adsorption. For this purpose, rate-limiting models have been utilized to investigate the rate-controlling step for CO2 adsorption.
For the intra-particle diffusion model, a plot of qe vs. t0.5 is presented in Figure 6 for both samples at 25 °C and 40 °C. The plot is non-linear for both samples with an intercept. Ideally, if pore diffusion is the sole rate-controlling step, a straight line passing through the origin is obtained. In any other case, other mechanisms control the overall rate of adsorption. The absence of a straight line here indicates that the other mechanisms are affecting the CO2 adsorption for both blank and IL-functionalized AC. Based on the curve-shaped plot, a three-step mechanism can be considered for CO2 adsorption: (i) bulk transport of CO2 molecules across the gas–solid interface to the surface of adsorption mainly controlled by film or boundary layer diffusion, (ii) gradual adsorption stage where intra-particle diffusion takes place and (iii) finally equilibrium stage where intra-particle diffusion is slowed down due to saturation of the active sites.
The contribution of film diffusion or external mass transfer is further confirmed via Boyd’s plot presented in Figure 7. Non-linear plots are obtained at both temperatures for blank and IL-functionalized AC. This confirms that film diffusion plays a significant role in controlling the rate of adsorption in addition to pore diffusion. The plot for AC-functionalized is identical at 25 °C and 40 °C except for a higher slope at 40 °C. This is expected because, at high temperatures, the film resistance is reduced due to faster mobility of CO2 molecules and a less viscous IL layer impregnated on AC. This leads to a higher CO2 adsorption rate and faster kinetics. From the above results, it is evident that, overall, two mechanisms control CO2 adsorption. At a given time, only one mechanism governs the adsorption process. Initially, film diffusion is the main resistance to mass transfer and controls the adsorption rate. Afterward, the fine porosity of the adsorbent makes intra-particle diffusion the rate-limiting mechanism until equilibrium is reached.

4. Conclusions

This study confirms the promising usage of surface-functionalized AC derived from biomass for CO2 adsorption. Overall, the Freundlich isotherm model provided the best fit for both samples under the tested conditions. The isotherm analysis confirmed that the CO2 was adsorbed by multilayer formation in a heterogeneous phase by spontaneous adsorption. The evaluated isotherm constants agree with earlier findings of biomass-derived AC reported in the literature. The best fit of experimental data of CO2 adsorption on blank AC over the tested conditions was obtained by applying a pseudo-second order model. The respective rate constant k2 increased with temperature. Similarly, CO2 adsorption using IL-functionalized AC was also best represented by a pseudo-second order model. The analysis of kinetic parameters indicated that CO2 adsorption mainly occurred through physisorption for blank AC. Whereas, for IL-functionalized AC, CO2 was adsorbed by a combination of both physisorption by the support and chemisorption by IL functionalities. Furthermore, the analysis of diffusion models revealed that mass transfer during CO2 adsorption proceeds with a diffusion-controlled process involving intra-particle diffusion and film diffusion mechanisms. Film diffusion dominates initially during the early stage of adsorption, and pore diffusion controls the overall rate of adsorption in later stages until equilibrium is reached. Lastly, the faster kinetics of CO2 adsorption on these adsorbents was confirmed, thus providing support to the conclusion drawn in our previous study regarding the suitability of surface-functionalized RSS-derived AC for CO2 separation. For future studies, cyclic experiments to estimate selectivity for CO2 adsorption and the lifetime of adsorbent should be performed.

Author Contributions

Conceptualization, A.B.; methodology, A.B., S.S.F. and M.A.; validation, A.B., M.A. and N.A.G.; formal analysis, A.B. and S.S.F.; investigation, S.S.F.; resources, A.B., M.A.; data curation, S.S.F.; writing—original draft preparation, S.S.F.; writing—review and editing, A.B. and M.A. and N.A.G.; visualization, A.B.; supervision, A.B. and M.A.; project administration, A.B.; funding acquisition, A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by YAYASAN Universiti Teknologi PETRONAS (YUTP) Malaysia under the grant number (YUTP-FRG 015LC0-068).

Data Availability Statement

Data sharing not applicable.

Acknowledgments

This study received funding from Yayasan Universiti Teknologi PETRONAS (YUTP-FRG-015LC0-068). The authors gratefully acknowledge Universiti Teknologi PETRONAS for the technical and facilities support. Support from Ministry of Higer Education, Malaysia, (MOHE) through the HICoE award to the Centre of Biofuel and Biochemical Research (CBBR) is also acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hansen, J.; Sato, M.; Ruedy, R.; Lo, K.; Lea, D.W.; Medina-Elizade, M. Global temperature change. Proc. Natl. Acad. Sci. USA 2006, 103, 14288–14293. [Google Scholar] [CrossRef] [PubMed]
  2. Diez, N.; Alvarez, P.; Granda, M.; Blanco, C.; Santamaria, R.; Menendez, R. CO2 adsorption capacity and kinetics in nitrogen-enriched activated carbon fibers prepared by different methods. Chem. Eng. J. 2015, 281, 704–712. [Google Scholar] [CrossRef]
  3. Yu, Y.; Zhang, T.; Wu, X.; Mu, D.; Zhang, Z.; Wang, G.G. Mass and heat transfer characteristic in MEA absorption of CO2 improved by meso-scale method. Int. J. Greenh. Gas Control 2016, 47, 310–321. [Google Scholar] [CrossRef]
  4. Gao, J.; Yin, J.; Zhu, F.; Chen, X.; Tong, M.; Kang, W.; Zhou, Y.; Lu, J. Experimental study of a hybrid solvent MEA-Methanol for post-combustion CO2 absorption in an absorber packed with three different packing: Sulzer BX500, Mellapale Y500, Pall rings 16 × 16. Sep. Purif. Technol. 2016, 163, 23–29. [Google Scholar] [CrossRef]
  5. Zhou, Z.; Mei, L.; Ma, C.; Xu, F.; Xiao, J.; Xia, Q.; Li, Z. A novel bimetallic MIL-101(Cr, Mg) with high CO2 adsorption capacity and CO2/N2 selectivity. Chem. Eng. Sci. 2016, 147, 109–117. [Google Scholar] [CrossRef]
  6. Xiao, G.; Singh, R.; Chaffee, A.; Webley, P. Advanced adsorbents based on MgO and K2CO3 for capture of CO2 at elevated temperatures. Int. J. Greenh. Gas Control 2011, 5, 634–639. [Google Scholar] [CrossRef]
  7. Ma, X.; Li, L.; Wang, S.; Lu, M.; Li, H.; Ma, W.; Keener, T.C. Ammonia-treated porous carbon derived from ZIF-8 for enhanced CO2 adsorption. Appl. Surf. Sci. 2016, 369, 390–397. [Google Scholar] [CrossRef]
  8. Bezerra, D.P.; Oliveira, R.S.; Vieira, R.S.; Cavalcante, C.L., Jr. Adsorption of CO2 on nitrogen-enriched activated carbon and zeolite 13X. Adsorption 2011, 17, 235–246. [Google Scholar] [CrossRef]
  9. Jiang, Q.; Rentschler, J.; Sethia, G.; Weinman, S.; Perrone, R.; Liu, K. Synthesis of T-type zeolite nanoparticles for the separation of CO2/N2 and CO2/CH4 by adsorption process. Chem. Eng. J. 2013, 230, 380–388. [Google Scholar] [CrossRef]
  10. Mohamedali, M.; Ibrahim, H.; Henni, A. Incorporation of acetate-based ionic liquids into a zeolite imidazolate framework (ZIF-8) as efficient sorbents for carbon dioxide capture. Chem. Eng. J. 2018, 334, 817–828. [Google Scholar] [CrossRef]
  11. Abutaleb, A.; Eldoma, M.A.; Imran, M.; Taha, K.K.; Bakather, O.Y.; Zouli, N.; Hegazi, S.E.F.; Hassan, M. Active adsorption performance of planetary ball milled Saudi Arabian bentonite clay for the removal of copper ions from aqueous solution. Europhys. Lett. 2021, 135, 30005. [Google Scholar] [CrossRef]
  12. Bai, B.C.; Cho, S.; Yu, H.-R.; Yi, K.B.; Kim, K.-D.; Lee, Y.-S. Effects of aminated carbon molecular sieves on breakthrough curve behavior in CO2/CH4 separation. J. Ind. Eng. Chem. 2013, 19, 776–783. [Google Scholar] [CrossRef]
  13. Zheng, X.; Huang, Y.; Duan, J.; Wang, C.; Wen, L.; Zhao, J.; Li, D. A microporous Zn(II)-MOF with open metal sites: Structure and selective adsorption properties. Dalton Trans. 2014, 43, 8311–8317. [Google Scholar] [CrossRef]
  14. Barcia, P.S.; Bastin, L.; Hurtado, E.J.; Silva, J.A.C.; Rodrigues, A.E.; Chen, B. Single and multicomponent sorption of CO2, CH4 and N2 in a microporous metal-organic framework. Sep. Sci. Technol. 2008, 43, 3494–3521. [Google Scholar] [CrossRef]
  15. Liu, Q.; Shi, J.; Wang, Q.; Tao, M.; He, Y.; Shi, Y. Carbon dioxide capture with polyethylenimine-functionalized industrial-grade mutiwalled carbon nanotubes. Ind. Eng. Chem. Res. 2014, 53, 17468–17475. [Google Scholar] [CrossRef]
  16. Tsai, C.L.; Chen, C.F. Characterization of bias-controlled carbon nanotubes. Diam. Relat. Mater. 2018, 12, 1615–1620. [Google Scholar] [CrossRef]
  17. Abutaleb, A.; Imran, M.; Zouli, N.; Khan, A.H.; Hussain, S.; Ali, M.A.; Bakather, O.; Gondal, M.A.; Khan, N.A.; Panchal, H.; et al. Fe3O4-multiwalled carbon nanotubes-bentonite as adsorbent for removal of methylene blue from aqueous solutions. Chemosphere 2023, 316, 137824. [Google Scholar] [CrossRef]
  18. Ao, W.; Fu, J.; Mao, X.; Kang, Q.; Ran, C.; Liu, Y.; Zhang, H.; Gao, Z.; Li, J.; Liu, G.; et al. Microwave assisted preparation of activated carbon from biomass: A review. Renew. Sustain. Energy Rev. 2018, 92, 958–979. [Google Scholar] [CrossRef]
  19. Borhan, A.; Yusup, S.; Lim, J.W.; Show, P.L. Characterization and modelling studies of activated carbon produced from rubber-seed shell using KOH for CO2 adsorption. Processes 2019, 7, 855. [Google Scholar] [CrossRef]
  20. Aroua, M.K.; Daud, W.M.; Yin, C.Y.; Adinata, D. Adsorption capacities of carbon dioxide, oxygen, nitrogen and methane on carbon molecular basket derived from polyethyleneimine impregnation on microporous palm shell activated carbon. Sep. Purif. Technol. 2008, 62, 609–613. [Google Scholar] [CrossRef]
  21. Boonpoke, A.; Chiarakorn, S.; Laosiripojana, N.; Towprayoon, S.; Chidthaisong, A. Synthesis of activated carbon and MCM-41 from bagasse and rice husk and their carbon dioxide adsorption. J. Sustain. Energy Environ. 2011, 2, 77–81. [Google Scholar]
  22. Sanz-Perez, E.S.; Arencibia, A.; Calleja, G.; Sanz, R. Tuning the textural properties of HMS mesoporous silica. Functionalization towards CO2 adsorption. Microporous Mesoporous Mater. 2018, 260, 235–244. [Google Scholar] [CrossRef]
  23. Wang, Q.; Luo, J.; Zhong, Z.; Borgna, A. CO2 capture by solid adsorbents and their applications: Current status and new trends. Energy Environ. Sci. 2011, 4, 42–55. [Google Scholar] [CrossRef]
  24. Plaza, M.G.; Pevida, C.; Arias, B.; Fermoso, J.; Rubiera, F.; Pis, J.J. A comparison of two methods for producing CO2 capture adsorbents. Energy Procedia 2009, 1, 1107–1113. [Google Scholar] [CrossRef]
  25. Malaysia, D.o.S. Monthly Rubber Statistics Malaysia, January 2020; Department of Statistics Malaysia: Putrajaya, Malaysia, 2020. [Google Scholar]
  26. Erto, A.; Silvestre-Albero, A.; Silvestre-Albero, J.; Rodriguez-Reinoso, F.; Balsamo, M.; Lancia, A.; Fabio, M. Carbon-supported ionic liquids as innovative adsorbents for CO2 separation from synthetic flue-gas. J. Colloid Interface Sci. 2015, 448, 41–50. [Google Scholar] [CrossRef]
  27. Yusuf, N.Y.; Masdar, M.S.; Isahak, W.N.R.W.; Nordin, D.; Husaini, T.; Majlan, E.H.; Rejab, S.A.M.; Chew, C.L. Ionic-liquid impregnated activated carbon for biohydrogen purification in an adsorption unit. IOP Conf. Ser. Mater. Sci. Eng. 2017, 206, 012071. [Google Scholar] [CrossRef]
  28. Arenillas, A.; Smith, K.M.; Drage, T.C.; Snape, C.E. CO2 capture using some fly ash-derived carbon materials. Fuel 2005, 84, 2204–2210. [Google Scholar] [CrossRef]
  29. Fatima, S.S.; Borhan, A.; Faheem, M. A comparison of two methods for the development of low-cost carbonaceous adsorbent from rubber seed shell (RSS). In Springer Proceedings in Complexity, Proceedings of the 6th international Conference on Fundamental and Applied Sciences: ICFAS 2020, Singapore, 14–16 July 2020; Springer: Singapore, 2021; pp. 107–116. [Google Scholar]
  30. Fatima, S.S.; Borhan, A.; Ayoub, M.; Ghani, N.A. CO2 adsorption performance on surface-functionalized activated carbon impregnated with pyrrolidinium-based ionic liquid. Processes 2022, 10, 2372. [Google Scholar] [CrossRef]
  31. Ammendola, P.; Raganati, F.; Chirone, R. CO2 adsorption on a fine activated carbon in a sound assisted fluidized bed: Thermodynamics and kinetics. Chem. Eng. J. 2017, 322, 302–313. [Google Scholar] [CrossRef]
  32. Zhu, J.; Xin, F.; Huang, J.; Dong, X.; Liu, H. Adsorption and diffusivity of CO2 in phosphonium ionic liquid modified silica. Chem. Eng. J. 2014, 246, 79–87. [Google Scholar] [CrossRef]
  33. Zhu, J.; He, B.; Huang, J.; Li, C.; Ren, T. Effect of immobilization methods and the pore structure on CO2 separation performance in silica-supported ionic liquids. Microporous Mesoporous Mater. 2018, 260, 190–200. [Google Scholar] [CrossRef]
  34. Langmuir, I. The constitution and fundamental properties of solids and liquids. part I. solids. J. Am. Chem. Soc. 1916, 38, 2221–2295. [Google Scholar] [CrossRef]
  35. Freundlich, M.F. Over the adsorption in solution. J. Phys. Chem. 1906, 57, 355–471. [Google Scholar]
  36. Tran, H.N.; You, S.-J.; Chao, H.-P. Thermodynamic parameters of cadmium adsorption onto orange peel calculated from various methods: A comparison study. J. Environ. Chem. Eng. 2016, 4, 2671–2682. [Google Scholar] [CrossRef]
  37. Chowdhury, S.; Mishra, R.; Saha, P.; Kushwaha, P. Adsorption thermodynamics, kinetics and isosteric heat of adsorption of malachite green onto chemically modified rice husk. Desalination 2011, 265, 159–168. [Google Scholar] [CrossRef]
  38. Adelodun, A.A.; Ngila, J.C.; Kim, D.G.; Jo, Y.M. Isotherm, thermodynamic and kinetic studies of selective CO2 adsorption on chemically modified carbon surfaces. Aerosol Air Qual. Res. 2016, 16, 3312–3329. [Google Scholar] [CrossRef]
  39. Monazam, E.R.; Spenik, J.; Shadle, L.J. Fluid bed adsorption of carbon dioxide on immobilized polyethylenimine (PEI): Kinetic analysis and breakthrough behavior. Chem. Eng. J. 2013, 223, 795–805. [Google Scholar] [CrossRef]
  40. Stevens, L.; Williams, K.; Han, W.Y.; Drage, T.; Snape, C.; Wood, J.; Wang, J. Preparation and CO2 adsorption of diamine modified montmorillonite via exfoliation grafting route. Chem. Eng. J. 2013, 215–216, 699–708. [Google Scholar] [CrossRef]
  41. Loganathan, S.; Tikmani, M.; Edubilli, S.; Mishra, A. CO2 adsoprtion kinetics on mesoporous silica under wide range of pressure and temperature. Chem. Eng. J. 2014, 256, 1–8. [Google Scholar] [CrossRef]
  42. Serna-Guerrero, R.; Sayari, A. Modeling adsorption of CO2 on amine-functionalized mesoporous silica. 2: Kinetics and breakthrough curves. Chem. Eng. J. 2010, 161, 182–190. [Google Scholar] [CrossRef]
  43. Blanchard, G.; Maunaye, M.; Martin, G. Removal of heavy metals from waters by means of natural zeolites. Water Res. 1984, 18, 1501–1507. [Google Scholar] [CrossRef]
  44. Singh, V.K.; Kumar, E.A. Comparative studies on CO2 adsorption kinetics by solid adsorbents. Energy Procedia 2016, 90, 316–325. [Google Scholar] [CrossRef]
  45. Lopes, E.C.N.; dos Anjos, F.S.C.; Vieira, E.F.S.; Cestari, A.R. An alternative avrami equation to evaluate kinetic parameters of the interaction of Hg(II) with thin chitosan membranes. J. Colloid Interface Sci. 2003, 263, 542–547. [Google Scholar] [CrossRef] [PubMed]
  46. Benedict, J.B.; Coppens, P. Kinetics of the single-crystal to single-crystal two-photon photodimerization of α-trans-Cinnamic Acid to α-Truxillic Acid. J. Phys. Chem. A 2009, 113, 3116–3120. [Google Scholar] [CrossRef] [PubMed]
  47. Song, G.; Zhu, X.; Chen, R.; Liao, Q.; Ding, Y.-D.; Chen, L. An investigation of CO2 adsorption kinetics on porous magnesium oxide. Chem. Eng. J. 2016, 283, 175–183. [Google Scholar] [CrossRef]
  48. Tsibranska, I.; Hristova, E. Comparison of different kinetic models for adsorption of heavy metals onto activated carbon from apricot stones. Bulg. Chem. Commun. 2011, 43, 370–377. [Google Scholar]
  49. Garces-Polo, S.I.; Villarroel-Rocha, J.; Sapag, K.; Korili, S.A.; Gil, A. A comparative study of CO2 diffusion from adsorption kinetic measurements on microporous materials at low pressures and temperatures. Chem. Eng. J. 2016, 302, 278–286. [Google Scholar] [CrossRef]
  50. Lunhong, A.; Li, M.; Li, L. Adsorption of methylene blue from aqueous solution with activated carbon/cobalt ferrite/alginate composite beads: Kinetics, isotherms, and thermodynamics. J. Chem. Eng. Data 2011, 56, 3475–3483. [Google Scholar] [CrossRef]
  51. Yu, H.; Wang, X.; Xu, C.; Chen, D.-L.; Zhu, W.; Krishna, R. Utilizing transient breakthroughs for evaluating the potential of kureha carbon for CO2 capture. Chem. Eng. J. 2015, 269, 135–147. [Google Scholar] [CrossRef]
  52. Hameed, B.H.; Tan, I.A.W.; Ahmad, A.L. Adsorption isotherm, kinetic modeling and mechanism of 2,4,6-trichlorophenol on coconut husk-based activated carbon. Chem. Eng. J. 2008, 144, 235–244. [Google Scholar] [CrossRef]
  53. Sharma, P.; Das, M.R. Removal of a cationic dye from aqueous solution using grephene oxide nanosheets: Investigation of adsorption parameters. J. Chem. Eng. Data 2013, 58, 151–158. [Google Scholar] [CrossRef]
  54. Perez, N.; Sanchez, M.; Rincon, G.; Delgado, L. Study of the behavior of metal adsorption on acid solutions on lignin using a comparison of different adsorption isotherms. Lat. Am. Appl. Res. 2015, 37, 157–162. [Google Scholar]
  55. Raganati, F.; Alfe, M.; Gargiulo, V.; Chirone, R.; Ammendola, P. Isotherms and thermodynamics of CO2 adsorption on a novel carbon-magnetite composite sorbent. Chem. Eng. Res. Des. 2018, 134, 540–552. [Google Scholar] [CrossRef]
  56. Rashidi, N.A.; Yusup, S.; Borhan, A. Isotherm and thermodynamic analysis of carbon dioxide on activated carbon. Procedia Eng. 2016, 148, 630–637. [Google Scholar] [CrossRef]
  57. Borah, J.M.; Sarma, J.; Mahiuddin, S. Adsorption comparison at the α-alumina/water interface: 3,4-Dihydroxybenzoic acid vs catechol Colloids Surf. A Physicochem. Eng. Asp. 2011, 387, 50–56. [Google Scholar] [CrossRef]
  58. Mengqi, W.; Qingbo, Y.; Wenjun, D.; Limin, H.; Kaijie, L.; Qin, Q.; Sihong, L.; Jinjie, D. 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]
Figure 1. Statistical data for monthly production of natural rubber in Malaysia from 2018–2021.
Figure 1. Statistical data for monthly production of natural rubber in Malaysia from 2018–2021.
Processes 11 02833 g001
Figure 2. (a) Experimental set-up of the high-pressure volumetric analyzer HPVA II (b) Temperature zones in the area above Valve 1, image is reproduced with permission from https://doi.org/10.3390/pr10112372.
Figure 2. (a) Experimental set-up of the high-pressure volumetric analyzer HPVA II (b) Temperature zones in the area above Valve 1, image is reproduced with permission from https://doi.org/10.3390/pr10112372.
Processes 11 02833 g002
Figure 3. A comparison of experimental CO2 adsorption isotherms and those predicted by different isotherm models for (a) AC-blank at 25 °C (b) AC-functionalized at 25 °C (c) AC-blank at 40 °C (d) AC-functionalized at 40 °C.
Figure 3. A comparison of experimental CO2 adsorption isotherms and those predicted by different isotherm models for (a) AC-blank at 25 °C (b) AC-functionalized at 25 °C (c) AC-blank at 40 °C (d) AC-functionalized at 40 °C.
Processes 11 02833 g003
Figure 4. Kinetic models fitted for the experimental CO2 adsorption capacity of AC-blank at 25 °C and 40 °C.
Figure 4. Kinetic models fitted for the experimental CO2 adsorption capacity of AC-blank at 25 °C and 40 °C.
Processes 11 02833 g004
Figure 5. Kinetic models fitted for the experimental CO2 adsorption capacity of AC-functionalized at 25 °C and 40 °C.
Figure 5. Kinetic models fitted for the experimental CO2 adsorption capacity of AC-functionalized at 25 °C and 40 °C.
Processes 11 02833 g005
Figure 6. Weber–Morris plot of CO2 adsorption on (a) AC-blank and (b) AC-functionalized at different adsorption temperatures.
Figure 6. Weber–Morris plot of CO2 adsorption on (a) AC-blank and (b) AC-functionalized at different adsorption temperatures.
Processes 11 02833 g006
Figure 7. Boyd’s film diffusion plots for (a) AC-blank and (b) AC-functionalized at different adsorption temperatures.
Figure 7. Boyd’s film diffusion plots for (a) AC-blank and (b) AC-functionalized at different adsorption temperatures.
Processes 11 02833 g007
Table 1. Isotherm model parameters for CO2 adsorption on blank and IL-functionalized AC.
Table 1. Isotherm model parameters for CO2 adsorption on blank and IL-functionalized AC.
SampleAC-BlankAC-Functionalized
Temperature (°C)25402540
Langmuir isotherm
qmax (mmol/g)2.59072.71961.35352.0117
kL (1/bar)3.29911.72392.96353.1927
R20.98710.97890.98150.9708
Freundlich isotherm
n1.68211.66812.10792.2681
1/n0.59450.44090.47440.4409
kF (mmol/g·bar)2.18711.56191.04541.5619
R20.98890.99640.99680.9998
Temkin isotherm
Bt (J/mol)130.5117103.1559222.0873144.0238
kT (mol/g·bar)59.870720.891236.164144.6218
R20.96080.96260.97380.9536
Table 2. Kinetic model parameters of CO2 adsorption on blank and IL-functionalized AC at 25 °C and 40 °C and 1 bar for pure CO2 feed.
Table 2. Kinetic model parameters of CO2 adsorption on blank and IL-functionalized AC at 25 °C and 40 °C and 1 bar for pure CO2 feed.
UnitsAC-BlankAC-Functionalized
Temperature°C25402540
qe,expcm3/g1.08024.60642.53513.5421
Two parameter models
Pseudo-second
order model
qe,calcm3/g0.84504.65122.31153.5044
k2g·min/cm39.842221.91417.36165.2338
R2-0.99810.99980.99971.000
Δq%23.455.884.985.06
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

Fatima, S.S.; Borhan, A.; Ayoub, M.; Ghani, N.A. Modeling of CO2 Adsorption on Surface-Functionalized Rubber-Seed Shell Activated Carbon: Isotherm and Kinetic Analysis. Processes 2023, 11, 2833. https://doi.org/10.3390/pr11102833

AMA Style

Fatima SS, Borhan A, Ayoub M, Ghani NA. Modeling of CO2 Adsorption on Surface-Functionalized Rubber-Seed Shell Activated Carbon: Isotherm and Kinetic Analysis. Processes. 2023; 11(10):2833. https://doi.org/10.3390/pr11102833

Chicago/Turabian Style

Fatima, Syeda Saba, Azry Borhan, Muhammad Ayoub, and Noraini Abd Ghani. 2023. "Modeling of CO2 Adsorption on Surface-Functionalized Rubber-Seed Shell Activated Carbon: Isotherm and Kinetic Analysis" Processes 11, no. 10: 2833. https://doi.org/10.3390/pr11102833

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