Next Article in Journal
Numerical Simulation for Void Coalescence (Water Treeing) in XLPE Insulation of Submarine Composite Power Cables
Next Article in Special Issue
Coproduction of Acrylic Acid with a Biodiesel Plant Using CO2 as Reaction Medium: Process Modeling and Production Cost Estimation
Previous Article in Journal
Optimization Design of Rib Width and Performance Analysis of Solid Oxide Electrolysis Cell
Previous Article in Special Issue
The Zr-Doped CaO CO2 Sorbent Fabricated by Wet High-Energy Milling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mass Transfer Correlation and Optimization of Carbon Dioxide Capture in a Microchannel Contactor: A Case of CO2-Rich Gas

by
Nattee Akkarawatkhoosith
1,
Wannarak Nopcharoenkul
2,
Amaraporn Kaewchada
2 and
Attasak Jaree
3,*
1
Department of Chemical Engineering, Faculty of Engineering, Mahidol University, 25/25 Phuttamonthon 4 Road, Nakhon Pathom 73170, Thailand
2
Department of Agro-Industrial, Food and Environmental Technology, King Mongkut’s University of Technology North Bangkok, Pracharat 1 Road, Wongsawang, Bansue, Bangkok 10800, Thailand
3
Department of Chemical Engineering, Faculty of Engineering, Kasetsart University, Chatuchak, Bangkok 10900, Thailand
*
Author to whom correspondence should be addressed.
Energies 2020, 13(20), 5465; https://doi.org/10.3390/en13205465
Submission received: 12 September 2020 / Revised: 7 October 2020 / Accepted: 17 October 2020 / Published: 19 October 2020
(This article belongs to the Special Issue Carbon Dioxide (CO2) Utilization)

Abstract

:
This work focused on the application of a microchannel contactor for CO2 capture using water as absorbent, especially for the application of CO2-rich gas. The influence of operating conditions (temperature, volumetric flow rate of gas and liquid, and CO2 concentration) on the absorption efficiency and the overall liquid-side volumetric mass transfer coefficient was presented in terms of the main effects and interactions based on the factorial design of experiments. It was found that 70.9% of CO2 capture was achieved under the operating conditions as follows; temperature of 50 °C, CO2 inlet fraction of 53.7%, total gas volumetric flow rate of 150 mL min−1, and adsorbent volumetric flow rate of 1 mL min−1. Outstanding performance of CO2 capture was demonstrated with the overall liquid-side volumetric mass transfer coefficient of 0.26 s−1. Further enhancing the system by using 2.2 M of monoethanolamine in water (1:1 molar ratio of MEA-to-CO2) boosted the absorption efficiency up to 88%.

1. Introduction

High concentration of carbon dioxide in the synthesis gas or biogas product is one of the major issues regarding the environmental pollution (greenhouse effect) and fuel quality, leading to the requirement of post-treatment process(es) for CO2 removal. For instance, high CO2 content in biogas degrades the fuel quality such as the calorific value and anti-knock properties of engine [1]. Over the past decades, many CO2 separation methods have been proposed such as absorption, adsorption, and membrane separation [2]. Among these methods, CO2 absorption is the most widely used due to the relatively lower operating cost and higher efficiency [3]. However, amid the successes, the use of toxic chemical solvents has been involved, causing serious impact on the environment (corrosion) and economic sustainability.
A wide range of toxic solvents have been reported for the use of CO2 removal with the high absorption capacity such as monoethanolamine (MEA) and diethanolamine (DEA) [3,4]. Various achievements of highly effective CO2 capture have been reported. For example, Sahraie et al. [5] investigated the effect of absorption variables on the CO2 absorption efficiency using MEA concentration in the range of 15–30 wt.%. It was found that the maximum of absorption efficiency (95%) was obtained with the requirement of 30 wt.% MEA. Similar results were reported for the use of high MEA concentration [6,7] (3–6 M) to achieve high absorption rate. The high concentration of solvent in the process raised an environmental concern. Moreover, the requirement of solvent purification and regeneration is associated with additional investment and operation costs [8].
A process of CO2 removal can be carried out using various types of unit operation such as bubble column, packed column, and tray column. Such macro-scale equipment is inherently associated with the low mass transfer coefficient [9]. Another type of contactor known as microreactor has been applied as a promising tool for enhancing the mass transfer coefficient in various applications particularly in multi-phase systems. Conceivably, the intensification of CO2 absorption could be achieved using a microtube contactor. Moreover, the issue of high solvent consumption could be alleviated with this technique as demonstrated in our previous work [10], in which microreactor technology was employed to enhance the CO2 capture process, particularly for CO2-rich gas application. Although our process exhibited the superior absorption efficiency (97%, 0.8 mol m−2 s−1) compared to the conventional reactors, the concentration of solvent was still high (4.3 M MEA). This hurdle could be handled by applying the eco-friendly solvent along with the microreactor technology.
Recently, ionic liquids as a green solvent have been reported for enhancing the CO2 absorption efficiency [9]. However, the high solvent cost is still the barrier for industrial applications. Among others, water is considered as the most promising solvent for CO2 absorption. Although this solvent exhibits relatively low CO2 absorption efficiency [11], it provides many important features for industrial applications such as being eco-friendly and readily available as well as allowing for simple unit operation. Hence, the process intensification for this solvent should be further investigated.
In the CO2-H2O separation process, the overall liquid-side volumetric mass transfer coefficient (KL) is the significant factor in determining the absorption efficiency since the major mass transfer resistance lies in the liquid phase [12]. This coefficient is strongly influenced by the operating conditions and flow characteristics in the system. For example, the enhancement of CO2-H2O process using ultrasonic contactor [13] was demonstrated to reduce the mass transfer resistance and to enhance the absorption rate. However, the feed concentration of CO2 was limited at 30 vol% which was not in the range for CO2-rich gas applications. Therefore, the application of microreactor for CO2-H2O absorption process should be extended to enhance CO2 absorption efficiency for CO2-rich gas applications.
The highlight of this research was to propose the simple and effective process for CO2 capture, particularly for CO2-rich gas (40–60 vol.%). In this work, microchannel technology was applied to enhance the CO2 absorption efficiency for the CO2-rich gas, which has not been investigated. Water was used as a green solvent. The effect of operating parameters on the absorption efficiency was investigated and the optimization via experimental design was carried out. The mass transfer coefficient was determined, and the reactor performance was compared with that of other absorbers based on the literature data. This investigation could provide an important information for applying with the other CO2 absorption systems using water. In addition, we demonstrated the improved performance of CO2 absorption by using low concentration of solvent (MEA).

2. Materials and Methods

2.1. Materials

The gases with high purity of CO2 (99.99%) and N2 (99.99%) were supplied by Praxair (Thailand) and Linde (Thailand), respectively. Water was used as a green absorbent. The CO2-N2 mixture (40–60 vol.%) was used to simulate high CO2 content gas at various CO2 concentration levels.
In this work, the rectangular microchannel made of aluminum alloy (AA1050) was used as a CO2 micro-contactor. The microchannel contactor system consisted of five sections including the inlet channels for CO2 gas mixture and absorbent, T-micromixer (500 μm) where the fluid streams were mixed, microchannel for CO2 capture (500 μm in width × 500 μm in depth × 60 mm in length) where the CO2-H2O absorption occurred, gas–liquid separation chamber, and outlet channels for gas and liquid streams. The microchannel contactor system and experimental set-up for CO2 absorption is shown in Figure 1.

2.2. CO2 Capture Process

In this work, nitrogen and CO2 were separately fed to the system via individual mass flow controllers. These two streams combined at a T-mixer to obtain a CO2-N2 gas mixture at a specific ratio or concentration. A stream of water (green solvent) was fed via an HPLC pump to mix with the CO2-N2 gas mixture at another T-mixer located in the front section of CO2 absorber. Note that both streams were separately preheated to the desired temperature prior to entering in the T-micromixer. The absorption process took place in the microtube contactor which was immersed in a hot water bath to control the absorption temperature. The exit-end of the microtube was connected to a separator that has two outlets; one for gas product and another for liquid product. The pressure of the system was controlled by means of a back-pressure regulator connected to the gas product stream. The gas mixture exiting through the back-pressure regulator was analyzed for CO2 content using a CO2 detector (COZIR Wide Range GC-0016). The absorption pressure was kept constant at 1.7 bar. Note that, the effect of pressure was assumed negligible in our system. The valve for liquid product was adjusted to ensure that no liquid was accumulated in the separator where the CO2 mass transfer was neglected due to the very short contact time of CO2 and water. The experimental range of absorption variables are summarized in Table 1.

2.3. Mass Transfer Coefficient Calculation

In this work, the physicochemical absorption took place in the CO2-H2O absorption system. The absorption of N2 in water was neglected in this work since the N2 can barely dissolve in water. The transport phenomena of CO2 can be described based on the two-film theory. First, CO2 from the bulk gas phase is transferred into the gas film. Then CO2 diffuses into the gas–liquid interface and is subsequently transferred across the liquid film out into the bulk of liquid. Hence, the CO2 mass transfer flux ( N CO 2 ) can be calculated via Equations (1) and (2).
N CO 2 = n CO 2 , in n CO 2 , out aV R
N CO 2 =   k G ( P CO 2 bulk   P CO 2 i )   =   k L ( C CO 2 i   C CO 2 bulk )
where n CO 2 is the mole of CO2; a is the gas–liquid interfacial area per unit reactor volume; VR is the reactor volume; and kG and kL are the individual gas-side and liquid-side mass transfer coefficient, respectively. In our absorption system, the gas-side mass transfer coefficient could be neglected due to the low solubility of CO2 in water. Consequently, the overall liquid-side mass transfer coefficient (KL) was approximately equal to the liquid-side mass transfer coefficient (kL). Since the concentration of CO2 at the gas–liquid interface is not readily measurable, the CO2 absorption flux can be expressed in terms of CO2 concentration ( C CO 2 i ) and the overall liquid-side mass transfer coefficient (KL) (see Equation (3)). Henry’s Law was used to describe the linear relationship between the partial pressure of CO2 in the gas phase and C CO 2 i (Equation (4) [14]).
N CO 2 =   K L ( C CO 2 i   C CO 2 bulk )
C CO 2 i =   H CO 2 P CO 2
where H CO 2 is the Henry’s coefficient. The average pressure of CO2 in the bulk gas can be expressed in terms of a logarithmic mean pressure difference based on the inlet and outlet conditions [15] as shown in Equation (5).
P CO 2 bulk   =   ( P CO 2 ,   in bulk P CO 2 ,   out bulk ln ( P CO 2 ,   in bulk P CO 2 ,   out bulk ) )
The Henry constant and diffusion coefficient of CO2 in H2O as a function of temperature were obtained from the literature by Versteeg and Swaaij [16], and Karlsson and Svensson [17] (Equations (6) and (7)). The absorption efficiency is expressed in Equation (8).
H CO 2 =   3 . 54   ×   10 7 exp ( 2044 T )  
D CO 2 H 2 O   =   2 . 35   ×   10 6 exp ( 2119 T )
%   absorption   efficiency   =   ( n CO 2 , in n CO 2 , out n CO 2 , in )   ×   100

3. Results and Discussion

3.1. Main Effect of Absorption Variables

The main effect of variables including the amount of CO2, volumetric flow rate of water, volumetric flow rate of gas mixture, and absorption temperature on the mean absorption efficiency (%) was evaluated based on the 3k factorial design (see Table 1). The main effect results shown in Figure 2a indicate a strong negative effect on the mean absorption efficiency when either the volumetric flow rate of liquid or gas was increased. This was due to the short absorption time (contact time) between gas phase and liquid phase. A similar trend was observed for the overall liquid-side volumetric mass transfer coefficient (KLa) (see Figure 2b) which decreased by increasing the volumetric flow rate. To further describe this behavior, the effect of volumetric flow rates of gas (150 to 200 mL min−1) and liquid (1 to 2 mL min−1) on the absorption efficiency was plotted when other variables were held constant (CO2 inlet fraction of 0.5 and absorption temperature of 40 °C) as shown in Figure 3. The significant decline of absorption efficiency was observed when the volumetric flow rate of liquid was increased. For instance, at the volumetric flow rate of gas of 175 mL min−1, the absorption efficiency was dramatically reduced from 62.1% to 16.2% when the volumetric flow rate of liquid was increased from 1 mL min−1 to 2 mL min−1, resulting in the poor mass transfer rate of CO2 to water (see Figure 3b). Similar behavior was observed when the total flow rate of liquid was increased instead of the volumetric flow rate of liquid (Figure 3).
For high CO2 concentration level, the adsorption capacity of CO2 in water proceeded via the hydration reaction. The series of reactions involved were as follows:
CO2 (g) ↔ CO2 (aq)
CO2 (aq) +H2O ↔ H2CO3
These reactions could enhance both the solubility of CO2 in water and the absorption efficacy. In order to confirm this, a separate CO2 absorption experiment was carried out while the conductivity of the liquid stream exiting the CO2 absorption apparatus was monitored. The operating conditions were at 40% of CO2 inlet concentration, temperature of 30 °C, 150 mL min−1 of gas flow rate, and 1 mL min−1 of liquid flow rate. The conductivity probe was placed inside a round-bottom glass tube with two ports on the side; one for the incoming liquid stream from the apparatus located near the bottom and another one (located at the position such that the probe was sufficiently immersed) for output stream going to the waste collector. The tube was initially filled with deionized water. It was observed that the conductivity of liquid stream significantly rose from 4.4 μS cm−1 (pure DI water) and reached equilibrium at 37.5 μS cm−1. The change in conductivity of the water was due to the existence of bicarbonate ion (carbonic acid). Note that, the CO2 (aq) in water cannot increase the conductivity of the water. This observation was also in line with the work of Bhaduri et al. [18] who investigated the improvement of CO2 capture by enhancing the CO2 hydration reaction. The high rate of CO2 hydration was evidently observed when the pure CO2 was used.
The low impact on the mean absorption efficiency was observed for the cases of the absorption temperature and CO2 fraction in the feed (Figure 2). In case of absorption temperature, the absorption efficiency was approximately unchanged at temperatures below 40 °C and slightly increased at 50 °C. This behavior involved the change of physical properties of CO2 with temperature. For instance, increasing temperature can help increase the CO2 diffusion coefficient [19]; however, the content of CO2 dissolved in water (CO2 solubility) is also decreased [20,21]. This means that the effect of diffusion was dominant over the effect of solubility at high absorption temperature. For the effect of CO2 inlet fraction, the mean absorption efficiency improved when the CO2 inlet fraction increased from 40% to 50% due to the large mass transfer flux ( N CO 2 ) as a result of high CO2 feed concentration as a driving force (see Equation (3)). Further increasing in the CO2 feed concentration (up to 60%) led to a slight decline of CO2 absorption efficiency, as the system was approaching the limit of CO2 solubility in water. These findings were confirmed when plotting the CO2 inlet fraction (40 to 60%) with the absorption temperature (30 to 50 °C) while holding the volumetric flow rate of 175 mL min−1 and 1.5 mL min−1 for gas and liquid streams, respectively (Figure 4). At the same CO2 inlet fraction, a slight effect of reaction temperature on the absorption efficiency was observed. For the case of CO2 fraction, the highest absorption efficiency was obtained for the CO2 inlet fraction of 50%.

3.2. Interaction Effect of Absorption Variables

The interaction effect of absorption variable pairs can be elaborated through both P-value and contour plot. When the P-value is larger than 0.05 or the slope of contour line is straight and parallel, the interaction effect is not statistically significant. The ANOVA results as shown in Table 2 indicate that the interaction effect between the total gas volumetric flow rate (G) and liquid volumetric flow rate (L) was statistically significant. This behavior was related to the contact time between gas phase and liquid phase which was inversely proportional to the volumetric flow rate. The contact time was relatively shorter at high volumetric flow rate of gas and high volumetric flow rate of liquid, and vice versa. The contour plot of this variable pair is shown in Figure 5a. At high level of the total gas volumetric flow rate, the absorption efficiency increased slightly with decreasing volumetric flow rate of liquid. A similar behavior was observed with higher sensitivity at low level of the total gas volumetric flow rate, signifying the interaction effect of these variables.
On the contrary, the interaction effect of other variable pairs was not statistically significant due to the large p-value (>0.05). This was confirmed by the contour plots in Figure 5b,c.

3.3. Flow Pattern of CO2 Capture Process

According to the literature, the superficial velocity of liquid and gas plays an important role on the gas-liquid flow pattern in a microreactor system. Hassan et al. [22] described the flow pattern of gas-liquid inside the microreactor system by analyzing the literature data for gas-liquid systems and developed a universal map of gas-liquid flow regime (using the superficial velocity of gas and liquid as coordinates) to predict the flow pattern of the system. Note that, similar flow pattern maps were also found in the literature for different hydraulic diameters between 0.1 mm to 1 mm, also with different materials [23,24]. Under the condition of high superficial velocity of gas and low superficial velocity of liquid, the slug-annular flow pattern was observed. This observation was confirmed by the experiment of Yue et al. [25], who studied the flow pattern of CO2-water inside a microchannel contractor with the hydraulic diameter of 0.667 mm and observed the slug-annular flow pattern when using high superficial velocity of gas and low superficial velocity of liquid. Our microreactor reactor had a hydraulic diameter of 0.5 mm and was operated with high superficial velocity of gas (10–13 m s−1) and high superficial velocity of liquid (0.07–0.13 m s−1). Under these conditions, our gas–liquid flow pattern would fall into the churn regime or slug-annular flow pattern, contributing to the large interfacial area between the liquid and gas phases [25].

3.4. Correlation Model

The correlation between the response (%absorption efficiency) and the operating variables of our absorption system (CO2 fraction (F), total gas volumetric flow rate (G), liquid volumetric flow rate (L), and temperature (T)) was be evaluated by regression analysis. The result is shown in Equation (11). The optimization model indicated that the volumetric flow rate of gas and the volumetric flow rate of liquid strongly influenced the absorption efficiency. This was in agreement with the optimization of other absorption systems reported in the literature [7,26]. The precision of this model was verified based on the coefficient of determination (R2) by the parity plot between the experimental and predicted values in terms of the response parameter as shown in Figure 6a. The R2 value of 0.9 suggested that the accuracy of model prediction was reasonable.
The overall liquid-side volumetric mass transfer coefficient of the system (KLa) can be expressed in terms of dimensionless parameters consisting of a group of physical properties and geometry of the system [27,28], i.e., Reynold number (Re), Sherwood number (Sh), Schmidt number (Sc), and capillary number (Ca). The correlation is presented in Equation (12). Note that, the KLa is embedded in the Sherwood number as modified Sherwood number (Sh*) (see Equation (13)). The correlation of our model was in line with those reported in the literature [29,30].
The accuracy of this model was validated as presented in Figure 6b. The mean square errors (MSE) of the efficiency model and overall liquid-side volumetric mass transfer coefficient of the system were 4.1% and 4.9%, respectively. The deviation of the model especially for the cases of low absorption efficiency and small KLa was possibly caused by the unstable pressure inside the system when the high volumetric flow rate of gas and liquid conditions were applied. The change of pressure can affect the hydrodynamic behavior, interfacial area, and mass transfer coefficient [31]. However, the effect of pressure on the absorption efficiency and KLa was little when compared to the other variables, i.e., temperature, and volumetric flow rate of gas and liquid [10].
%absorption efficiency = 595.16 − 0.45F − 3.98G − 198.37L − 11.68T + 0.019F × G − 2.41F × L + 0.07F × T + 1.50G × L + 0.09G × T + 3.36L × T+0.004F × G × L − 0.0008F × G × T + 0.03F × L × T − 0.03G × L × T
Sh L * = 3 . 67 Re G 0.26 Re L 0.46 Sc L 0.44 Ca L 1.31
Sh L * =   K L aD d

3.5. Optimization

The objective function for optimization was set to maximize the absorption efficiency. The optimization was performed based on the response surface methodology and the optimal conditions with the maximum %absorption of 70.9% were found at the temperature of 50 °C, CO2 fraction of 53.7%, the total volumetric flow rate of gas of 150 mL min−1, and the volumetric flow rate of water of 1 mL min−1. Under these conditions, KLa of 0.26 s−1 was calculated. This will be compared with other systems of CO2 absorption using water in Section 3.6. It is noted that this technique may be particularly useful for reducing the costs associated with chemical absorption for CO2 removal [32,33]. Although the quality of water in practical applications can also varied drastically such as tap water, underground water, and wastewater, this result can be used as a guideline for choosing the operating conditions.

3.6. Comparison of the Liquid-Side Volumetric Mass Transfer Coefficient for Different Systems

The performance of our system for CO2 capture application was compared with that of the other systems as summarized in Table 3. Although the pressure of the system was not the same in each absorption process, the effect of pressure on the KL was not significantly observed when the low pressure was used (1–2 bar). This was in line with several reports [10,34]. Note that, the effect of pressure on the KL would be noticeable when the absorption system is operated at pressures exceeding 10 bar [35].
As compared among different absorption systems, our system offered the largest KLa coefficient for CO2 absorption in water (0.02–0.29 s−1), implying superior CO2 capture efficiency. This was due to the unique characteristics of microchannels. For instance, high surface-to-volume ratio of microchannel (3400–9000 m2 m−3), which is considerably larger than the other absorption devices, promotes heat and mass transfer rates [25,34]. This was in line with the KLa results and surface-to-volume ratio of the other absorbers such as packed column (0.0127 s−1, 10–350 m2 m−3) [36] and stirred tank (0.0056–0.0333 s−1, 100–2000 m2 m−3) [37]. Our work also presents a superior absorption performance when compared to the physical-chemical absorption in a microchamber reported by Zhu et al. [33] who investigated the CO2 capture efficiency under a broad range of gas volumetric flow rate (up to 300 mL min−1). Note that when the extremely short absorption time was applied, the physical absorption was dominant especially in the region of the high volumetric flow rate of gas and liquid. This was owing to the fact that the contact time in our system was longer, allowing for better performance. Lower absorption efficiency of our system compared to that of the hollow fiber membrane was because of the lower interfacial area of gas and liquid; however, the difficulty related to the flow bypassing and channeling of the liquid is a trade-off.
In terms of water utilization, our system offered much lower flow ratio between liquid and gas. Regarding the footprint, the size of absorber can be substantially reduced using microchannel absorber as compared to the conventional packed bed. For example, based on our experimental results, only a small bundle of approximately 454 microchannels operating in parallel can handle the input gas flow rate of 6.8 × 104 mL min−1. This small footprint also offers the potential for reduced power consumption, investment and maintenance cost. Therefore, microchannel can be effectively applied as an alternative device for CO2 capture applications.

3.7. Physicochemical Absorption

The findings from our experiments indicate that the maximum absorption efficiency of 70.9% (gas purity of 84%) was obtained by using water as a green absorbent. Since the purity of product gas at the maximum absorption efficiency was lower than 90%, we then extended our findings to the physicochemical absorption using MEA, where the required amount of solvent used significantly affects the process viability. Note that, in general chemisorption process, relatively high concentration of solvent is required to achieve high CO2 absorption efficiency, i.e., 30 wt.% of MEA solution [3,4]. Our assumption was that a slight addition of chemical solvent (much less than what is commonly used) to our CO2 absorption system would fulfill the requirement in terms of absorption efficiency.
To verify this statement, we adapted our CO2 absorption system by adding a little amount of MEA into the water. Three sets of experiment were carried out to demonstrate the possibility of using this system to further enhance the absorption efficiency. The experiments were carried out at constant temperature of 40 °C, MEA concentration of 2.2 M (1:1 molar ratio of MEA-to-CO2), liquid flow rate of 1 mL min−1, and volumetric flow rate of gas of 150 mL min−1, while the CO2 fraction (40–60 vol.%) were varied. As shown in Figure 7, the addition of small amount of MEA could significantly enhance the %absorption by more than 37% for various levels of CO2 feed concentration. For example, for the CO2 fraction of 40%, using MEA of 1 mol% could enhance the %absorption from 56.2% to 80.2%, suggesting that small amount of solvent was adequate to achieve high absorption efficiency. It is also possible to extend our findings for the application CO2 absorption using wastewater containing ammonia which can be related to many industries such as fertilizer, rubber processing, leather manufacturing, etc. However, this requires further investigation on the effect of operating conditions and characteristics of absorbent on the efficiency of CO2 absorption.

4. Conclusions

The findings of this work indicate that the absorption of CO2 from the gas using water in a microchannel contactor was efficient compared to other absorbers in terms of the overall liquid-side volumetric mass transfer coefficient. The influence of operating conditions including temperature, gas and liquid volumetric flow rate on the %absorption and overall liquid-side volumetric mass transfer coefficient was significant. The maximum %absorption of 70.9% was achieved at temperature of 50 °C, CO2 fraction of 53.7%, total volumetric flow rate of gas of 150 mL min−1, and volumetric flow rate of water of 1 mL min−1. It was also demonstrated that this system can be much improved by adding little amount of MEA in the liquid absorbent. The %absorption up to 88% was achieved by using 2.2 M of MEA (1:1 molar ratio of MEA-to-CO2).

Author Contributions

Conceptualization, N.A., A.K. and A.J.; methodology, N.A., A.K. and A.J.; investigation N.A., A.K. and A.J.; resources, A.J.; writing—original draft preparation, N.A.; writing—review and editing, N.A., W.N., A.K. and A.J.; visualization, N.A., W.N. and A.J.; supervision, A.K. and A.J.; project administration, A.J.; funding acquisition, A.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Kasetsart University Research and Development Institute (KURDI).

Acknowledgments

Financial support from Kasetsart University Research and Development Institute was acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature and Units

aInterfacial area per reactor volume, m2 m−3
CCO2 concentrations, M
CaCapillary number
dAbsorber diameter, m
DDiffusivity, m2 s−1
HHenry’s Law constant, mol Pa−1 m−3
hHeight of absorber, m
kGGas-side mass transfer coefficient, mol m−2 s−1 Pa−1
KLOverall liquid-side mass transfer coefficient, m s−1
KLaOverall gas-liquid volumetric mass transfer coefficient, s−1
kLLiquid-side mass transfer coefficient, m s−1
nMolar flow rate, mol h−1
N CO 2 CO2 absorption flux, mol m−2 h−1
PPartial pressure, Pa
QVolumetric flow rate, m3 s−1
ReReynold number
ScSchmidt number
ShSherwood number
Sh*Modified Sherwood number
TTemperature, °C
VRReactor volume, m3
LHSVLiquid Hourly Space Velocity, h−1
GHSVGas Hourly Space Velocity, h−1
LMPDLogarithmic mean pressure difference, Pa
μViscosity, kg m−1 s−1
ρDensity, kg m−3
GGas phase
LLiquid phase
CO2Carbon dioxide
H2OWater
H2OWater
inInlet
outOutlet
iAt interface
bulkAt bulk phase

References

  1. Noorollahi, Y.; Kheirrouz, M.; Asl, H.F.; Yousefi, H.; Hajinezhatd, A. Biogas production potential from livestock manure in Iran. Renew. Sustain. Energ. Rev. 2015, 50, 748–754. [Google Scholar] [CrossRef]
  2. Khraisheh, M.; Mukherjee, S.; Kumar, A.; Momani, F.A.; Walker, G.; Zaworotko, M.J. An overview on trace CO2 removal by advanced physisorbent materials. J. Environ. Manag. 2020, 255, 109874. [Google Scholar] [CrossRef] [PubMed]
  3. Pashaei, H.; Ghaemi, A. CO2 absorption into aqueous diethanolamine solution with nano heavy metal oxide particles using stirrer bubble column: Hydrodynamics and mass transfer. J. Environ. Chem. Eng. 2020, 8, 104110. [Google Scholar] [CrossRef]
  4. Ma, D.; Zhu, C.; Fu, T.; Yuan, X.; Ma, Y. An effective hybrid solvent of MEA/DEEA for CO2 absorption and its mass transfer performance in microreactor. Sep. Purif. Technol. 2020, 242, 116795. [Google Scholar] [CrossRef]
  5. Sahraie, S.; Rashidi, H.; Sheyda, P.V. An optimization framework to investigate the CO2 capture performance by MEA: Experimental and statistical studies using Box-Behnken design. Process Saf. Environ. Prot. 2019, 122, 161–168. [Google Scholar] [CrossRef]
  6. Hemmati, A.; Rashidi, H.; Behradfar, K.; Kazem, A. A comparative study of different mass transfer and liquid hold-up correlations in modeling CO2 absorption with MEA. J. Nat. Gas Sci. Eng. 2019, 62, 92–100. [Google Scholar] [CrossRef]
  7. Chen, P.C.; Lin, S.Z. Optimization in the absorption and desorption of co2 using sodium glycinate solution. Appl. Sci. 2018, 8, 2041. [Google Scholar] [CrossRef] [Green Version]
  8. Xu, M.; Wang, S.; Xu, L. Screening of physical-chemical biphasic solvents for CO2 absorption. Int. J. Greenh. Gas Control 2019, 85, 199–205. [Google Scholar] [CrossRef]
  9. Aghel, B.; Heidaryan, E.; Sahraie, S.; Mir, S. Application of the microchannel reactor to carbon dioxide absorption. J. Clean. Prod. 2019, 231, 723–732. [Google Scholar] [CrossRef]
  10. Akkarawatkhoosith, N.; Kaewchada, A.; Jaree, A. High-throughput CO2 capture for biogas purification using monoethanolamine in a microtube contactor. J. Taiwan Inst. Chem. Eng. 2019, 98, 113–123. [Google Scholar] [CrossRef]
  11. Remacha, M.J.N.; Kulkarni, A.A.; Jensen, K.F. Gas-liquid flow and mass transfer in an advanced-flow reactor. Ind. Eng. Chem. Res. 2013, 52, 8996–9010. [Google Scholar] [CrossRef]
  12. Luo, X.; Hartono, A.; Svendsen, H.F. Comparative kinetics of carbon dioxide absorption in unloaded aqueous monoethanolamine solutions using wetted wall and string of discs columns. Chem. Eng. Sci. 2012, 82, 31–43. [Google Scholar] [CrossRef]
  13. Yusof, S.M.M.; Lau, K.K.; Shariff, A.M.; Tay, W.H.; Mustafa, N.F.A.; Lock, S.S.M. Novel continuous ultrasonic contactor system for CO2 absorption: Parametric and optimization study. J. Ind. Eng. Chem. 2019, 79, 279–287. [Google Scholar] [CrossRef]
  14. Wangfeng, C.; Jiao, Z.; Xubin, Z.; Yan, W. Enhancement of CO2 absorption under taylor flow in the presence of fine particles. Chin. J. Chem. Eng. 2013, 21, 135–143. [Google Scholar]
  15. Luo, X.; Hartono, A.; Hussain, S.; Svendsen, H.F. Mass transfer and kinetics of carbon dioxide absorption into loaded aqueous monoethanolamine solutions. Chem. Eng. Sci. 2015, 123, 57–69. [Google Scholar] [CrossRef]
  16. Versteeg, G.F.; Swaaij, W.P.M.V. Solubility and diffusivity of acid gases (COP, N20) in aqueous alkanolamine solutions. J. Chem. Eng. Data 1988, 33, 29–34. [Google Scholar] [CrossRef] [Green Version]
  17. Karlsson, H.; Svensson, H. Rate of absorption for CO2 absorption systems using a wetted wall column. Energy Procedia 2017, 114, 2009–2023. [Google Scholar] [CrossRef]
  18. Bhaduri, G.A.; Siller, L. Nickel nanoparticles catalyse reversible hydration of carbon dioxide for mineralization carbon capture and storage. Catal. Sci. Technol. 2013, 3, 1234–1239. [Google Scholar] [CrossRef]
  19. Cadogan, S.P.; Maitland, G.C.; Trusler, J.P.M. Diffusion Coefficients of CO2 and N2 in Water at Temperatures between 298.15 K and 423.15 K at Pressures up to 45 MPa. J. Chem. Eng. Data 2014, 59, 519–525. [Google Scholar] [CrossRef] [Green Version]
  20. Mansourizadeh, A.; Ismail, A.F.; Matsuura, T. Effect of operating conditions on the physical and chemical CO2 absorption through the PVDF hollow fiber membrane contactor. J. Membr. Sci. 2010, 353, 192–200. [Google Scholar] [CrossRef]
  21. Diamond, L.W.; Akinfiev, N.N. Solubility of CO2 in water from −1.5 to 100 °C and from 0.1 to 100 MPa: Evaluation of literature data and thermodynamic modelling. Fluid Phase Equilibria 2003, 208, 265–290. [Google Scholar] [CrossRef]
  22. Hassan, I.; Vaillancourt, M.; Pehlivan, K. Two-phase flow regime transitions in microchannels: A comparative experimental study. Microscale Thermophys. Eng. 2005, 9, 165–182. [Google Scholar] [CrossRef]
  23. Kashid, M.N.; Renken, A.; Minsker, L.K. Gas-liquid and liquid-liquid mass transfer in microstructured reactors. Chem. Eng. Sci. 2011, 66, 3876–3897. [Google Scholar] [CrossRef]
  24. Shao, N.; Gavriilidis, A.; Angeli, P. Flow regimes for adiabatic gas-liquid flow in microchannels. Chem. Eng. Sci. 2009, 64, 2749–2761. [Google Scholar] [CrossRef]
  25. Yue, J.; Chen, G.; Yuan, Q.; Luo, L.; Gonthier, Y. Hydrodynamics and mass transfer characteristics in gas-liquid flow through a rectangular microchannel. Chem. Eng. Sci. 2007, 62, 2096–2108. [Google Scholar] [CrossRef]
  26. Sheyda, P.V.; Afshari, A. A detailed screening on the mass transfer modeling of the CO2 absorption utilizing silica nanofluid in a wetted wall column. Process Saf. Environ. Prot. 2019, 127, 125–132. [Google Scholar] [CrossRef]
  27. Puxty, G.; Rowland, R.; Attalla, M. Comparison of the rate of CO2 absorption in to aqueous ammonia and monoethanolamine. Chem. Eng. Sci. 2010, 65, 915–992. [Google Scholar] [CrossRef]
  28. Arachchige, U.S.P.R.; Aryal, N.; Eimer, D.A.; Melaaen, M.C. Viscosities of pure and aqueous solutions of monoethanolamine (MEA), diethanolamine (DEA) and N-methyldiethanolamine (MDEA). Nord. Rheol. Soc. 2013, 21, 299–306. [Google Scholar]
  29. Li, C.; Zhu, C.; Ma, Y.; Liu, D.; Gao, X. Experimental study on volumetric mass transfer coefficient of CO2 absorption into MEA aqueous solution in a rectangular microchannel reactor. Int. J. Heat Mass Transf. 2014, 78, 1055–1059. [Google Scholar] [CrossRef]
  30. Ji, X.Y.; Ma, Y.G.; Fu, T.T.; Zhu, C.Y.; Wang, D.J. Experimental investigation on liquid volumetric mass transfer coefficient for upward gas-liquid two-phase flow in rectangular microchannels. Br. J. Chem. Eng. 2010, 27, 573–582. [Google Scholar] [CrossRef] [Green Version]
  31. Isa, F.; Suleman, H.; Zabiri, H.; Maulud, A.S.; Ramasamy, M.; Tufa, L.D.; Shariff, A.M. An overview on CO2 removal via absorption: Effect of elevated pressures in counter-current packed column. J. Nat. Gas Sci. Eng. 2016, 33, 666–677. [Google Scholar] [CrossRef]
  32. Guangwen, C.; Jun, Y.; Quan, Y. Gas-Liquid Microreaction Technology: Recent Developments and Future Challenges. Chin. J. Chem. Eng. 2008, 16, 663–669. [Google Scholar]
  33. Zhu, K.; Yao, C.; Liu, Y.; Chen, G. Theoretical approach to CO2 absorption in microreactors and reactor volume prediction. Chem. Eng. Process. 2020, 150, 107904. [Google Scholar] [CrossRef]
  34. Elhajj, J.; Hindi, M.A.; Azizi, F. A Review of the absorption and desorption processes of carbon dioxide in water systems. Ind. Eng. Chem. Res. 2014, 53, 2–22. [Google Scholar] [CrossRef]
  35. Yao, C.; Dong, Z.; Zhao, Y.; Chen, G. Gas-liquid flow and mass transfer in a microchannel under elevated pressures. Chem. Eng. Sci. 2015, 123, 137–145. [Google Scholar] [CrossRef]
  36. Charpentier, J.C. Mass-transfer rates in gas–liquid absorbers and reactors. Adv. Chem. Eng. 1981, 11, 1–133. [Google Scholar]
  37. Kies, F.K.; Benadda, B.; Otterbein, M. Experimental study on mass transfer of a co-current gas–liquid contactor performing under high gas velocities. Chem. Eng. Process. 2004, 43, 1389–1395. [Google Scholar] [CrossRef]
  38. Salimi, J.; Salimi, F. CO2 capture by water-based Al2O3 and Al2O3-SiO2 mixture nanofluids in an absorption packed column. Rev. Mex. Ing. Quim. 2016, 15, 185–192. [Google Scholar]
  39. Hill, G.A. measurement of overall volumetric mass transfer coefficients for carbon dioxide in a well-mixed reactor using a pH probe. Ind. Eng. Chem. Res. 2006, 45, 5796–5800. [Google Scholar] [CrossRef]
  40. Wang, F.; Kang, G.; Liu, D.; Li, M.; Cao, Y. Enhancing CO2 Absorption Efficiency Using a Novel PTFE Hollow Fiber Membrane Contactor at Elevated Pressure. AIChE J. 2018, 64, 2135–2145. [Google Scholar] [CrossRef]
  41. Belaissaoui, B.; Baro, J.C.; Hernando, A.L.; Zaidiza, D.A.; Chabanon, E.; Castel, C.; Rode, S.; Roizard, D.; Favre, E. Potentialities of a dense skin hollow fiber membrane contactor for biogas purification by pressurized water absorption. J. Membr. Sci. 2016, 513, 236–249. [Google Scholar] [CrossRef]
  42. Vandu, C.O.; Liu, H.; Krishna, R. Mass transfer from Taylor bubbles rising in single capillaries. Chem. Eng. Sci. 2005, 60, 6430–6437. [Google Scholar] [CrossRef]
  43. Constantinou, A.; Gavriilidis, A. CO2 absorption in a microstructured mesh reactor. Ind. Eng. Chem. Res. 2010, 49, 1041–1049. [Google Scholar] [CrossRef]
Figure 1. Equipment and experimental set-up for CO2 absorption process.
Figure 1. Equipment and experimental set-up for CO2 absorption process.
Energies 13 05465 g001
Figure 2. Main effect of absorption variables on the responses; (a) absorption efficiency, (b) overall liquid-side mass transfer coefficient.
Figure 2. Main effect of absorption variables on the responses; (a) absorption efficiency, (b) overall liquid-side mass transfer coefficient.
Energies 13 05465 g002
Figure 3. Effects of total flow rate of gas and liquid on the responses: (a) absorption efficiency, (b) overall liquid-side mass transfer coefficient.
Figure 3. Effects of total flow rate of gas and liquid on the responses: (a) absorption efficiency, (b) overall liquid-side mass transfer coefficient.
Energies 13 05465 g003
Figure 4. Effects of CO2 fraction and absorption temperature on the responses: (a) absorption efficiency, (b) overall liquid-side mass transfer coefficient.
Figure 4. Effects of CO2 fraction and absorption temperature on the responses: (a) absorption efficiency, (b) overall liquid-side mass transfer coefficient.
Energies 13 05465 g004
Figure 5. Interaction effect of variable pairs; (a) volumetric flow rate of gas and liquid, (b) volumetric flow rate of gas and CO2 fraction, (c) volumetric flow rate of liquid and temperature.
Figure 5. Interaction effect of variable pairs; (a) volumetric flow rate of gas and liquid, (b) volumetric flow rate of gas and CO2 fraction, (c) volumetric flow rate of liquid and temperature.
Energies 13 05465 g005
Figure 6. Parity plot of responses; (a) %absorption efficiency (%), (b) liquid-side volumetric mass transfer coefficient (s−1).
Figure 6. Parity plot of responses; (a) %absorption efficiency (%), (b) liquid-side volumetric mass transfer coefficient (s−1).
Energies 13 05465 g006
Figure 7. Effect of MEA addition on the %absorption under various CO2 fractions.
Figure 7. Effect of MEA addition on the %absorption under various CO2 fractions.
Energies 13 05465 g007
Table 1. Experimental range of absorption variables.
Table 1. Experimental range of absorption variables.
VariableUnitSymbolRange and Level
−101
Independent Variables
CO2 Fractionvol.%F405060
Total gas volumetric flow ratemL min−1G150175200
Liquid volumetric flow ratemL min−1L1.01.52.0
Temperature°CT304050
Dependent Variable
%Absorption%E%
Table 2. ANOVA results for 3k factorial design.
Table 2. ANOVA results for 3k factorial design.
VariableDFSeq SSAdj SSAdj MSp-Value
F2216.96216.96108.480.144
G26587.366587.363293.680
L212,334.7912,334.796167.40
T2665.8665.8332.90.008
F × G4314.59314.5978.650.225
F × L452.5952.5913.150.895
F × T4302.21302.2175.550.241
G × L41410.611410.61352.650.002
G × T4282.32282.3270.580.27
L × T4146.72146.7236.680.577
F × G × L8216.02216.02270.805
F × G × T8920.61920.61115.080.071
F × L × T8495.67495.6761.960.332
G × L × T8634.36634.3679.30.2
Error16790.51790.5149.41
Total8025,371.12
Table 3. Comparison of KLa and efficiency for absorption system with the various reactor types.
Table 3. Comparison of KLa and efficiency for absorption system with the various reactor types.
ReactorConditionsKLa (s−1)Efficiency (%)Reference
Packed TowerSystem: CO2-water
Diameter of Column: 30 mm
Length of column: 900 mm
Gas flow rate: 1000 mL min−1
Liquid flow rate: 300 mL min−1
Temperature: 25 °C
Pressure: 1 bar
GHSV: 93.75 h−1
LHSV: 28.125 h−1
Contact Time: 30 s
0.0055-[38]
Well-mixed ReactorSystem: CO2-water
Impeller speed: 150–600 rpm
Gas flow rate: 200–2000 mL min−1
CO2 concentration: 10 vol.%
Temperature: 15–40 °C
Pressure: 0.003–0.2 bar
GHSV: 4.9–49 h−1
Contact Time: N/A
0.0056–0.0333-[39]
Hollow Fiber
Membrane
System: CO2-water
Length 580 mm, diameter 25 mm
Liquid Flow Rate: 0.2–0.8 L min−1
Gas Flow Rate: 0.7–2.8 L min−1
CO2 concentration: 40 vol.%
Temperature: 25 °C
-55–97[40]
Hollow Fiber
Membrane
System: CO2-water
Length 240 mm, diameter 36 mm
Liquid Flow Rate: 3 × 10−3–1 × 10−2 m s−1
Gas Flow Rate: 1 × 10−4–4 × 10−4 m s−1
CO2 concentration: 30 vol.%
Temperature: 22 °C
-10–80[41]
MicrotubeSystem: Air-water
Diameter of channel: 1 mm
Length of channel: 200 mm
Temperature: 25 °C
Pressure: 1 bar
GHSV: 1782 h−1
LHSV: 15,330 h−1
Contact Time: 1.08 s
0.38-[42]
MicrochannelSystem: CO2-NaOH
Width of channel: 5.48 mm
Depth of channel: 1.05 mm
Length of channel: 90 mm
Liquid Flow Rate: 1.2–2.5 mL min−1
Gas Flow Rate: 177–354 mL min−1
CO2 concentration: 20 vol.%
NaOH concentration: 2 M
Temperature: 20 °C
Pressure: 1 bar
-15–50[43]
Microchannel aSystem: CO2-DEA
Diameter of channel: 0.6 mm
Length of channel: 100 mm
Liquid Flow Rate: 0.9–1.2 mL min−1
Gas Flow Rate: 150–300 mL min−1
CO2 concentration: 16.4 vol.%
DEA concentration: 30 wt.%
Temperature: 25 °C
Pressure: 1 bar
GHSV: 3.18 × 105–6.37 × 105 h−1
LHSV: 1.91 × 103–2.55 × 103 h−1
Contact Time: 3.3 × 10−5 –6.7 × 10−5 s
-5–10[33]
MicrochannelSystem: CO2-water
Diameter of channel: 0.5 mm
Length of channel: 60 mm
Liquid Flow Rate: 1–2 mL min−1
Gas Flow Rate: 150–200 mL min−1
CO2 concentration: 40–60 vol.%
Temperature: 30–50 °C
Pressure: 1.7 bar
GHSV: 6 × 105–8 × 105 h−1
LHSV: 4 × 103–8 × 103 h−1
Contact Time: 4.4 × 10−3–5.9 × 10−3 s
0.02–0.264.8–70.9This work
a Physical-chemical absorption operated at high volumetric flow rate of gas and liquid.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Akkarawatkhoosith, N.; Nopcharoenkul, W.; Kaewchada, A.; Jaree, A. Mass Transfer Correlation and Optimization of Carbon Dioxide Capture in a Microchannel Contactor: A Case of CO2-Rich Gas. Energies 2020, 13, 5465. https://doi.org/10.3390/en13205465

AMA Style

Akkarawatkhoosith N, Nopcharoenkul W, Kaewchada A, Jaree A. Mass Transfer Correlation and Optimization of Carbon Dioxide Capture in a Microchannel Contactor: A Case of CO2-Rich Gas. Energies. 2020; 13(20):5465. https://doi.org/10.3390/en13205465

Chicago/Turabian Style

Akkarawatkhoosith, Nattee, Wannarak Nopcharoenkul, Amaraporn Kaewchada, and Attasak Jaree. 2020. "Mass Transfer Correlation and Optimization of Carbon Dioxide Capture in a Microchannel Contactor: A Case of CO2-Rich Gas" Energies 13, no. 20: 5465. https://doi.org/10.3390/en13205465

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