Next Article in Journal
Experimental and Modeling Study of Acetonitrile Separation from Water with Ionic Liquids: VLE Data for Binary and Ternary Systems
Previous Article in Journal
A Data-Driven Approach for Integrated Design and Dynamic Optimization Under Water Demand Uncertainty of Renewable Electrodialysis Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Intensified CO2 Absorption Process Using a Green Solvent: Rate-Based Modelling, Sensitivity Analysis, and Scale-Up

by
Morteza Afkhamipour
1,
Mohammad Shamsi
2,
Seyedsaman Mousavian
3 and
Tohid N. Borhani
4,*
1
National Iranian Gas Company, South Pars Gas Complex, Asaluyeh 75391-311, Iran
2
Process Engineering Department, Faculty of Chemical Engineering, Tarbiat Modares University, Tehran 14115-143, Iran
3
Department of Chemical Engineering, Gac. C., Islamic Azad University, Gachsaran 7581863876, Iran
4
Centre for Engineering Innovation and Research, School of Engineering, Computing and Mathematical Sciences, University of Wolverhampton, Wolverhampton WV1 1LY, UK
*
Author to whom correspondence should be addressed.
Processes 2025, 13(12), 3774; https://doi.org/10.3390/pr13123774
Submission received: 20 October 2025 / Revised: 11 November 2025 / Accepted: 19 November 2025 / Published: 22 November 2025
(This article belongs to the Special Issue CO2 Capture and Low-Carbon Hydrogen Production Processes)

Abstract

Ionic liquids (ILs) are recognized as environmentally friendly solvents due to their high CO2 absorption capacity, ease of recovery, and chemical stability, making them a promising alternative to conventional solvents for CO2 capture. In this study, a rate-based mathematical model was developed for a rotating packed bed (RPB) absorber employing 1-n-butyl-3-methylimidazolium hexafluorophosphate ([bmim][PF6]) as the solvent. The model incorporates mass, energy, and momentum balances, coupled with a thermodynamic model whose parameters were determined using experimental data. The rate-based model was validated against experimental results obtained from the RPB absorber. To enhance predictive accuracy, a sensitivity analysis of various mass transfer correlations was conducted, and simulations were performed based on the outcomes of this analysis. The model provided detailed radial profiles of pressure, gas and liquid flow rates, CO2 concentration, temperature, volumetric mass transfer coefficients, and both gas- and liquid-phase resistances. The results indicated that the CO2 capture efficiency and mass transfer coefficients in both phases increased with rotational speed along the bed’s radial direction. Furthermore, the RPB was designed for a flue gas stream from a fired heater in a petrochemical unit containing 10.74 mol % CO2. The optimal liquid-to-gas ratio at a large scale was found to be 0.3 kg/kg, achieving a CO2 removal efficiency of 98%. Under these conditions, the required motor power at an outer radius of 1.55 m was approximately 24.6 kW. Furthermore, comparison with a conventional packed bed showed that the liquid-phase volumetric mass transfer coefficient in the RPB was significantly higher, confirming its superior mass transfer performance.

1. Introduction

1.1. Background

Over the past century, the atmospheric concentration of CO2 has increased significantly due to the growing energy consumption in the industrial sector. The majority of CO2 emissions originate from the combustion of fossil fuels, which is recognized as one of the primary drivers of climate change and global warming [1,2,3,4]. To achieve the targets set by the Paris Agreement, CO2 emissions must be strictly controlled in the short term, before 2050, to slow down the pace of global warming. In the long term, large-scale CO2 capture and storage (CCS) is required to reduce the overall concentration of CO2 in the atmosphere. Therefore, the development of environmentally safe and stable technologies and approaches for CCS is both urgent and essential [5]. Conventional solutions for reducing CO2 in the atmosphere include CCS, efficient utilization of existing energy resources, and the use of non-fossil fuels such as renewable energy and hydrogen [6]. The carbon capture, utilization, and storage (CCUS) technology has been recognized as an effective and promising approach for large-scale CO2 mitigation, enabling the continued use of fossil fuels, at least in the short to medium term [7,8].
Selecting optimal methods for CO2 capture is crucial in terms of both operational and capital costs. A wide range of approaches, including pressure-swing adsorption, absorption, membrane separation, and cryogenic distillation, have been proposed for CO2 capture [9].
Among the various technologies available, CO2 capture using solvents is a well-established and user-friendly method. Its industrial applications include dehydration, natural gas sweetening, carbon capture, and so on [10].
Current CO2 capture systems at an industrial scale predominantly utilize amine-based absorbents, owing to their significant reactivity with CO2. However, several drawbacks, such as high regeneration heat demand, significant solvent losses, and corrosiveness, are associated with amine-based systems due to their intrinsic properties. These environmental and economic challenges have motivated researchers to explore alternative processes and novel absorbents [11].
The core of absorption processes lies in the use of high-performance solvents; therefore, exploring advanced new solvents to enhance the efficiency of absorption systems is essential [10]. In recent years, advances in green chemical technologies have led to the development of a new class of compounds known as ILs [12].
Due to their unique advantages, such as adequate thermal stability, a wide liquid range, very low vapor pressure, and high and tunable absorption capacity, ILs have attracted increasing attention as absorbents in gas separation processes [13].
In other words, ILs represent a new class of molten salts with melting points near room temperature. Owing to their non-volatility, they are appealing for a wide range of applications, including extraction processes, electrolytic media, heat transfer fluids, and thermal storage media [14,15].
They are entirely composed of anions and cations. The most commonly used ILs are formed by cations such as phosphonium, ammonium, pyridinium, and imidazolium combined with inorganic or organic anions such as nitrate, ethanoate, halide, or tetrafluoroborate [12,14].
Moreover, numerous combinations of anions and cations can produce ILs, and this structural flexibility allows for tuning their physical and chemical properties, making them suitable for designing high-energy-efficiency liquid absorbents for CO2 capture processes [15].
Specifically, ILs are being extensively investigated as absorbents for CO2 capture, aiming to overcome the drawbacks associated with amine-based systems and to advance next-generation absorption technologies [13].
Overall, IL-based absorbents exhibit significantly lower CO2 absorption enthalpies compared to amine solvents, resulting in reduced energy requirements for solvent regeneration. In addition, the low volatility and thermal stability of ILs may decrease the release of volatile components and reduce solvent make-up costs. Finally, ILs are less likely to produce impurities during their reaction with CO2 and are less corrosive [11]. ILs are typically high-viscosity fluids with poor fluidity. A significant limitation for their large-scale application in CO2 absorption processes is the low gas–liquid mass transfer rate in conventional gas–liquid contactors, as well as the high mass transfer resistance due to their elevated viscosity. Therefore, the use of an RPB as an efficient contactor with high mass transfer performance is crucial for such high-viscosity systems. RPBs have been successfully employed in processes such as reactive crystallization, polymer devolatilization, distillation, absorption, and others [3]. Next, the studies conducted on the application of IL-based solvents for the CO2 capture process are reviewed.

1.2. Literature Survey

Researchers have extensively studied ILs containing ammonium and imidazolium for CO2 absorption. Blanchard et al. [16] studied the solubility of CO2 in the IL such as [bmim][PF6] and reported this solvent as a green and environmentally friendly option. Cadena et al. [17] measured the solubility of CO2 in several ILs, including [bmim][PF6], [bmim][BF4], and [emim][Tf2N]. Their results demonstrated that ILs containing the [PF6] anion exhibit higher CO2 absorption due to strong interactions with CO2.
Shiflett and Yokozeki [18] measured the CO2 solubility in [bmim][PF6] and [bmim][BF4] over a temperature range of 283.15–348.15 K and at pressures below 2 MPa. Their results indicated that CO2 solubility decreases with increasing temperature, whereas it increases with increasing pressure.
Ziobrowski et al. [19] studied CO2 absorption in traditional packed beds using ILs and compared them with a 15 wt.% MEA solution. Their results showed that ILs and MEA exhibit comparable CO2 absorption capacities, although the liquid-side mass transfer coefficients for ILs are several times lower, highlighting the need for low-viscosity ILs with improved chemisorption properties.
Seo et al. [11] developed an integrated rate-based and thermodynamic modeling framework for CO2 capture in packed-bed absorbers using ILs. The framework was applied for economic optimization and sensitivity analysis, highlighting the influence of key IL properties—including chemical absorption enthalpy, viscosity, heat capacity, and molar volume—on process performance and providing insights relevant for comparison with advanced amine-based processes.
Afkhamipour et al. [20] applied a rate-based model, incorporating the Deshmukh–Mather thermodynamic method, to simulate CO2 absorption in a packed column using a Diethylenetriamine (DETA) solution and the ionic liquid [bmim][PF6]. Sensitivity analysis was performed to assess the impact of mass transfer correlations on temperature and concentration profiles. The performance of DETA was compared with that of the ionic solvent, revealing that [bmim][PF6] exhibits higher absorption efficiency under the same conditions. The study emphasizes the importance of optimizing absorber column parameters to maximize CO2 removal efficiency while considering the specific advantages and limitations of each.
Hospital-Benito et al. [21] conducted a process modeling study of integrated absorber–stripper systems using six different ILs for CO2 capture. Targeting a 90% CO2 recovery, they evaluated the ILs based on solvent demand, regeneration energy, and column size. Their findings highlighted that the viscosity of ILs and their reaction enthalpy with CO2 are key parameters governing solvent selection and overall process performance.

1.3. Motivation of the Current Study

The lack of modeling studies on CO2 absorption using green solvents highlights the importance of the present research. While amine-based CO2 absorption processes have been extensively investigated, previous findings indicate that the development of IL-based absorption systems could offer significant operational advantages over conventional chemical absorption. However, to date, no comprehensive investigation has been reported on CO2 absorption processes employing green solvents in RPBs, covering the pathway from laboratory-scale validation to industrial-scale applications. This study aims to evaluate the performance of an RPB for CO2 absorption and to assess the feasibility of process intensification for capturing CO2 from flue gas at an industrial scale, thereby highlighting the advantages of HiGee technology over conventional absorption processes. For the first time, this work investigates the variations in key parameters, including KGa, KLa, CO2 concentration in the gas phase, gas and liquid flow rates, gas-phase temperature, CO2 removal efficiency, gas- and liquid-phase resistances, pressure drop, and liquid holdup along the packing. Furthermore, the influence of different mass transfer correlations on the system performance is systematically evaluated. In addition, to evaluate the effectiveness of applying HiGee technology with green solvents, a rigorous iterative approach was implemented to estimate the dimensions of the RPB at an industrial scale. During the scale-up process, the optimal L/G ratio, the rotor’s optimal power consumption, and the variation of CO2 removal efficiency were thoroughly analyzed. Finally, the performance of the RPB absorber using the [bmim][PF6] solvent was compared with that of conventional packed columns. This comparison aims to provide a clearer perspective on the potential adoption and effectiveness of green solvents in RPB-based CO2 absorption processes at both laboratory and industrial scales.

2. Methodology

The rate-based and thermodynamic models are developed in MATLAB software (V14) by considering relevant equations and parameters.

2.1. Thermodynamic Modeling

Determining the mass transfer flux requires knowledge of both the driving force and the mass transfer coefficient. The driving force is defined as the difference between the CO2 concentration in the gas phase and its equilibrium concentration in the liquid phase [22]. In this study, the equilibrium CO2 concentration in the liquid phase was estimated using the UNIQUAC model, following the approach adopted by Yunus et al. [23], for CO2 solubility in a pyridinium-based ionic liquid system. The absorption of CO2 by [bmim][PF6] was analyzed using available solubility data to determine the unknown parameters in the model equations. All relevant equations employed for the thermodynamic modeling in this study are as follows [18]:
f 2 L =   f 2 G     x 2 γ 2 f 2 = y 2 P
f 2 = H e 2 , r γ 2 , r     x 2 = y 2 p H e 2 , r   ( γ 2   / γ 2 , r )
l n   γ i = l n   γ i C + l n   γ i R
l n   γ i C = l n Φ i x i + z 2 q i   l n θ i Φ i + l i Φ i x i   j x j l j   l i = z 2 r i q i r i 1 ;   z = 10   ; r i = k ν k ( i ) R k   ; q i = k ν k ( i ) Q k   Φ i = x i   r i   / j x j   r j   ;   θ i = x i   q i   / j x j   q j  
ln γ i R = q i   1 l n j θ j   τ j i j θ j   τ i j   / k θ k   τ k j τ j i = e x p a j i T   a n d   τ i j = e x p a i j T  
O F = 1 n i = 1 n x 2 e x p x 2 c a l x 2 e x p 2
Equation (1) shows that, at equilibrium, the fugacity of the gas phase is equal to that of the liquid phase when a gaseous solute (component 2) dissolves in a liquid solvent (component 1). Here, x2 and γ2 represent the mole fraction and activity coefficient of the gaseous component 2 in solvent 1, respectively. When the temperature exceeds the critical temperature of the gas, the parameter f 2 corresponds to the fugacity of a hypothetical pure liquid. Since f 2 cannot be determined experimentally, the expression is reformulated to estimate Henry’s law constant for a reference solvent. In Equation (2), γ 2 , r denotes the infinite dilution activity coefficient of the gas in the reference solvent, while H e 2 , r   represents Henry’s law constant of the gas in that solvent.
The UNIQUAC model is employed in this study to determine the activity coefficient. Equation (3) presents the activity coefficient of component i, where l n γ i C represents the combinatorial contribution and l n γ i R accounts for the residual contribution. The combinatorial term, given by Equation (4), is defined using Φ i , the segment fraction, and θ i , the area fraction. The volume and surface area parameters for component i are denoted by r i and q i , respectively. The residual contribution of the UNIQUAC equation is calculated using Equation (5), where τ j i represents the binary interaction parameter. Thus, aji and aij are the two interaction parameters required by the UNIQUAC model for a binary mixture. To minimize errors in estimating these interaction parameters, an objective function (OF), as shown in Equation (6), was employed. Using the UNIQUAC equations, γ 2 and x 2   can be calculated, representing the activity coefficient and solubility of CO2 in the ionic liquid, respectively. The optimized model parameters, including the binary interaction parameters (aij (K)), volume parameters ( r i ), and surface area parameters ( q i ), were obtained in this study and are summarized in Table 1.

2.2. Process Modeling

Traditional absorption columns are associated with operational challenges and high costs due to their large size compared to RPBs. The use of a suitable solvent and an appropriate rotational speed—taking into account the solvent’s viscosity—can enhance mass transfer and consequently reduce operating costs. Modeling and simulating such processes using a detailed rate-based approach significantly improves the accuracy of process design and performance analysis [22,24]. Depending on the system configuration, the gas–liquid flow may occur in co-current, counter-current, or cross-current modes [25]. In this study, a counter-current flow configuration was considered for model development, based on the available experimental data for CO2 absorption in [bmim][PF6] [3].
Figure 1 illustrates the counter-current flow in an RPB for CO2 absorption using [bmim][PF6]. The liquid enters the column and is distributed radially from the inner to the outer radius across the packing surface, while the gas enters from the outer radius and comes into radial contact with the liquid. This liquid distribution is driven by the centrifugal force generated by the rotor and the resulting pressure gradient. Under the modelling assumptions, a differential volume element is defined, and mass and energy balances are derived for this element in the radial direction. Prior to formulating the governing balance equations, the following assumptions were considered in the modeling of the absorber column:
  • The system operates under steady-state conditions.
  • The gas phase follows ideal behavior, whereas the liquid phase demonstrates non-ideal thermodynamic characteristics.
  • The absorber operates under adiabatic conditions.
  • Variations in both liquid and gas flow rates along the column are incorporated into the modeling equations.
The main equations developed for the process modeling of CO2 absorption using [bmim][PF6] in an RPB are presented as follows:
d G d r = 2 π r H j C O 2 a e ϵ G
d L d r = 2 π r H j C O 2 a e ϵ L
d y C O 2 d r = 2 π r H a e ϵ G G y C O 2 + 1 j C O 2
d x C O 2 d r = 2 π r H a e ϵ L L ( x C O 2 1 ) j C O 2
d x I L d r = 2 π r H a e ϵ L L ( x I L 1 ) j C O 2
d T G d r = 2 π r H a e h G ϵ G C p G G T G T L
d T L d r = 2 π r H a e ϵ L ( j C O 2 T L ϵ L L j C O 2 T G C p L L h G T G T L C p L L )
Since the first-order ordinary differential equations for mass, energy, and momentum balances require initial conditions, these were taken as the operating conditions at the inlet of the packed bed for both the gas and liquid phases. The mass transfer flux of CO2 ( j C O 2 ) was calculated using the film theory, as expressed by the following equation:
j C O 2 = K G , C O 2 ( P C O 2 P C O 2 * )
The calculation of the CO2 mass flux requires evaluation of K G , C O 2 , which accounts for resistances in both phases. Since the ionic liquid solvent is highly viscous, the resistance is often dominated by the liquid phase, and the gas-phase resistance may be neglected. However, for greater accuracy, both resistances are considered, as expressed by the following equation:
1 K G , C O 2 = 1 k G , C O 2 + H e C O 2 I L k L , C O 2
Here, the terms H e C O 2 I L / k L , C O 2 and 1 / k G , C O 2 represent the liquid- and gas-phase mass transfer resistances, respectively. The gas-phase mass transfer coefficient, k G , C O 2 , was calculated using the correlation proposed by Onda et al. [26]. For the liquid-side mass transfer coefficient, k L , C O 2 , two correlations were adopted, as reported by Onda et al. [26] and Tung and Mah [27]. A sensitivity analysis was performed using these two correlations for k L , C O 2 and five correlations for the effective interfacial area,   a e , as detailed in Table 2. The purpose of the sensitivity analysis was to evaluate the model’s response to the selected correlations, which have been previously employed by various researchers in absorption column studies [20,22,24,28]. To ensure the accuracy of the results, the model was first validated, and all subsequent predictions and design calculations were carried out using the most appropriate mass transfer correlations. Accurate estimation of the heat and mass transfer coefficients requires detailed knowledge of key physical and hydrodynamic properties, including the density, viscosity, surface tension, and heat capacity of the ionic solvent, as well as gas diffusivity and liquid holdup. The relevant equations and parameters are summarized in Table 3. The regressed correlations for viscosity, density, surface tension, and diffusion coefficient were derived from the experimental data reported by Zhang et al. [3], covering a temperature range of 293–336 K. The other correlations required for modeling the RPB absorber are provided in Table 3.

2.2.1. Numerical Solution

The CO2 absorption system using an ionic liquid in an RPB was modelled by simultaneously solving Equations (7)–(15) under the stated assumptions. Given the counter-current flow configuration, the operating conditions of the gas stream were defined at the outer radius, while those of the IL were specified at the inner radius. In this study, these equations are first-order ordinary differential equations (ODEs) that require initial conditions for their solution. The initial conditions correspond to the operating conditions of the inlet gas and solvent entering the RPB. Since the calculations begin at the outer radius, where only the gas inlet conditions are known and the solvent outlet conditions are unknown, initial guesses for the solvent outflow at this radius were assumed using the shooting method. Once the initial conditions at the outer radius are estimated, the equations are solved using the finite difference method, and the results are computed point by point along the radial direction until the inner radius is reached, where the predicted solvent inlet conditions are compared with the operational conditions obtained from experimental data. The error is then defined and iteratively minimized until convergence is achieved, ensuring the accuracy of the predicted outcomes.

2.2.2. Experimental Data Source

Based on the experimental data reported by Zhang et al. [3], the CO2 absorption using [bmim][PF6] in an RPB equipped with stainless steel wire mesh packing was modelled and simulated. The operating conditions and other parameters of the packed bed, used as inputs for the model, are summarized in Table 4 and Table 5.

2.3. Sensitivity Analysis

Sensitivity analysis was performed using the correlations summarized in Table 2. Among the various correlations available in the literature, those proposed by Tung and Mah [27] and Onda et al. [26] were adopted to estimate the liquid-phase mass transfer coefficient within the framework of the rate-based model. This approach is adopted because the absorption process with ionic solvents is film-controlled. In addition, five correlations were selected for the effective surface area, as detailed in Table 2. For the gas phase, the correlation developed by Onda et al. [26] was applied in all sensitivity analysis cases.

2.4. Iterative RPB Scale-Up Approach

A rigorous design methodology, proposed by Shamsi et al. [41], was employed in this study to estimate the dimensions of the RPB absorber at an industrial scale. The corresponding iterative procedure is illustrated in Figure 2. The inlet flue gas operating conditions for the RPB absorber were based on those of an existing industrial unit in a petrochemical complex [42].
Key operating parameters of the flue gas are summarized in Table 6. Following the proposed methodology, the internal radius of the packing was first estimated using the correlation reported by Agarwal et al. [43]. After determining the specific surface area of the packing and the liquid holdup, the gas superficial velocity at flooding conditions was calculated using the correlation established by Jassim et al. [44]. The axial height (H) of the RPB absorber was then estimated based on the superficial gas velocity, assuming operation at approximately 80% of the flooding velocity. Accordingly, the RPB was designed to operate below flooding conditions to ensure efficient and stable performance. Initially, a value for r o   was assumed. After calculating the mass transfer coefficients and Henry’s constant, the NTU and ATU were subsequently determined using Equations (16) and (17).
N T U = ln Y 2 Y 1
A T U = F G H P K G
Finally, the r o , n e w is determined through an iterative approach, as detailed in Figure 2. As can be seen, the design of an RPB absorber for dimension estimation, specifically regarding axial height, outer radius, and inner radius, depends on the correlations, which significantly affect its separation performance at the industrial scale.

3. Results and Discussion

3.1. Thermodynamic Modeling Outcomes

The validity of the thermodynamic model for CO2 absorption by [bmim][PF6] was assessed using experimental data reported by Shiflett and Yokozeki [18], as shown in Figure 3. As evident from the figure, the results are plotted for different temperatures; each of them is a specific isotherm. For each isotherm, solubility increases with increasing pressure, whereas at a constant pressure, solubility decreases with rising temperature. At 348 K, the decrease in solubility is sufficiently pronounced that the corresponding plot appears nearly linear. In this study, data were evaluated over a pressure range from low pressures up to 2025 kPa. Data in the low-pressure range are particularly critical for the rate-based model, as they enable accurate prediction of CO2 solubility and improve the estimation of Henry’s constant. Given the low partial pressure of CO2 in feed gas, the rate-based model primarily relies on low-pressure solubility data. In contrast, the thermodynamic model was validated over a broader pressure range to provide sufficient information for accurate parameter fitting. By encompassing the entire pressure range, the thermodynamic model can be effectively integrated with the rate-based model, enabling it to utilize low-pressure data while also accommodating higher pressures.

3.2. Process Modeling Outcomes

Figure 4 illustrates the validation of the developed model for CO2 absorption in an RPB by comparing the predicted and experimental kla values as a function of rpm at different liquid and gas flow rates. As shown, the model accurately reproduces the experimental data reported by Zhang et al. [3] under operating conditions of G = 0.95 L.min−1 and L = 43.8 mL.min−1, with an AARD of 7.29%, and at G = 0.85 L.min−1 and L = 43.8 mL.min−1, with an AARD of 6.11%. These outcomes confirm the strong predictive capability of the model. Increasing the rotational speed from 1000 to 3300 r/min consistently enhances kla. However, kLa values are lower at higher gas flow rates, primarily due to the increased liquid-phase resistance during CO2 absorption by the [bmim][PF6] solvent.
Figure 5 illustrates the validation of the developed model for the physical absorption process. As shown, the developed model for CO2 absorption in the RPB accurately predicts the experimental kLa vs. temperature data, achieving an AARD of 7.88% at L = 43.8 mL.min−1 and 8.64% at L = 58.4 mL.min−1. In both cases, increasing the temperature from 293 to 335 K leads to a rise in kLa. This enhancement in the liquid-side mass transfer coefficient is attributed to the reduction in the viscosity of the ionic solvent with rising temperature, which leads to an enhancement in the diffusion coefficient of CO2 in the solvent.
Following validation and confirmation of the model’s accuracy for CO2 absorption, the effects of different parameters along the radial direction of the packing were investigated. As shown in Figure 6, kLa decreases along the radial direction at all rotational speeds. The mass transfer efficiency reaches its maximum near the inner radius of the bed due to the longer contact time between the gas and liquid phases and then declines toward the outer periphery, where the effective interfacial contact is reduced. At any given radial position, increasing the rotational speed (from 1100 to 3300 r/min) leads to an increase in kLa due to the reduction in H e C O 2 I L / k L , C O 2 for the CO2-IL system. Higher rpm increases the centrifugal force, which in turn enlarges the gas–liquid interfacial area and turbulence, thereby enhancing mass transfer. This demonstrates a direct correlation between rotational speed and mass transfer performance. The reduction in mass transfer resistance can be attributed to a decreased diffusion depth, a thinner liquid film, and smaller liquid droplets, all of which contribute to improved kLa. However, an excessive increase in rotational speed has a limited effect on kLa, as the reduction in mass transfer resistance is offset by the decreased gas–liquid contact time, which is unfavorable for the absorption process [3].
Figure 7 illustrates the variation in the kGa along the radial packing length at different rotational speeds. As can be seen, the value of kGa decreases along the radius direction (from ri to ro) at all rotational speeds. The kGa is a function of the L/G ratio, solvent concentration, and rotational speed. Increasing the rotational speed leads to a reduction in mass transfer resistance, enhanced liquid dispersion, smaller droplet sizes, and an increased effective contact area [24]. Consequently, kGa increases with higher motor speed, indicating a direct correlation between rotational speed and kGa in this system.
Figure 8 shows the radial distribution of CO2 mole fraction for different inlet concentrations of flue gas, ranging from 0.05 to 0.2. As the gas passes through the packing, CO2 is efficiently absorbed by the IL, resulting in a significant decrease in its concentration up to a radial distance of approximately 0.02 m. Beyond this radial position, toward the inner radius, the flue gas exits the packed bed with nearly negligible CO2 levels. A steeper radial gradient in CO2 mole fraction is observed at higher inlet concentrations, reflecting the strong absorption capacity of the [bmim][PF6] solvent over a wide range of flue gas compositions. These results demonstrate the effectiveness of the system in handling flue gases with varying CO2 content.
Figure 9 illustrates the radial distribution of the gas-phase flow rate for various inlet flue gas flow rates entering the absorber bed. As noted earlier, the highest CO2 concentration and mass transfer driving force are observed near the external radius of the bed. Due to the counter-current radial flow configuration, where the gas and liquid phases move in opposite directions, CO2 from the incoming flue gas is progressively absorbed as it passes through the packing of the RPB. Consequently, both the CO2 concentration and the gas-phase flow rate decrease along the radial direction, with the treated gas ultimately exiting the system at the inner radius. As shown in Figure 9, a higher inlet flue gas flow rate causes a steeper initial gradient in the gas-phase flow rate along the radial direction. The gradient then levels off near the mid-radius of the packing, with the exact position of this transition determined by the inlet gas flow rate.
Figure 10 illustrates the variation in the liquid-phase flow rate along the packing length at different inlet solvent flow rates. As shown, the liquid flow rate increases from the internal radius toward the external radius due to CO2 absorption by the solvent. In this system, mass transfer resistance is predominantly in the liquid phase, causing kLa to increase from the outer to the inner radius. This rise in kLa reduces the mass transfer driving force, which explains the observed trend in the liquid-phase flow rate.
Figure 11 illustrates the radial temperature distribution of the IL-based absorption system for different inlet gas temperatures. As the flue gas flows through the packing, its temperature increases along the radial direction due to heat exchange with the liquid phase. Once thermal equilibrium is reached, the gas exits the packing from the inner radius with a nearly constant temperature. As the inlet gas temperature rises, the temperature difference between the packing’s inlet and outlet also grows. For instance, when the gas enters the packing at 293 K, the temperature difference is approximately 2.8 K, whereas for an inlet temperature of 323 K, it increases to about 9 K. Although increasing the temperature lowers the solvent viscosity and enhances the gas diffusion coefficient, it simultaneously decreases the gas solubility in the solvent. Generally, in such systems, an increase in temperature leads to a decrease in absorption efficiency. Although increasing the temperature lowers the solvent viscosity and enhances the gas diffusion coefficient, it simultaneously decreases the gas solubility in the solvent. Therefore, it is preferable to operate the system at lower temperatures, especially given the high rotational speed of the system. For the IL-based solvent, operating close to ambient temperature is preferred to maximize CO2 absorption performance.
Figure 12 shows the CO2 capture level along the radial position at diverse motor speeds. The gradual absorption of CO2 throughout the packing bed results in solvent saturation, a reduction in the driving force, and the establishment of equilibrium, ultimately leading to improved overall absorption efficiency along the radial path. Increasing the rotational speed in RPB absorbers promotes the formation of smaller liquid droplets, enlarges the gas–liquid interfacial area, and reduces mass transfer resistance, all of which contribute to enhanced CO2 removal performance [22]. As shown in Figure 12, at a rotational speed of 3500 r/min, the absorption efficiency reaches nearly 100% around the mid-radius of the packing (r ≈ 0.02 m) due to the increased centrifugal force and intensified mass transfer. In this system, the unique physical properties of the IL-based solvent—compared to those of conventional chemical solvents—enable the use of higher rotational speeds. However, in centrifugal systems such as RPBs, excessively high rotational speeds increase energy consumption without a proportional improvement in removal efficiency. Therefore, determining the optimal rotational speed is essential to minimizing energy penalties and ensuring efficient operation.
Figure 13a,b illustrate the influence of rotational speed on the mass transfer resistance in the liquid and gas phases along the radial direction of the packing. The liquid-phase resistance constitutes the dominant contribution, while the gas-phase resistance remains the lowest, confirming that the process is governed by liquid film control. Across all examined rotational speeds (1000–3500 rpm), the liquid-phase mass transfer resistance decreases along the radial direction from the external to the internal radius of the packing. At the outer radius, the liquid film thickens due to CO2 absorption, which reduces the liquid velocity and consequently increases the mass transfer resistance compared to the inner radius.
In contrast, the gas-phase resistance exhibits an opposite trend, decreasing toward the outer radius. The high viscosity of the IL-based solvent suppresses diffusion in the liquid phase, thereby lowering the mass transfer coefficient and increasing the overall resistance. In this system, an inverse relationship exists between the rotational speed and the thickness of the liquid film and droplets, directly influencing the mass transfer coefficient. However, given the quantitatively small thickness of the liquid film, increasing the rotational speed does not significantly alter the liquid-phase resistance. Consequently, as shown in Figure 13, the high viscosity of the fluid results in a narrow range of resistance variation, with values remaining tightly clustered.
At any given rotational speed, the liquid-phase mass transfer resistance increases from the inner to the outer region of the packing. This behavior can be attributed to a reduction in both the liquid film thickness and the liquid-phase mass transfer coefficient. Conversely, the gas-phase resistance decreases due to an increase in its mass transfer coefficient.
Figure 14 illustrates the radial variation in pressure drop across the packing bed under different rotational speeds. Since pressure drop directly influences energy consumption, design optimization, and the selection of suitable packing materials, its analysis is crucial for evaluating the performance of an RPB [22]. The pressure drop in an RPB is influenced by several factors, including rotational speed, packing characteristics (type and size), fluid flow rates, and liquid viscosity. This pressure loss arises from the resistance the gas phase encounters while flowing through the packed bed, which becomes more pronounced at elevated rotational speeds.
The effects of the rotational speeds on holdup at two distinct liquid flow rates are shown in Figure 15. Several parameters, including rotational speed, liquid flow rate, and solvent viscosity, influence the holdup in an RPB. Increasing the rotational speed promotes the formation of dispersed liquid droplets, diminishes the viscous resistance encountered by the liquid, and thins the liquid film on the packing surface [22]. Consequently, the holdup within the bed decreases as the flow regime transitions from the pore regime to the film regime. Moreover, at a constant temperature, increasing the liquid flow rate results in a corresponding rise in holdup. Overall, near the inner radius of the packing, where the local liquid flow rate is higher, the holdup attains its maximum value and then gradually decreases along the radial direction. This behavior underscores the positive influence of liquid flow rate and the inverse effect of rotational speed on holdup.

3.3. Sensitivity Analysis Results

Figure 16 illustrates the impact of rotational speed on the kla, based on experimental data reported in the literature [3]. In this analysis, the liquid-phase mass transfer coefficient was determined using the correlation proposed by Tung and Mah [27]. Five simulation runs were conducted using the effective interfacial area (ae) values listed in Table 2, corresponding to five distinct case studies. The best agreement between model predictions and experimental data was achieved for Case 1, where the correlation proposed by Billet and Schultes [29] was applied to estimate ae. Consequently, Case 1 was selected to validate the model’s performance. The purpose of this section is to identify and evaluate the most relevant empirical correlations reported in the literature for the present system. Furthermore, the rate-based model predictions presented here were obtained using the set of correlations associated with Case 1.
Figure 17 illustrates the radial variation in the kLa for five distinct case studies. In Figure 17a, kla is calculated using the correlation proposed by Onda et al. [26], while Figure 17b shows the results based on the correlation by Tung and Mah [27]. Both subfigures reveal an increasing trend in kLa from radially inward, consistent with the trend previously discussed in Figure 6. Among the five cases, Case 1 exhibits the best performance across the packing, as corroborated by comparison with the experimental data shown in Figure 16.

3.4. Scale-Up Results

Figure 18 depicts the variation in the outer radius as a function of the liquid-to-gas ratio. The scale-up results were derived using the iterative procedure outlined in the flowchart shown in Figure 2. Mass transfer coefficients were calculated based on the correlations from Case 1 of the sensitivity analysis, conducted at a rotational speed of 2000 rpm. One of the main challenges in scaling up an RPB absorber is determining the optimal L/G ratio. Figure 18 illustrates the impact of varying the L/G ratio from 0.1 to 1.0 vs. the external radius of the packing. At the optimal L/G ratio, the estimated industrial-scale dimensions of the RPB absorber, calculated based on the flue gas operating conditions provided in Table 6, are an axial height of 0.48 m, an outer radius of 1.55 m, and an inner radius of 0.53 m.
Figure 19 illustrates the radial profile of CO2 removal efficiency in the large-scale RPB absorber. The lowest CO2 removal efficiency is observed at the outer radius, where the flue gas enters the packing. Efficiency increases progressively along the radial direction, reaching approximately 98% at the inner radius, where the treated gas exits. This high removal efficiency underscores the effective performance of the selected IL solvent under industrial-scale operating conditions.
The relationship between motor power consumption and the external radius of the packing in the RPB absorber is presented in Figure 20. At an outer radius of 1.55 m, the power consumption reaches an optimal value of 24.6 kW. This finding corroborates the optimal outer radius determined via the rigorous design approach, confirming its suitability for industrial-scale operation. It should be noted that due to the higher viscosity of ILs, reducing drag forces requires higher rotational speeds compared to amine systems; consequently, greater motor power is clearly needed.

3.5. Comparison of kLa Value Between RPB and Conventional PB

To evaluate and compare the mass transfer performance of a conventional absorption column and an RPB using an IL solvent, kLa values were calculated for both systems under identical operating conditions, as presented in Table 7. The operating parameters in Table 7 were derived from the values reported in Table 4 and Table 5. Based on the modeling and simulation results for the RPB at two distinct rotational speeds, the data in Table 7 indicate that the mass transfer coefficients are substantially higher than those of the conventional packed bed, confirming the superior performance of the RPB system. These simulation results further demonstrate the model’s capability to accurately predict and compare the performance of both column types across a wide range of operating conditions, offering valuable insights for process optimization.

4. Conclusions and Recommendations for Future Work

A rate-based mathematical model was developed to simulate CO2 absorption in an RPB using [bmim][PF6] as the solvent. The UNIQUAC model parameters were optimized, and the validated results showed excellent agreement with experimental data. The rate-based model accurately predicts experimental outcomes under various operating conditions, achieving an AARD of 7.48%, thereby confirming its robustness. Increasing the rotational speed from 1100 to 3300 r/min enhances mass transfer by reducing liquid-phase resistance, improving liquid distribution, and enlarging the effective interfacial area, which collectively lead to higher values of both kLa and kGa. Additionally, higher gas flow rates cause a more pronounced initial decline in flow velocity, which stabilizes around the mid-radius of the packing, while the liquid flow rate increases toward the outer radius due to CO2 absorption. Although elevated temperatures reduce solvent viscosity and enhance gas diffusivity, they simultaneously decrease CO2 solubility, indicating that lower operating temperatures are more favorable, particularly at high rotational speeds. As the rotational speed increases, liquid-phase mass transfer resistance rises because of liquid film thinning and a reduction in the liquid-side mass transfer coefficient, whereas gas-phase resistance decreases. Liquid holdup increases with higher liquid flow rates but decreases with increasing rotational speed. Sensitivity analysis identified the correlations of Tung and Mah [27] for the liquid-side mass transfer coefficient, Onda et al. [26] for the gas-side, and Billet and Schultes [29] for the interfacial area as the most reliable. These findings highlight the potential of RPB technology for intensified CO2 capture, particularly when utilizing IL solvents. For future research, it is recommended to conduct an economic assessment of green solvents in comparison to conventional chemical solvents. Such an evaluation would provide clearer insights into the industrial-scale adoption of HiGee technology and support more informed decision-making for stakeholders in the field of carbon capture.

Author Contributions

Conceptualization, M.A.; Methodology, M.A., M.S. and T.N.B.; Software, M.A.; Formal analysis, M.S.; Investigation, M.A., M.S. and S.M.; Data curation, M.S.; Writing—original draft, M.A., M.S., S.M. and T.N.B.; Writing—review & editing, T.N.B.; Supervision, T.N.B.; Project administration, T.N.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

All authors certify that they have no affiliations with or involvement in any organisation or entity with any financial interest or non-financial interest in the subject matter or materials discussed in this manuscript.

Nomenclature

AARDabsolute average relative deviation, %
ATUarea of a transfer unit
[bmim][PF6]1-n-butyl-3-methylimidazolium hexafluorophosphate
[bmim][BF4]1-n-butyl-3-methylimidazolium tetrafluoroborate
[emim][Tf2N]1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl)imide
a e effective interfacial area of packing, m2/m3
a t total surface area of packing, m2/m3
CCUScarbon capture, utilization, and storage
CFDcomputational fluid dynamics
C p G molar heat capacity of the gas phase, J/kmol.K
C p L molar heat capacity of the liquid phase, J/kmol.K
DETAdiethylenetriamine
D L IL diffusivity, m2/s
D C O 2 I L CO2 diffusivity in IL, m2/s
d h hydraulic diameter, m = 4ε/ a t
d p effective diameter of packing, m = 6(1-ε)/ a t
f d fraction of cross-sectional area of RPB eye occupied by distributor, s
f 2 property of a hypothetical pure liquid at temperatures above the critical temperature of the gas
f 2 L liquid-phase fugacity
f 2 G gas-phase fugacity
F G gas molar flow rate, kmol/s
gacceleration due to gravity, m/s2
g 0 characteristic centrifugal acceleration, 100 m/s2
Gmolar flow rate of flue gas, kmol/s
Haxial height, m
H e 2 , r Henry’s law constant of the gas in the reference solvent
H e C O 2 I L Henry’s law constant, Pa.m3/mol
ILsionic liquids
j C O 2 CO2 molar flux, kmol/m2.s
K G , C O 2 overall mass transfer coefficient of component i in the gas phase, kmol/m2.s.kPa
k G , C O 2 gas-side mass transfer coefficient, kmol/m2.s.Pa
k L , C O 2 liquid-side mass transfer coefficient, kmol/m2.s.Pa
KGavolumetric gas-side mass transfer coefficient, 1/s
KLavolumetric liquid-side mass transfer coefficient, 1/s
L/Gliquid-to-gas ratio, kg/kg
L m * liquid mass flow rate per unit tangential section area, kg/m2.s
MEAmonoethanolamine
nnumber of data points
Nrotating speed of RPB, rpm
NTUnumber of transfer units
ODEsordinary differential equations
OFobjective function
ptotal pressure of the system, Pa
PBpacked bed
P C O 2 CO2 partial pressure in the bulk gas phase, Pa
P C O 2 * equilibrium CO2 partial pressure
qiarea parameter for component i
Q L liquid volumetric flow rate, m3/s
Q k area parameter for group k
RPBrotating packed bed
rpmrevolutions per minute
r i volume parameter for component i
R k volume parameter for group k
T G gas-side temperature, K
T L liquid-side temperature, K
U G gas superficial velocity, m/s = U G =   Q L / 2 π r H
u L liquid velocity, m/s
v j e t liquid jet velocity, m/s
ν k ( i ) number of groups of type k in component i
x2mole fraction and the activity coefficient of gaseous 2 in solvent 1
y 2 mole fraction of the gas solute in the gas phase
Y 1 CO2 mole fraction in the inlet flue gas stream
Y 2 CO2 mole fraction in the outlet flue gas stream
Greek letters
γ 2 , r activity coefficient of the gas at infinite dilution in the reference solvent
γ 2 activity coefficient of gaseous 2 in solvent 1
γ i C activity coefficient of combinatorial part of the UNIQUAC model
γ i R activity coefficient of residual part of the UNIQUAC model
ϵ G gas holdup
ϵ L liquid holdup
ε packing porosity, m3/m3
θ i average area function
μ L liquid dynamic viscosity, Pa s
v L liquid kinematic viscosity
ρ L liquid density, kg/m3
σ L liquid surface tension, N/m
σ c critical surface tension for packing material
Φiaverage segment fraction
ω rotating speed, r/min
Superscripts and subscripts
calcalculated
expexperimental
Ggas phase
iinner
Lliquid phase
oouter

References

  1. Bazargani, Z.; Sabzi, F. Thermodynamic modeling of CO2 absorption in 1-butyl-3-methylimidazolium-based ionic liquids. J. Mol. Liq. 2016, 223, 235–242. [Google Scholar] [CrossRef]
  2. Zhang, X.; Bao, D.; Huang, Y.; Dong, H.; Zhang, X.; Zhang, S. Gas–liquid mass-transfer properties in CO2 absorption system with ionic liquids. AIChE J. 2014, 60, 2929–2939. [Google Scholar] [CrossRef]
  3. Zhang, L.-L.; Wang, J.-X.; Xiang, Y.; Zeng, X.-F.; Chen, J.-F. Absorption of carbon dioxide with ionic liquid in a rotating packed bed contactor: Mass transfer study. Ind. Eng. Chem. Res. 2011, 50, 6957–6964. [Google Scholar] [CrossRef]
  4. Shamsi, M.; Sheidaei, M.; Karami, B.; Cheraghdar, A.; Bakhsheshi, S.; Afshardoost, A. Development of an off-grid polygeneration system utilizing multi-waste heat recovery from low-grade heat sources for sustainable production of e-methanol, potable water, liquefied CO2, and utilities. Renew. Energy Focus 2025, 55, 100728. [Google Scholar] [CrossRef]
  5. Xie, C.; Dong, Y.; Zhang, L.; Chu, G.-W.; Luo, Y.; Sun, B.-C.; Zeng, X.; Chen, J.-F. Low-concentration CO2 capture from natural gas power plants using a rotating packed bed reactor. Energy Fuels 2018, 33, 1713–1721. [Google Scholar] [CrossRef]
  6. Babamohammadi, S.; Shamiri, A.; Aroua, M.K. A review of CO2 capture by absorption in ionic liquid-based solvents. Rev. Chem. Eng. 2015, 31, 383–412. [Google Scholar] [CrossRef]
  7. Du, M.; Sun, B.; Luo, Y.; Zhang, L.; Chu, G.; Zou, H.; Chen, J. Zinc (II)-1,4,7,10-tetraazacyclododecane complexes catalyzed tertiary amine-based CO2 capture in a rotating packed bed. Sep. Purif. Technol. 2025, 354, 129454. [Google Scholar] [CrossRef]
  8. Ilea, F.-M.; Dragan, S.; Cormos, A.-M. Assessment of mass transfer intensification potential for a CO2 capture process using three-phase fluidized bed. Chem. Eng. Process. Process. Intensif. 2020, 157, 108115. [Google Scholar] [CrossRef]
  9. Zhang, L.; Wang, J.; Liu, Z.; Lu, Y.; Chu, G.; Wang, W.; Chen, J. Efficient capture of carbon dioxide with novel mass-transfer intensification device using ionic liquids. AIChE J. 2013, 59, 2957–2965. [Google Scholar] [CrossRef]
  10. Zhang, X.; Ding, X.; Song, Z.; Zhou, T.; Sundmacher, K. Integrated ionic liquid and rate-based absorption process design for gas separation: Global optimization using hybrid models. AIChE J. 2021, 67, e17340. [Google Scholar] [CrossRef]
  11. Seo, K.; Tsay, C.; Hong, B.; Edgar, T.F.; Stadtherr, M.A.; Baldea, M. Rate-based process optimization and sensitivity analysis for ionic-liquid-based post-combustion carbon capture. ACS Sustain. Chem. Eng. 2020, 8, 10242–10258. [Google Scholar] [CrossRef]
  12. Salgado, J.; Regueira, T.; Lugo, L.; Vijande, J.; Fernández, J.; García, J. Density and viscosity of three (2,2,2-trifluoroethanol + 1-butyl-3-methylimidazolium) ionic liquid binary systems. J. Chem. Thermodyn. 2014, 70, 101–110. [Google Scholar] [CrossRef]
  13. Moya, C.; Palomar, J.; Gonzalez-Miquel, M.; Bedia, J.; Rodriguez, F. Diffusion coefficients of CO2 in ionic liquids estimated by gravimetry. Ind. Eng. Chem. Res. 2014, 53, 13782–13789. [Google Scholar] [CrossRef]
  14. Qiao, Y.; Yan, F.; Xia, S.; Yin, S.; Ma, P. Densities and viscosities of [Bmim][PF6] and binary systems [Bmim][PF6] + ethanol, [Bmim][PF6] + benzene at several temperatures and pressures: Determined by the falling-ball method. J. Chem. Eng. Data 2011, 56, 2379–2385. [Google Scholar] [CrossRef]
  15. Brennecke, J.F.; Gurkan, B.E. Ionic liquids for CO2 capture and emission reduction. J. Phys. Chem. Lett. 2010, 1, 3459–3464. [Google Scholar] [CrossRef]
  16. Blanchard, L.A.; Hancu, D.; Beckman, E.J.; Brennecke, J.F. Green processing using ionic liquids and CO2. Nature 1999, 399, 28–29. [Google Scholar] [CrossRef]
  17. Cadena, C.; Anthony, J.L.; Shah, J.K.; Morrow, T.I.; Brennecke, J.F.; Maginn, E.J. Why is CO2 so soluble in imidazolium-based ionic liquids? J. Am. Chem. Soc. 2004, 126, 5300–5308. [Google Scholar] [CrossRef] [PubMed]
  18. Shiflett, M.B.; Yokozeki, A. Solubilities and diffusivities of carbon dioxide in ionic liquids: [bmim][PF6] and [bmim][BF4]. Ind. Eng. Chem. Res. 2005, 44, 4453–4464. [Google Scholar] [CrossRef]
  19. Ziobrowski, Z.; Krupiczka, R.; Rotkegel, A. Carbon dioxide absorption in a packed column using imidazolium based ionic liquids and MEA solution. Int. J. Greenh. Gas Control. 2016, 47, 8–16. [Google Scholar] [CrossRef]
  20. Afkhamipour, M.; Seifi, E.; Esmaeili, A.; Shamsi, M.; Borhani, T.N. Comparison of CO2 absorption in DETA solution and [bmim]-[PF6] using thermodynamic and process modelling. Fuel 2024, 356, 129607. [Google Scholar] [CrossRef]
  21. Hospital-Benito, D.; Lemus, J.; Moya, C.; Santiago, R.; Palomar, J. Process analysis overview of ionic liquids on CO2 chemical capture. Chem. Eng. J. 2020, 390, 124509. [Google Scholar] [CrossRef]
  22. Shamsi, M.; Darian, J.T.; Afkhamipour, M. Process modeling and sensitivity analysis of intensified post-combustion CO2 capture process by polyamine solution in a rotating packed bed. Fuel 2025, 397, 135294. [Google Scholar] [CrossRef]
  23. Yunus, N.M.; Shaharun, M.S.; Mutalib, M.A.; Murugesan, T. Modeling of CO2 solubility in pyridinium-based ionic liquids using UNIQUAC. In AIP Conference Proceedings, Proceedings of the 4th International Conference on Fundamental and Applied Sciences (ICFAS2016), Kuala Lumpur, Malaysia, 15–17 August 2016; AIP Publishing LLC: Melville, NY, USA, 2016; p. 040014. [Google Scholar] [CrossRef]
  24. Shamsi, M.; Darian, J.T.; Afkhamipour, M. Assessment of mass transfer performance using the two-film theory and surrogate models for intensified CO2 capture process by amine solutions in rotating packed beds. Chem. Eng. Process.-Process Intensif. 2025, 208, 110080. [Google Scholar] [CrossRef]
  25. Kolawole, T.O. Intensified Post-Combustion Carbon Capture Using a Pilot Scale Rotating Packed Bed and Monoethanolamine Solutions. Ph.D. Thesis, Newcastle University, Newcastle upon Tyne, UK, 2019. [Google Scholar]
  26. Onda, K.; Takeuchi, H.; Okumoto, Y. Mass transfer coefficients between gas and liquid phases in packed columns. J. Chem. Eng. Jpn. 1968, 1, 56–62. [Google Scholar] [CrossRef]
  27. Tung, H.-H.; Mah, R.S. Modeling liquid mass transfer in HIGEE separation process. Chem. Eng. Commun. 1985, 39, 147–153. [Google Scholar] [CrossRef]
  28. Afkhamipour, M.; Mofarahi, M. Sensitivity analysis of the rate-based CO2 absorber model using amine solutions (MEA, MDEA and AMP) in packed columns. Int. J. Greenh. Gas Control. 2014, 25, 9–22. [Google Scholar] [CrossRef]
  29. Billet, R.; Schultes, M. Prediction of mass transfer columns with dumped and arranged packings: Updated summary of the calculation method of Billet and Schultes. Chem. Eng. Res. Des. 1999, 77, 498–504. [Google Scholar] [CrossRef]
  30. Rajan, S.; Kumar, M.; Ansari, M.J.; Rao, D.P.; Kaistha, N. Limiting gas liquid flows and mass transfer in a novel rotating packed bed (HiGee). Ind. Eng. Chem. Res. 2011, 50, 986–997. [Google Scholar] [CrossRef]
  31. Puranik, S.; Vogelpohl, A. Effective interfacial area in irrigated packed columns. Chem. Eng. Sci. 1974, 29, 501–507. [Google Scholar] [CrossRef]
  32. Kolev, N. Wirkungsweise von füllkörperschüttungen. Chem. Ing. Tech. 1976, 48, 1105–1112. [Google Scholar] [CrossRef]
  33. Troncoso, J.; Cerdeiriña, C.A.; Sanmamed, Y.A.; Romaní, L.; Rebelo, L.P.N. Thermodynamic properties of imidazolium-based ionic liquids: Densities, heat capacities, and enthalpies of fusion of [bmim][PF6] and [bmim][NTf2]. J. Chem. Eng. Data 2006, 51, 1856–1859. [Google Scholar] [CrossRef]
  34. Tagiuri, A.M. Studies of Solubility of CO2 in Ionic Liquids, Kinetics, and Heat of Reactions of CO2 in Promising Cyclic Amines. Ph.D. Thesis, University of Regina, Regina, SK, Canada, 2019. [Google Scholar]
  35. Kidnay, A.J.; Parrish, W.R.; McCartney, D.G. Fundamentals of Natural Gas Processing; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar] [CrossRef]
  36. Moisseyeva, Y.; Saitova, A.; Strokin, S. Calculating densities and viscosities of natural gas with a high content of C2+ to predict two-phase liquid-gas flow pattern. Petroleum 2023, 9, 579–591. [Google Scholar] [CrossRef]
  37. Fuller, E.N.; Schettler, P.D.; Giddings, J.C. New method for prediction of binary gas-phase diffusion coefficients. Ind. Eng. Chem. 1966, 58, 18–27. [Google Scholar] [CrossRef]
  38. Llerena-Chavez, H.; Larachi, F. Analysis of flow in rotating packed beds via CFD simulations—Dry pressure drop and gas flow maldistribution. Chem. Eng. Sci. 2009, 64, 2113–2126. [Google Scholar] [CrossRef]
  39. Burns, J.; Jamil, J.; Ramshaw, C. Process intensification: Operating characteristics of rotating packed beds—Determination of liquid hold-up for a high-voidage structured packing. Chem. Eng. Sci. 2000, 55, 2401–2415. [Google Scholar] [CrossRef]
  40. Singh, S.P.; Wilson, J.H.; Counce, R.M.; Lucero, A.J.; Reed, G.D.; Ashworth, R.A.; Elliott, M.G. Removal of volatile organic compounds from groundwater using a rotary air stripper. Ind. Eng. Chem. Res. 1992, 31, 574–580. [Google Scholar] [CrossRef]
  41. Shamsi, M.; Darian, J.T.; Afkhamipour, M. A process intensification approach for industrial plant decarbonization: Scale-up, techno-economic, and environmental assessment. Results Eng. 2025, 28, 107153. [Google Scholar] [CrossRef]
  42. Shamsi, M.; Naeiji, E.; Rooeentan, S.; Shahandashty, B.F.; Namegoshayfard, P.; Bonyadi, M. Proposal and investigation of CO2 capture from fired heater flue gases to increase methanol production: A case study. Energy 2023, 274, 127375. [Google Scholar] [CrossRef]
  43. Agarwal, L.; Pavani, V.; Rao, D.P.; Kaistha, N. Process intensification in HiGee absorption and distillation: Design procedure and applications. Ind. Eng. Chem. Res. 2010, 49, 10046–10058. [Google Scholar] [CrossRef]
  44. Jassim, M.S.; Rochelle, G.; Eimer, D.; Ramshaw, C. Carbon dioxide absorption and desorption in aqueous monoethanolamine solutions in a rotating packed bed. Ind. Eng. Chem. Res. 2007, 46, 2823–2833. [Google Scholar] [CrossRef]
Figure 1. Schematic of the RPB and the corresponding domain distribution of the packing bed (reused with permission from [22]).
Figure 1. Schematic of the RPB and the corresponding domain distribution of the packing bed (reused with permission from [22]).
Processes 13 03774 g001
Figure 2. Iterative scale-up approach for CO2 absorption in IL solvent used in RPB. References cited in the figure: Agarwal et al. [43]; Jassim et al. [44].
Figure 2. Iterative scale-up approach for CO2 absorption in IL solvent used in RPB. References cited in the figure: Agarwal et al. [43]; Jassim et al. [44].
Processes 13 03774 g002
Figure 3. Thermodynamic model results vs. experimental data for CO2-[bmim][PF6] system.
Figure 3. Thermodynamic model results vs. experimental data for CO2-[bmim][PF6] system.
Processes 13 03774 g003
Figure 4. Validation of kLa vs. rotational speeds at various liquid and gas flow rates.
Figure 4. Validation of kLa vs. rotational speeds at various liquid and gas flow rates.
Processes 13 03774 g004
Figure 5. Validation of kLa vs. temperature at various liquid and gas flow rates.
Figure 5. Validation of kLa vs. temperature at various liquid and gas flow rates.
Processes 13 03774 g005
Figure 6. Prediction of kLa in the radial direction at various rotational speeds.
Figure 6. Prediction of kLa in the radial direction at various rotational speeds.
Processes 13 03774 g006
Figure 7. Prediction of kGa in the radial direction at various rotational speeds.
Figure 7. Prediction of kGa in the radial direction at various rotational speeds.
Processes 13 03774 g007
Figure 8. Prediction of CO2 mole fraction in the gas phase for diverse inlet CO2 concentrations.
Figure 8. Prediction of CO2 mole fraction in the gas phase for diverse inlet CO2 concentrations.
Processes 13 03774 g008
Figure 9. Prediction of the gas flow rate in the radial direction for different inlet gas flow rates.
Figure 9. Prediction of the gas flow rate in the radial direction for different inlet gas flow rates.
Processes 13 03774 g009
Figure 10. Prediction of liquid flow rate profiles along the radial position for various inlet liquid flow rates.
Figure 10. Prediction of liquid flow rate profiles along the radial position for various inlet liquid flow rates.
Processes 13 03774 g010
Figure 11. Prediction of the temperature in the radial direction for varying inlet gas temperatures.
Figure 11. Prediction of the temperature in the radial direction for varying inlet gas temperatures.
Processes 13 03774 g011
Figure 12. Prediction of CO2 removal efficiency in the radial path for varying rotational speeds.
Figure 12. Prediction of CO2 removal efficiency in the radial path for varying rotational speeds.
Processes 13 03774 g012
Figure 13. Prediction of mass transfer resistance in the radial direction at various rotational speeds: (a) liquid-phase resistance; (b) gas-phase resistance.
Figure 13. Prediction of mass transfer resistance in the radial direction at various rotational speeds: (a) liquid-phase resistance; (b) gas-phase resistance.
Processes 13 03774 g013
Figure 14. Prediction of pressure variation in the radial direction at different rotational speeds.
Figure 14. Prediction of pressure variation in the radial direction at different rotational speeds.
Processes 13 03774 g014
Figure 15. Prediction of the holdup in the radial direction at different rotational speeds.
Figure 15. Prediction of the holdup in the radial direction at different rotational speeds.
Processes 13 03774 g015
Figure 16. Sensitivity analysis for prediction of the kLa at various rotational speeds (cases 1–5, with Tung and Mah correlation [27]).
Figure 16. Sensitivity analysis for prediction of the kLa at various rotational speeds (cases 1–5, with Tung and Mah correlation [27]).
Processes 13 03774 g016
Figure 17. Sensitivity analysis for prediction of the kLa (a): cases 1–5, with Onda et al. [26] correlation, and (b): cases 1–5, with Tung and Mah correlation [27]).
Figure 17. Sensitivity analysis for prediction of the kLa (a): cases 1–5, with Onda et al. [26] correlation, and (b): cases 1–5, with Tung and Mah correlation [27]).
Processes 13 03774 g017
Figure 18. The optimal outer radius determined using a rigorous design approach.
Figure 18. The optimal outer radius determined using a rigorous design approach.
Processes 13 03774 g018
Figure 19. CO2 removal efficiency for an industrial-scale RPB using IL solvent.
Figure 19. CO2 removal efficiency for an industrial-scale RPB using IL solvent.
Processes 13 03774 g019
Figure 20. Motor electrical power consumption for an industrial-scale RPB using IL solvent based on the optimal outer radius.
Figure 20. Motor electrical power consumption for an industrial-scale RPB using IL solvent based on the optimal outer radius.
Processes 13 03774 g020
Table 1. Optimized parameters for the thermodynamic model.
Table 1. Optimized parameters for the thermodynamic model.
r-ILq-ILr-CO2q-CO2a (CO2-IL)a (IL-CO2)
7.1084.53.1391.0205.3142.153
Table 2. Mass transfer correlations set used in the developed model.
Table 2. Mass transfer correlations set used in the developed model.
AuthorsCorrelationsRef.
k G
Onda et al. k G R T a t D G = 5.23 V m * μ G a t 0.7 μ G ρ G D G 1 3 a t d p 2 [26]
k L
1Tung and Mah k L d p D L = 0.918 μ L D L ρ L 1 2 L m * μ L a t 1 3 a t a e 1 3 d p 3 ρ L 2 r ω 2 μ L 2 1 6 [27]
2Onda et al. k L ρ L μ L r ω 2 1 3 = 0.0051 L m * a μ L 2 3 μ L ρ L D L 1 2 a t d p 0.4 [26]
a e
1Billets and Schultes a e a t = 1.5 a t d h 0.5 ρ L u L d h μ L 0.2 ρ L u L 2 d h σ L 0.75 u L 2 r ω 2 d h 0.45 [29]
2Onda et al. a e a t = 1 e x p 1.45 σ c σ L 0.75 L m * a t μ L 0.1 a t L m 2 * r ω 2 ρ L 2 0.05 L m 2 * σ L ρ L a t 0.2 [26]
3Rajan et al. a e a t = 54999 ρ L d p u L μ L 2.2186 u L 2 r ω 2 d p 0.1748 ( ρ L d p u L 2 σ L ) 1.3160 [30]
4Puranik and Vogelpohl a e a t = 1.045 L m * a t μ L 0.041 L m 2 * σ L ρ L a t 0.133 σ C σ L 0.182 [31]
5Kolev a e a t = 0.584 a t u L 2 r ω 2 0.196 ρ L r ω 2 a t 2 γ L 0.49 ( a t d p ) 0.42 [32]
Table 3. Physical and hydrodynamic properties utilized in the developed model.
Table 3. Physical and hydrodynamic properties utilized in the developed model.
PropertyCorrelations/MethodRef.
[bmim][PF6] density, kg/m3 ρ I L = 1.19 T + 1721.1 [3]
[bmim][PF6] surface tension, N/m σ I L = 2.81 × 10 4 T + 0.1305 [3]
[bmim][PF6] viscosity, Pa·s μ I L = 2.55 × 10 6 exp 3091 T [3]
Diffusivity of CO2 in IL solvent, m2/s D C O 2 I L = 4.0 × 1 0 33 T 8.9507 [3]
Heat capacity of IL, J/kmol·K C p I L = 388.01 4.016 × 10 1 T + 1.567 × 10 3 T 2 × 1000 [33]
Henry’s constant of CO2 in [bmim][PF6], Pa H e C O 2 I L = ( 0.0042 T 3 3.7734 T 2 + 1132.5 T 113406 ) × 1 0 6 [34]
Molar heat capacity of gas, J/kmol·KDIPPR method (empirical equation)[35]
Thermal conductivity of the gas, W/m·KDIPPR method (empirical equation)[35]
Gas viscosityHerning–Zipperer method[36]
Gas diffusivityFuller correlation[37]
Pressure drop Δ P R P B = 150 μ G 1 ε 2 d P 2 ε 3 Q G 2 п H l n r o r i
+ ε 0.08 Q G + 2000 R P M 1.22 + ω 1.22 Q G 2
+ 1.75 1 ε ρ L d P ε 3 Q G 2 п H 2 1 r i 1 r o + 1 2 ρ G ω 2 r o 2 r i 2
[38]
Holdup ε = 0.039 g g 0 0.05 U L U 0 0.6 ( v L v 0 ) 0.22 [39]
Rotation energy P m o t o r = 1.2 + 1.1 × 10 3 ρ L r o 2 ω 2 Q L [40]
Table 4. Operating parameters are used for modeling the RPB.
Table 4. Operating parameters are used for modeling the RPB.
Operating VariablesValue
Temperature, K293
Pressure, atm1
Liquid flow rate, mL/min29.2−102.2
Gas flow rate, L/min0.6−1.0
CO2 mole fraction 0.1
Solvent mole fraction1.0 (pure)
Pressure drop in the bed, kPa<2
Table 5. Dimensions of the RPB.
Table 5. Dimensions of the RPB.
ParameterValue
r i , cm1
r o , cm3
H, cm2
Packing type Stainless wire mesh
Table 6. Operating conditions of flue gas from the petrochemical complex [42].
Table 6. Operating conditions of flue gas from the petrochemical complex [42].
Flue Gas Specifications
Temperature, °C40
Pressure, bar abs1
Total flow, kmol/h6447.8
Component (mole %)
CO210.74
N265.73
O21.53
H2O21.18
Ar0.82
Table 7. Comparison of RPB and PB based on the rate-based model for CO2 absorption by IL solvent.
Table 7. Comparison of RPB and PB based on the rate-based model for CO2 absorption by IL solvent.
Operating ParametersRPBPB
Liquid flow rate, mL/min5050
Gas flow rate, L/min0.50.5
CO2 concentration, mol %1010
Packing surface area, m2/m38502000
Packing volume, cm350.2698.17
Pressure, atm11
Temperature, K298298
kLa, 1/s0.024 at rpm = 2000
0.032 at rpm = 3000
0.0019
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

Afkhamipour, M.; Shamsi, M.; Mousavian, S.; Borhani, T.N. Intensified CO2 Absorption Process Using a Green Solvent: Rate-Based Modelling, Sensitivity Analysis, and Scale-Up. Processes 2025, 13, 3774. https://doi.org/10.3390/pr13123774

AMA Style

Afkhamipour M, Shamsi M, Mousavian S, Borhani TN. Intensified CO2 Absorption Process Using a Green Solvent: Rate-Based Modelling, Sensitivity Analysis, and Scale-Up. Processes. 2025; 13(12):3774. https://doi.org/10.3390/pr13123774

Chicago/Turabian Style

Afkhamipour, Morteza, Mohammad Shamsi, Seyedsaman Mousavian, and Tohid N. Borhani. 2025. "Intensified CO2 Absorption Process Using a Green Solvent: Rate-Based Modelling, Sensitivity Analysis, and Scale-Up" Processes 13, no. 12: 3774. https://doi.org/10.3390/pr13123774

APA Style

Afkhamipour, M., Shamsi, M., Mousavian, S., & Borhani, T. N. (2025). Intensified CO2 Absorption Process Using a Green Solvent: Rate-Based Modelling, Sensitivity Analysis, and Scale-Up. Processes, 13(12), 3774. https://doi.org/10.3390/pr13123774

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