Next Article in Journal
Special Issue “Natural Plant Substances—Structural and Application Aspects: A Theme Issue in Honor of Professor Wieslaw Oleszek”
Next Article in Special Issue
Trends on CO2 Capture with Microalgae: A Bibliometric Analysis
Previous Article in Journal
Merging Electron Deficient Boronic Centers with Electron-Withdrawing Fluorine Substituents Results in Unique Properties of Fluorinated Phenylboronic Compounds
Previous Article in Special Issue
Do Bio-Ethanol and Synthetic Ethanol Produced from Air-Captured CO2 Have the Same Degree of “Greenness” and Relevance to “Fossil C”?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization and Application of Amine-Functionalized Hierarchically Micro-Mesoporous Silicon Composites for CO2 Capture in Flue Gas

1
School of Ecological Environment and Urban Construction, Fujian University of Technology, Fuzhou 350118, China
2
Fuzhou Smart Environmental Industry Technology Innovation Center, Fuzhou 350118, China
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(11), 3429; https://doi.org/10.3390/molecules27113429
Submission received: 7 May 2022 / Revised: 23 May 2022 / Accepted: 24 May 2022 / Published: 26 May 2022
(This article belongs to the Special Issue The CO2 Economy: CO2 Capture and Reuse Technologies)

Abstract

:
An efficient CO2 adsorbent with a hierarchically micro-mesoporous structure and a large number of amine groups was fabricated by a two-step synthesis technique. Its structural properties, surface groups, thermal stability and CO2 adsorption performance were fully investigated. The analysis results show that the prepared CO2 adsorbent has a specific hierarchically micro-mesoporous structure and highly uniformly dispersed amine groups that are favorable for the adsorption of CO2. At the same time, the CO2 adsorption capacity of the prepared adsorbent can reach a maximum of 3.32 mmol-CO2/g-adsorbent in the actual flue gas temperature range of 303–343 K. In addition, the kinetic analysis results indicate that both the adsorption process and the desorption process have rapid adsorption/desorption rates. Finally, the fitting of the CO2 adsorption/desorption experimental data by Avrami’s fractional kinetic model shows that the CO2 adsorption rate is mainly controlled by the intra-particle diffusion rate, and the temperature has little effect on the adsorption rate.

1. Introduction

For climate change, the ultimate global goal is to avoid dangerous disturbances to the climate system. To achieve this goal, UNFCCC member states have shown strong support for measurable guardrail targets, such as “Avoid 1.5 °C” or “Avoid 450 ppm CO2”. However, the CO2 concentration in the atmosphere has increased sharply from 340 ppm in 1980 to 408 ppm in 2019 [1], and if existing human activities do not sufficiently and quickly change, the global temperature increase will exceed 1.5 °C. Therefore, a variety of methods should be adopted to reduce CO2 levels, such as planting more trees, replacing fossil energy with renewable energy, improving the efficiency of coal-fired plants, and using CO2 capture and storage (CCS) to reduce CO2 content [2,3].
CCS technology is an attractive method because it helps reduce the release of CO2 into the environment and allows the continued use of coal to meet the world’s energy needs [2,4]. According to different coal-fired plant configurations, three main approaches can be applied to reduce CO2 emissions from the flue gas of these plants, including pre-combustion capture, oxyfuel combustion, and post-combustion capture [5]. Among these CCS technologies, post-combustion capture technology (which uses wet/dry adsorbents to absorb CO2 from the flue gas that is produced by the combustion of fossil fuels) is the most popular industrial method [6]. Post-combustion capture technologies generally include absorption [7], porous materials adsorption [8], membrane separation [9], cryogenics and hydration etc. [5]. At present, some experiments and computational studies have been conducted on the above-mentioned technologies [10,11,12].
Among these technologies, the adsorption-based technology is the most attractive because it has the potential to lower costs while avoiding the defects of aqueous amine absorption [13]. Adsorbents are the foundation for the adsorption process. In recent years, many porous solid adsorbents have been investigated in order to capture CO2 from flue gas, including carbon materials [14,15,16], zeolites [17], mesoporous silicon [18], pillared lamellar clays [19], and the metal-organic frameworks (MOFs) [20,21].
The ideal solid adsorbent for CO2 capture should have good CO2 selectivity, a high CO2 adsorption capacity, and a faster CO2 adsorption/desorption rate [22]. The faster the CO2 adsorption/desorption process the more economical the process of CCS [23]. Usually, the microporous materials, such as the MCM series and ZSM series, have a large CO2 adsorption capacity because of their high surface area and high porosity [24]. However, due to their rapid decrease in adsorption capacity at higher temperatures, these microporous materials are mainly used at low temperatures (below 303 K) [25]. Mesoporous materials such as SBA series and KIT series have unique mass-transfer properties due to their large pore diameters [26]. It is therefore expected that a new type of micro-mesoporous composite material with the advantages of high surface area and rapid mass transfer can be prepared through the combination of microporous and mesoporous materials production. In addition, this material may display a better CO2 adsorption performance after surface amine functionalization; this is because amines are nucleophilic, and they can strongly interact with electrophilic CO2 through nucleophilic substitution reactions to increase the CO2 adsorption capacity.
Therefore, the objective of this study is to synthesize an amine-functionalized hierarchically micro-mesoporous adsorbent and discuss its CO2 adsorption behavior. In order to achieve this target, firstly, the nano-scale microporous silicon precursor was assembled into an ordered cubic mesoporous structure by a two-step hydrothermal crystallization method to synthesize the micro-mesoporous silicon material. Then, the prepared micro-mesoporous silicon material was used as a support, an amino compound was used as a modifier, and the amine-functionalized hierarchically micro-mesoporous silicon composites was fabricated by an impregnation method. After that, the pore structure, surface characteristics and CO2 adsorption performance of the prepared adsorbent were analyzed. Finally, the experimental data of isothermal CO2 adsorption/desorption was simulated by a mathematical model, and the kinetic characteristics of CO2 adsorption/desorption were discussed.

2. Materials and Methods

2.1. Materials

Tetrapropylammonium hydroxide solution (TPAOH); aluminium isopropoxide; tetraethylorthosilicate (TEOS); polyethylene-polypropylene glycol (P123); hydrochloric acid (HCl); and n-Butanol were used for the synthesis of a hierarchically micro-mesoporous silicon.
Ethylenediamine (EDA); diethylenetriamine (DETA); tetraethylenepentamine (TEPA), pentaethylenehexamine (PEHA); ethylene imine polymer (PEI); and ethanol were used for the synthesis of an amine-functionalized hierarchically micro-mesoporous adsorbent. The chemicals used in the experiment are listed in Table 1. All chemicals were used without further purification.

2.2. Sample Preparation

2.2.1. Synthesis of Support Material

The preparation of the hierarchically micro-mesoporous silicon material was as follows: First, 4.54 g of TPAOH was dissolved in 16.87 g of deionized water at room temperature and stirred for 30 min, and then 0.17 g of aluminum isopropoxide and 0.1 g of NaOH were added. After stirring for 30 min, the mixture was heated to 303 K, and 8.5 g of TEOS was dropped into the mixture. After 16 h of vigorous stirring, the final solid–liquid mixture was transferred to a Teflon kettle and subjected to hydrothermal treatment at 423 K for 24 h. The microcrystalline emulsion was then cooled to room temperature and the final resultant was labeled as microcrystalline emulsion A.
Second, 4.0 g of P123 (5800) was dissolved in a mixture of 144 g of H2O and 7.9 g of hydrochloric acid solution (35 wt.%) at 313 K under stirring. After the P123 was completely dissolved, 4.0 g of n-Butanol was added. After stirring continuously for 1 h, the above microcrystalline emulsion A was added and sonicated for 30 min. After vigorously stirring at 313 K for 24 h the resultant was transferred to a Teflon kettle and subjected to hydrothermal treatment at 373 K for 24 h. The solid–liquid mixture was filtered to obtain a solid product and the solid product was dried at 373 K for 10 h. After that, the obtained solid sample was calcined at 823 K for 6 h under a temperature programmed system (5 K/min) to remove the template. The resulting material is herein denoted as micro-mesoporous silicon (MMS).

2.2.2. Preparation of Amine-Functionalized Hierarchically Micro-Mesoporous Adsorbent

The amine-loaded hierarchically micro-mesoporous adsorbent was produced by a wet impregnation method. The ethylenediamine (EDA); diethylenetriamine (DETA); tetraethylenepentamine (TEPA); pentaethylenehexamine (PEHA); and polyetherimide (PEI) were used as amine-modified materials. In a typical experiment, 1 g of amine-modified material was dissolved in 50 mL of ethanol at room temperature, and after stirring for 30 min, 1 g of MMS was added, and then the mixture was stirred for 2 h. After that, the mixture was evaporated at 353 K and then dried in air at 373 K for 1 h. Finally, a white composite was obtained. The obtained composites are denoted as AMMS-E, AMMS-D, AMMS-T, AMMS-P and AMMS-PEI, respectively. Here, the AMMS represents amine-functionalized micro-mesoporous silicon, and the E, D, T, P, and PEI stand for EDA, DETA, TEPA, PEHA, and PEI, respectively.

2.3. Characterization

The surface crystallinity of the AMMS was analyzed by the X-ray diffraction (XRD) measurement using a Rigaku powder diffractometer (D8 ADVANCE, Bruker, Bremen, Germany) with Cu Kα radiation (λ = 0.15406 nm). The tube voltage was 45 kV and the current was 40 mA. The XRD diffraction patterns were taken in the 2θ range of 0.5–10° and 5–60° at a scan speed of 2°/min. The surface morphology was determined with a scanning electron microscope (S4800, Hitachi, Tokyo, Japan).
The surface area analyzer (Autosorb iQ, The Quantachrome Instruments U.S., Boynton Beach, FL, USA) was used for the nitrogen adsorption/desorption test. Before the measurement, the sample was degassed at 573 K under nitrogen flow for 3 h. The surface area of the powder was obtained using the Brunauer-Emmett-Teller (BET) method. The pore size distributions were calculated by the Barrett-Joyner-Halenda (BJH) equation. The total pore volume was determined from the amount of adsorbed N2 at P/P0 = 0.99.
A TG (TG 209F3, Netzsch, Selb, Germany) instrument was used to analyze the thermal stability of AMMS, and the thermogravimetric analysis test was performed at a temperature of 303 K to 973 K with a heating rate of 10 K/min under a dynamic N2 atmosphere.
The surface chemical groups were analyzed by Fourier Transform Infrared (FTIR, Nicolet IS50, Thermo Fisher Scientific, Waltham, MA, USA) spectroscopies.

2.4. Adsorption/Desorption of CO2

A TGA device (TG209F3, Netzsch) was used for the CO2 adsorption/desorption test. A total of 10 mg of AMMS samples were loaded into a 0.05 cm3 platinum sample pan. After stabilization, each sample was heated to 423 K in a nitrogen stream (120 cm3 min−1), then cooled to the adsorption temperature, and a CO2 (99.99%)/N2 (99.99%) mixture (10 Vol.% CO2) was introduced until the mass of the sample no longer increased (about 50 min). The adsorption capacity was calculated based on the AMMS sample mass increase using Equation (1).
q a = M 1 M 0 44 × M 0 × 10 3
where qa is the adsorption capacity of CO2, mmol g−1; M0 is the mass of pure adsorbent, and g; M1 is the mass of the sample after adsorbing CO2 g.
Once the adsorption process reached equilibrium, the temperature programmed desorption (TPD) test was performed on the same experimental system. The adsorbent was heated to the required desorption temperature (383 K) in a nitrogen stream (120 cm3 min−1). The total amount of CO2 was calculated based on the loss of sample mass using Equation (1).

2.5. Adsorption/Desorption Kinetic Models

Avrami’s fractional kinetic model (Equation (2)) is used to study the kinetics of CO2 adsorption on AMMS materials. Usually, Equation (2) can be transformed into Equation (3) in the form of desorption component y. Thus, Equation (3) can be employed to further understand the desorption kinetics of the CO2 desorption process [27].
q t = q e [ 1 e ( k a t ) n ]
y = 1 e ( k a t ) n
where q e and q t (mmol·g−1) denote the equilibrium capacity and the adsorption capacity at any time t (s), respectively, k a is the kinetic constant of the Avrami model, the n represents as the Avrami exponent, which is often in fractional form, and reflects possible mechanism changes in the adsorption process [28].

3. Results and Discussions

3.1. Characterization of the AMMS

3.1.1. XRD Analysis

Figure 1 provides the small-angle XRD pattern of the MMS and AMMS powder. The spectrum for the MMS and AMMS powder reveals a strong peak at about 2θ = 0.55° that corresponds to the (100) crystal face, indicating that the composites materials MMS and AMMS may have a p6 mm hexagonal-symmetry mesoporous structure that is similar to SBA-15 [29]. However, we cannot find the (110) and (200) peaks that are normally observed in SBA-15 from the spectra of the MMS and AMMS powders, which may indicate that the mesoporous system of MMS and AMMS has become disordered [30]. This disorder may be attributed to the formation of microporous structures that are similar to ZSM-5 materials in MMS and AMMS materials. Figure 1 also displays the XRD patterns in the wide-angle of the MMS and AMMS powder. The sharp peaks that are clearly visible at 2θ = 7.9°, 8.8°, 20.3°, 23.1° and 23.9° correspond to (101), (200), (103), (501), and (303) crystal plains and consistent with the patterns of ZSM-5 that are reported in the literature [31].
The locations of the characteristic Bragg diffraction peaks of the MMS and AMMS samples are almost the same, but the peak intensity (XRD patterns in wide-angle) of the AMMS is reduced, indicating that the amine-modified material is loaded into the micro-pores of MMS. At the same time, the micro-mesoporous structure of MMS is retained.

3.1.2. SEM Analysis

The surface morphology of MMS and AMMS-T is presented in Figure 2, and the SEM images of other AMMS are shown in Figure S1. The SEM micrograph of MMS shows that this solid is formed by the aggregation of small amorphous material units (like the ZSM-5 material [31], and a certain number of pores are formed between the units. In contrast to the surface morphology of the MMS material, the SEM micrograph of AMMS-T shows that the surface of AMMS-T is distributed with highly dispersed and uniform particles. This result indicates that the amine-modified material was successfully impregnated into the pores of the MMS material, and the loading of the amine-modified material resulted in a decrease in the surface area and pore volume of the structure, which was further confirmed in the BET surface area analysis. At the same time, the highly uniformly dispersed amine on the solid surface may be beneficial to the adsorption of CO2 [32]. The subsequent CO2 adsorption analysis also proved this possibility.

3.1.3. BET Analysis

Through nitrogen adsorption experiments at 77 K, the structural characteristics of MMS and AMMS, such as surface area, average pore size and pore volume were analyzed. The nitrogen-adsorption isotherms of MMS and AMMS-T are presented in Figure 3, and the nitrogen adsorption isotherms of other AMMS are displayed in Figure S2. All adsorption isotherms represent the typical type IV features of mesoporous silica. The pore size distribution of MMS and AMMS-T are also displayed in Figure 3. The pore size distribution curve of MMS has two obvious peaks at 1.6 nm and 3.5 nm, indicating that its pore structure is mainly composed of micro-pores and meso-pores. However, the pore size distribution of AMMS-T is relatively random and irregular. This phenomenon may be caused by the disordered accumulation of amine-modified materials on the surface of the pores during the impregnation process.
Table 2 summarizes the surface area, average pore size and pore volume of all the materials. It was observed that after amine modification, the surface area and pore volume of AMMS were significantly reduced. This can be explained by amine impregnation, where amine material accumulates on the surface of the MMS porous channels, resulting in a decrease in surface area and pore volume. These results are consistent with the XRD measurement results.

3.1.4. TG Analysis

The thermal behavior of AMMS was studied by TG analysis. Figure 4 shows the mass loss curve of AMMS. Two stages of mass loss are observed in the TG curve of AMMS. Due to the removal of adsorbed water and CO2, the mass loss of 3–8 wt.% in the first stage is approximately between 303 K and 373 K. We noticed that when the temperature is higher than 373 K, the mass loss curves display a relatively straight line within a certain temperature range, indicating that there is basically no mass loss in the sample within this temperature range. For AMMS-T, the temperature range is 373 K to 423 K, AMMS-P is 373 K to 483 K, and AMMS-PEI is 373 K to 543 K. As the temperature continues to rise, the samples begin to undergo the second stage of the mass loss process. The results suggest that AMMS-T, AMMS-E, AMMS-D, AMMS-P and AMMS-PEI have a thermal stability of 423 K, 483 K, 403 K, 483 K, and 543 K, respectively. Finally, when the sample mass remains constant, the total mass loss of AMMS-T, AMMS-P and AMMS-PEI is about 55%, 61% and 55%, respectively. The reason for the different total mass loss of the samples may be attributed to the volatilization and decomposition of different amine-modified materials.

3.1.5. FTIR Analysis

The FTIR spectrums of MMS and AMMS in the domain of 400–4000 cm−1 are demonstrated in Figure 5. For MMS, the adsorption peak around 462 cm−1, 804 cm−1 and 1092 cm−1 could be attributed to the Si-O-Si bending vibrations [33], the symmetric stretching vibrations of Si–O–Si bonds [33], and the asymmetric stretching vibrations of Si–O–Si bonds [34]. Moreover, the adsorption band at 552 cm−1 represents the typical vibration band of five- or six-membered rings of X–O–X, where X can be Al or Si [35]. In addition, the adsorption band at 972 cm−1 is ascribed to the defective Si–OH group [36]. Furthermore, the adsorption bands at 1632 cm−1 and 3448 cm−1 are due to the H–O–H bending vibration and the typical of O–H stretching vibration, respectively [37].
The FTIR spectrum of AMMS displays some new adsorption peaks at 1314 cm−1, 1475 cm−1, 1589 cm−1, 1673 cm−1, 2839 cm−1, and 2958 cm−1. The peak at 1314 cm−1 is attributed to C–N tensile vibration. The adsorption peaks at 1475 cm−1 and 1589 cm−1 represent N–H stretching vibrations which are associated with asymmetric and symmetric bending of the primary amines (–NH2). In addition, the adsorption peak at 1673 cm−1 is related to the bending of secondary amines (–N(R)H) in amine-modified materials. Moreover, the adsorption peaks at 2839 cm−1 and 2958 cm−1 are due to the CH2 asymmetric and symmetric stretching modes of the amine chain in the AMMS. Compared with the MMS [38], we cannot observe the peak at 1632 cm−1 from the FTIR spectrum of AMMS. At the same time, the intensity of the peak at 3448 cm−1 decreases with the impregnation of amine. The reason for this may be due to the interaction of amine and MMS. Overall, the above FTIR analysis indicates that the amine-modified material is effectively loaded into the pores of the MMS.

3.2. Adsorption Properties Analysis

CO2 adsorption experiments were carried out at the required temperature with C0 = 10 vol.% of CO2 (inlet CO2 concentration). The CO2 adsorption capacities (qa) of MMS and AMMS are presented in Table 3. In addition, the adsorption curves of MMS and AMMS-T at three temperatures (303 K, 323 K, and 343 K) are shown in Figure 6, and the adsorption curves of the other AMMS are displayed in Figure S3.
It can be seen from Table 3 that the qa of MMS (at different temperatures) is very limited. At the same time, as the temperature increases from 303 K to 343 K, the qa decreases significantly. This may be due to the exothermic behavior of the adsorption phenomenon. However, the qa of AMMS-T increases with increasing temperature. The reason for this may be that the loaded amine-modified material has higher molecular activity and mobility at higher temperatures, so the qa of AMMS shows an increasing trend when the temperature increases from 303 K to 343 K.
From Table 3, we observed that under similar conditions AMMS-T has the best adsorption capacity (qa = 3.32 mmol-CO2/g-adsorbent at 343 K) among all the AMMS. This result is consistent with the results of the SEM analysis. The previous SEM analysis results indicated that among all the surface morphologies of AMMS samples, the surface particle dispersion of AMMS-T was the most consistent and uniform. It is precisely because of the highly uniform dispersion of the amine-modified material particle that agglomeration of the particles is avoided, so that more amine groups of the amine-modified material can contact and react with CO2 molecules, thereby improving the ability to adsorb CO2.
The adsorption capacity of AMMS-T was compared with the values that were obtained under similar conditions for the CO2 adsorbents proposed in the literature (Table 4). We noticed that AMMS-T exhibited a relatively good adsorption capacity under similar conditions.
The abscissa in Figure 6 denotes the time and the ordinate represents the qa. The overall adsorption curve profiles for 303 K, 323 K, and 343 K display similar adsorptive behavior. All the adsorption processes are completed in a short time, and the adsorption rate is approximately the same. This means that although a higher temperature can increase the CO2 adsorption capacity, the adsorption rate is not controlled by the adsorption temperature. The adsorption rate is further analyzed in the subsequent kinetic analysis of the adsorption process.

3.3. Adsorption/Desorption Kinetics Analysis

Previous adsorption capacity studies have shown that AMMS-T exhibits the best adsorption capacity among all AMMS (qa = 3.32 mmol-CO2/g-adsorbent, at 343 K). Therefore, AMMS-T was used as a CO2 adsorbent to analyze the characteristics of the adsorption/desorption kinetics of the CO2 adsorption/desorption process in this study. Multiple adsorption kinetics models have been employed to quantitatively analyze the adsorption and to explore the adsorption mechanism; for example, Lagergren’s pseudo-first-order model [48]; Ho’s pseudo-second-order model [49]; the classical intracrystalline diffusion model [50]; and Avrami’s fractional-order kinetic model [51]. Among these models, since the fractional order of the Avrami model can be used to characterize the complexity of the reaction mechanism or the simultaneous occurrence of multiple reaction paths [52,53], the Avrami fractional-order kinetic model is well suited to analyzing the CO2 adsorption kinetics and adsorption mechanisms [54]. For the Avrami model, if n = 1, it means that the adsorption process has uniform adsorption [55,56]. If n = 2, then the adsorption process may show a perfect one-dimensional growth at the adsorption point after uniform adsorption [57].
The kinetics character of CO2 adsorption on AMMS-T were investigated by isothermal adsorption at several T (303, 323 and 343 K) and at various C0 (10, 15, 20, and 40 vol.%). The experimental data were simulated using the Avrami model and the results can be found in Figure 7 and Table 5.
As described in Table 5, the adsorption rate constant ka increases with the increase in C0 at the same T. However, the value of ka only slightly changes with the increase in T under the same C0. The high partial pressure of CO2 can promote the faster diffusion of CO2 to the active sites on the surface of the solid pores. Therefore, it can be deduced from the simulation results that ka is mainly controlled by the intra-particle diffusion process, and temperature has little effect on the ka.
It can be seen from Table 5 that most Avrami exponent n are greater than 1 and less than 2, indicating that additional adsorption may be preferentially provided on the existing adsorption sites and resulting in one-dimensional growth at these adsorption points after uniform adsorption. The value of n is less than 2, because as the adsorption progresses, not all the adsorption sites that have adsorbed CO2 start a new uniform one-dimensional adsorption, but only some sites continue to adsorb CO2; the rest do not continue to absorb CO2.
The ka can be demonstrated by the Arrhenius equation:
k a = A e ( E a / R T )
where A denotes the Arrhenius pre-exponential factor, Ea is the activation energy, and R is the universal ideal gas constant. A plot of lnka versus 1/T is shown in Figure 8. Table 6 presents the A and Ea that were obtained by linear regression of the experimental data.
As shown in Table 6, the absolute value of Ea decreases with the increase in C0, indicating that a high CO2 partial pressure promotes the adsorption of CO2. The result is consistent with the previous analysis results, that is, the CO2 adsorption rate is mainly controlled by the intra-particle diffusion process. The Ea of AMMS-T is less than that of liquid ammonia absorption (40−50 kJ·mol−1 [58,59]), which indicates that AMMS-T may be a promising and effective CO2 adsorbent.
Figure 9 displays the TPD curve, the desorption curve at 383 K and the fitting curve of the Avrami fractional kinetic model. The TPD curve is basically a horizontal straight-line segment between 343 K and 353 K, indicating that the adsorbent that was saturated with CO2 had almost no mass loss, which means that when T ≤ 353 K, no desorption process occurs. Then, the TPD curve is a downwardly sloping line segment between 353 K and 383 K, and mass loss begins to appear, indicating that CO2 is gradually desorbed from the surface of the adsorbent. Finally, when the temperature is maintained at 383 K, the TPD curve presents a vertical line segment, indicating that CO2 is quickly and completely desorbed from the surface of the adsorbent. Moreover, according to the desorption curve, the CO2 is almost completely desorbed within 6 min, which means that the desorption rate is very fast. All these results suggest that the optimal desorption temperature for AMMS-T is about 383 K.
From Figure 9, we also noticed that the R2 values for the fitted Avrami fractional-order model are 0.99725, which verifies that the Avrami model can describe the experimental data well.

3.4. Cyclic CO2 Adsorption/Desorption Behavior of the AMMS-T

The persistent cyclic adsorption/desorption behavior of the adsorbent is essential for long-term operation. Figure 10 describes the adsorption capacity of the AMMS-T in repeated cycles of CO2 adsorption at 343 K and desorption at 383 K. The cycle data indicate that the adsorption performance of the AMMS-T is quite stable, and the adsorption capacity is still about 3.05 mmol g−1 after five adsorption/regeneration cycles.

4. Conclusions

This study presents an effective amine-functionalized micro-mesoporous silicon adsorbent for CO2 capture. The adsorbent was prepared by the impregnation method using micro-mesoporous silicon as a support and TEPA as a modifier. The resulting sample was characterized with the XRD, SEM, TG, nitrogen adsorption and FTIR analysis. The characterization results show that the prepared AMMS-T maintains a micro-mesoporous structure after amine loading, and there are highly uniformly dispersed amine groups distributed on the surface of the pore which is beneficial to the adsorption of CO2.
The CO2 adsorption capacity of AMMS-T displays an upward trend with the increase in temperature in the range of 303–343 K, and the maximum adsorption capacity is 3.32 mmol-CO2 g−1-adsorbent. The results of the adsorption kinetic analysis illustrate that Avrami’s fractional kinetic model can fit the CO2 adsorption experimental data well. In addition, the simulation results reveal that the CO2 adsorption rate of AMMS-T is mainly affected by the intra-particle diffusion rate, and temperature has little effect on it. The optimal regeneration temperature of AMMS-T is about 383 K and the CO2 desorption process can be well simulated by the Avrami model.
In summary, the prepared AMMS-T shows excellent CO2 adsorption/desorption performance. The good agreement between the results of the Avrami model and the experimental data shows that the kinetic constants that were obtained by the simulation are highly effective for designing the actual CO2 adsorption process.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27113429/s1, Figure S1: SEM images of AMMS-E, AMMS-D, AMMS-PEI and AMMS-P; Figure S2: N2 adsorption isotherms of AMMS-E, AMMS-D, AMMS-P and AMMS-PEI; Figure S3: CO2 adsorption curves of MMS, AMMS-E, AMMS-D, AMMS-P and AMMS-PEI.

Author Contributions

Conceptualization, Y.L. and Y.C.; Methodology, Y.L.; Software, Y.L. and J.W.; Validation, Y.L.; formal analysis, M.L.; investigation, Y.C., M.L. and X.W.; data curation, J.W. and X.W.; writing—original draft preparation, Y.C.; writing—review and editing, Y.L.; visualization, Y.C.; supervision, M.L.; project administration, Y.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Fujian Province grant number 2019J01774 And The research start-up funds of the Fujian University of Technology grant number GY-Z20081.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author, [Liu Y.], upon reasonable request.

Conflicts of Interest

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

References

  1. NOAA (National Ocean and Atmospheric Administration); Earth System Research Laboratory (ESRL). Global Monitoring Division. Available online: https://gml.noaa.gov/ccgg/ (accessed on 1 May 2022).
  2. Bhatia, S.K.; Bhatia, R.K.; Jeon, J.M.; Kumar, G.; Yang, Y.H. Carbon Dioxide Capture and Bioenergy Production Using Biological System—A Review. Renew. Sustain. Energy Rev. 2019, 08, 143–158. [Google Scholar] [CrossRef]
  3. Herzog, H. Advanced Post-Combustion CO2 Capture. Work 2009, 1, 39. [Google Scholar]
  4. Mac Dowell, N.; Fennell, P.S.; Shah, N.; Maitland, G.C. The Role of CO2 Capture and Utilization in Mitigating Climate Change. Nat. Clim. Chang. 2017, 7, 243–249. [Google Scholar] [CrossRef] [Green Version]
  5. Song, C.; Liu, Q.; Ji, N.; Deng, S.; Zhao, J.; Li, Y.; Song, Y.; Li, H. Alternative Pathways for Efficient CO2 Capture by Hybrid Processes—A Review. Renew. Sustain. Energy Rev. 2018, 82, 215–231. [Google Scholar] [CrossRef]
  6. Hu, X.E.; Liu, L.; Luo, X.; Xiao, G.; Shiko, E.; Zhang, R.; Fan, X.; Zhou, Y.; Liu, Y.; Zeng, Z.; et al. A Review of N-Functionalized Solid Adsorbents for Post-Combustion CO2 Capture. Appl. Energy 2020, 260, 114244. [Google Scholar] [CrossRef]
  7. Jaffary, B.; Jaafari, L.; Idem, R. CO2 Capture Performance Comparisons of Polyamines at Practical Concentrations for Use as Activators for Methyldiethanolamine for Natural Gas Sweetening. Energy Fuels 2021, 35, 8081–8094. [Google Scholar] [CrossRef]
  8. Raganati, F.; Miccio, F.; Ammendola, P. Adsorption of Carbon Dioxide for Post-combustion Capture: A Review. Energy Fuels 2021, 35, 12845–12868. [Google Scholar] [CrossRef]
  9. Embaye, A.S.; Martínez-Izquierdo, L.; Malankowska, M.; Téllez, C.; Coronas, J. Poly(ether-block-amide) Copolymer Membranes in CO2 Separation Applications. Energy Fuels 2021, 35, 17085–17102. [Google Scholar] [CrossRef]
  10. Liu, M.; Hohenshil, A.; Gadikota, G. Integrated CO2 Capture and Removal via Carbon Mineralization with Inherent Regeneration of Aqueous Solvents. Energy Fuels 2021, 35, 8051–8068. [Google Scholar] [CrossRef]
  11. Hughes, R.; Kotamreddy, G.; Ostace, A.; Bhattacharyya, D.; Siegelman, R.L.; Parker, S.T.; Didas, S.A.; Long, J.R.; Omell, B.; Matuszewski, M. Isotherm, Kinetic, Process Modeling, and Techno-Economic Analysis of a Diamine-Appended Metal–Organic Framework for CO2 Capture Using Fixed Bed Contactors. Energy Fuels 2021, 35, 6040–6055. [Google Scholar] [CrossRef]
  12. Chen, Y.; Long, Y.; Sun, J.; Bai, S.; Chen, Y.; Chen, Z.; Zhao, C. Core-in-Shell, Cellulose-Templated CaO-Based Sorbent Pellets for CO2 Capture at Elevated Temperatures. Energy Fuels 2021, 35, 13215–13223. [Google Scholar] [CrossRef]
  13. Alghamdi, T.; Baamran, K.S.; Okoronkwo, M.U.; Rownaghi, A.A.; Rezaei, F. Metal-Doped K–Ca Double Salts with Improved Capture Performance and Stability for High-Temperature CO2 Adsorption. Energy Fuels 2021, 35, 4258–4266. [Google Scholar] [CrossRef]
  14. Balou, S.; Babak, S.E.; Priye, A. Synergistic Effect of Nitrogen Doping and Ultra-Microporosity on the Performance of Biomass and Microalgae-Derived Activated Carbons for CO2 Capture. ACS Appl. Mater. Interfaces 2020, 12, 42711–42722. [Google Scholar] [CrossRef] [PubMed]
  15. Cai, W.; Ding, J.; He, Y.; Chen, X.; Yuan, D.; Chen, C.; Cheng, L.; Du, W.; Wan, H.; Guan, G. Nitrogen-Doped Microporous Carbon Prepared by One-Step Carbonization: Rational Design of a Polymer Precursor for Efficient CO2 Capture. Energy Fuels 2021, 35, 8857–8867. [Google Scholar] [CrossRef]
  16. Li, J.; Zhang, W.; Bao, A. Design of Hierarchically Structured Porous Boron/Nitrogen-Codoped Carbon Materials with Excellent Performance for CO2 Capture. Ind. Eng. Chem. Res. 2021, 60, 2710–2718. [Google Scholar] [CrossRef]
  17. Luzzi, E.; Aprea, P.; de Luna, M.; Caputo, D.; Filippone, G. Mechanically Coherent Zeolite 13X/Chitosan Aerogel Beads for Effective CO2 Capture. ACS Appl. Mater. Interfaces 2021, 13, 20728–20734. [Google Scholar] [CrossRef]
  18. Rosu, C.; Pang, S.H.; Sujan, A.R.; Sakwa-Novak, M.A.; Ping, E.W.; Jones, C.W. Effect of Extended Aging and Oxidation on Linear Poly (Propylenimine)-Mesoporous Silica Composites for CO2 Capture from Simulated Air and Flue Gas Streams. ACS Appl. Mater. Interfaces 2020, 12, 38085–38097. [Google Scholar] [CrossRef]
  19. Gil, A.; Arrieta, E.; Vicente, M.A.; Korili, S.A. Synthesis and CO2 Adsorption Properties of Hydrotalcite-like Compounds Prepared from Aluminum Saline Slag Wastes. Chem. Eng. J. 2018, 334, 1341–1350. [Google Scholar] [CrossRef]
  20. Park, J.; Park, J.R.; Choe, J.H.; Kim, S.; Kang, M.; Kang, D.W.; Kim, J.Y.; Jeong, Y.W.; Hong, C.S. Metal-Organic Framework Adsorbent for Practical Capture of Trace Carbon Dioxide. ACS Appl. Mater. Interfaces 2020, 12, 50534–50540. [Google Scholar] [CrossRef]
  21. Zulys, A.; Yulia, F.; Muhadzib, N.; Nasruddin. Biological Metal–Organic Frameworks (Bio-MOFs) for CO2 Capture. Ind. Eng. Chem. Res. 2020, 60, 37–51. [Google Scholar] [CrossRef]
  22. Serna-Guerrero, R.; Sayari, A. Modeling Adsorption of CO2 on Amine-Functionalized Mesoporous Silica. 2: Kinetics and Breakthrough Curves. Chem. Eng. J. 2010, 161, 182–190. [Google Scholar] [CrossRef]
  23. Samanta, A.; Zhao, A.; Shimizu, G.K.H.; Sarkar, P.; Gupta, R. Post-Combustion CO2 Capture Using Solid Sorbents: A Review. Ind. Eng. Chem. Res. 2012, 51, 1438–1463. [Google Scholar] [CrossRef]
  24. Hefti, M.; Marx, D.; Joss, L.; Mazzotti, M. Adsorption Equilibrium of Binary Mixtures of Carbon Dioxide and Nitrogen on Zeolites ZSM-5 and 13X. Microporous Mesoporous Mater. 2015, 215, 215–228. [Google Scholar] [CrossRef]
  25. Cen, Q.; Fang, M.; Wang, T.; Majchrzak-Kucęba, I.; Wawrzyńczak, D.; Luo, Z. Thermodynamics and Regeneration Studies of CO2 Adsorption on Activated Carbon. Greenh. Gases Sci. Technol. 2016, 6, 787–796. [Google Scholar] [CrossRef]
  26. Son, W.-J.; Choi, J.-S.; Ahn, W.-S. Adsorptive Removal of Carbon Dioxide Using Polyethyleneimine-Loaded Mesoporous Silica Materials. Microporous Mesoporous Mater. 2008, 113, 31–40. [Google Scholar] [CrossRef]
  27. Kinefuchi, I.; Yamaguchi, H.; Sakiyama, Y.; Takagi, S.; Matsumoto, Y. Inhomogeneous Decomposition of Ultrathin Oxide Films on Si(100): Application of Avrami Kinetics to Thermal Desorption Spectra. J. Chem. Phys. 2008, 128, 164712. [Google Scholar] [CrossRef] [PubMed]
  28. de Menezes, E.W.; Lima, E.C.; Royer, B.; de Souza, F.E.; dos Santos, B.D.; Gregório, J.R.; Costa, T.M.H.; Gushikem, Y.; Benvenutti, E.V. Ionic Silica Based Hybrid Material Containing the Pyridinium Group Used as an Adsorbent for Textile Dye. J. Colloid Interface Sci. 2012, 378, 10–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G.H.; Chmelka, B.F.; Stucky, G.D. Triblock Copolymer Syntheses of Mesoporous Silica with Periodic 50 to 300 Angstrom Pores. Science 1998, 279, 548–552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Qu, F.; Zhu, G.; Lin, H.; Zhang, W.; Sun, J.; Li, S.; Qiu, S. A Controlled Release of Ibuprofen by Systematically Tailoring the Morphology of Mesoporous Silica Materials. J. Solid State Chem. 2006, 179, 2027–2035. [Google Scholar] [CrossRef]
  31. Van Grieken, R.; Sotelo, J.L.; Menéndez, J.M.; Melero, J.A. Anomalous Crystallization Mechanism in the Synthesis of Nanocrystalline ZSM-5. Microporous Mesoporous Mater. 2000, 39, 135–147. [Google Scholar] [CrossRef]
  32. Wang, X.; Ma, X.; Song, C.; Locke, D.R.; Siefert, S.; Winans, R.E.; Möllmer, J.; Lange, M.; Möller, A.; Gläser, R. Molecular Basket Sorbents Polyethylenimine–SBA-15 for CO2 Capture from Flue Gas: Characterization and Sorption Properties. Microporous Mesoporous Mater. 2013, 169, 103–111. [Google Scholar] [CrossRef]
  33. Murugesan, V.; Umamaheswari, V.; Palanichamy, M. Isopropylation of M-Cresol over Mesoporous Al-MCM-41 Molecular Sieves. J. Catal. 2002, 210, 367–374. [Google Scholar] [CrossRef]
  34. Zhang, D.; Duan, A.; Zhao, Z.; Xu, C. Synthesis, Characterization, and Catalytic Performance of NiMo Catalysts Supported on Hierarchically Porous Beta-KIT-6 Material in the Hydrodesulfurization of Dibenzothiophene. J. Catal. 2010, 274, 273–286. [Google Scholar] [CrossRef]
  35. Liu, Y.; Zhang, W.; Pinnavaia, T.J. Steam-Stable MSU-S Aluminosilicate Mesostructures Assembled from Zeolite ZSM-5 and Zeolite Beta Seeds. Angew. Chem. Int. Ed. 2001, 40, 1255–1258. [Google Scholar] [CrossRef]
  36. Liu, J.; Yu, L.; Zhao, Z.; Chen, Y.; Zhu, P.; Wang, C.; Luo, Y.; Xu, C.; Duan, A.; Jiang, G. Potassium-Modified Molybdenum-Containing SBA-15 Catalysts for Highly Efficient Production of Acetaldehyde and Ethylene by the Selective Oxidation of Ethane. J. Catal. 2012, 285, 134–144. [Google Scholar] [CrossRef]
  37. Knöfel, C.; Martin, C.; Hornebecq, V.; Llewellyn, P.L. Study of Carbon Dioxide Adsorption on Mesoporous Aminopropylsilane- Functionalized Silica and Titania Combining Microcalorimetry and in Situ Infrared Spectroscopy. J. Phys. Chem. C 2009, 113, 21726–21734. [Google Scholar] [CrossRef]
  38. Wang, J.; Chen, H.; Zhou, H.; Liu, X.; Qiao, W.; Long, D.; Ling, L. Carbon Dioxide Capture Using Polyethylenimine-Loaded Mesoporous Carbons. J. Environ. Sci. (China) 2013, 25, 124–132. [Google Scholar] [CrossRef]
  39. Sanz-Perez, E.S.; Arencibia, A.; Calleja, G.; Sanz, R. Tuning the textural properties of HMS mesoporous silica. Functionalization towards CO2 adsorption. Microporous Mesoporous Mater. 2018, 260, 235–244. [Google Scholar] [CrossRef]
  40. Zhang, H.; Goeppert, A.; Kar, S.; Prakash, G.K.S. Structural parameters to consider in selecting silica supports for polyethylenimine based CO2 solid adsorbents importance of pore size. J. CO2 Util. 2018, 26, 246–253. [Google Scholar] [CrossRef]
  41. Zhang, W.; Gao, E.; Li, Y.; Bernards, M.T.; He, Y.; Shi, Y. CO2 capture with polyamine-based protic ionic liquid functionalized mesoporous silica. J. CO2 Util. 2019, 34, 606–615. [Google Scholar] [CrossRef]
  42. Cheng, J.; Liu, N.; Hu, L.; Li, Y.; Wang, Y.; Zhou, J. Polyethyleneimine entwine thermally-treated Zn/Co zeolitic imidazolate frameworks to enhance CO2 adsorption. Chem. Eng. J. 2019, 364, 530–540. [Google Scholar] [CrossRef]
  43. Meng, Y.; Yan, Y.; Wu, X.; Sharmin, N.; Zhao, H.; Lester, E.; Wu, T.; Pang, C. Synthesis and functionalization of cauliflower-like mesoporous siliceous foam materials from oil shale waste for post-combustion carbon capture. J. CO2 Util. 2020, 40, 101–199. [Google Scholar] [CrossRef]
  44. Jeon, S.; Min, J.; Kim, S.H.; Lee, K.B. Introduction of cross-linking agents to enhance the tability of polyethyleneimine-impregnated CO2 adsorbents: Effect of different alkyl chain lengths. Chem. Eng. J. 2020, 398, 125531. [Google Scholar] [CrossRef]
  45. Park, J.M.; Jhung, S.H. CO2 adsorption at low pressure over polymers-loaded mesoporous metal organic framework PCN-777: Effect of basic site and porosity on adsorption. J. CO2 Util. 2020, 42, 101332. [Google Scholar] [CrossRef]
  46. Gaikwad, S.; Kim, Y.; Gaikwad, R.; Han, S. Enhanced CO2 capture capacity of amine-functionalized MOF-177 metal organic framework. J. Environ. Chem. Eng. 2021, 9, 105523. [Google Scholar] [CrossRef]
  47. Zhou, C.; Yu, S.; Ma, K.; Liang, B.; Tang, S.; Liu, C.; Yue, H. Amine-functionalized mesoporous monolithic adsorbents for post-combustion carbon dioxide capture. Chem. Eng. J. 2021, 413, 127675. [Google Scholar] [CrossRef]
  48. Yuh-Shan, H. Citation Review of Lagergren Kinetic Rate Equation on Adsorption Reactions. Scientometrics 2004, 59, 171–177. [Google Scholar] [CrossRef]
  49. Ho, Y.S. Review of Second-Order Models for Adsorption Systems. J. Hazard. Mater. 2006, 136, 681–689. [Google Scholar] [CrossRef] [Green Version]
  50. Thomas, W.J.; Crittenden, B. Adsorption Technology and Design; Butterworth-Heinemann: Oxford, UK, 1998. [Google Scholar] [CrossRef]
  51. Plazinski, W.; Rudzinski, W.; Plazinska, A. Theoretical Models of Sorption Kinetics Including a Surface Reaction Mechanism: A Review. Adv. Colloid Interface Sci. 2009, 152, 2–13. [Google Scholar] [CrossRef]
  52. Lopes, E.C.N.; Dos Anjos, F.S.C.; Vieira, E.F.S.; Cestari, A.R. An Alternative Avrami Equation to Evaluate Kinetic Parameters of the Interaction of Hg(II) with Thin Chitosan Membranes. J. Colloid Interface Sci. 2003, 263, 542–547. [Google Scholar] [CrossRef]
  53. Cestari, A.R.; Vieira, E.F.; Vieira, G.S.; Almeida, L.E. The Removal of Anionic Dyes from Aqueous Solutions in the Presence of Anionic Surfactant Using Aminopropylsilica—A Kinetic Study. J. Hazard. Mater. 2006, 138, 133–141. [Google Scholar] [CrossRef]
  54. Liu, Q.; Shi, J.; Zheng, S.; Tao, M.; He, Y.; Shi, Y. Kinetics Studies of CO2 Adsorption/desorption on Amine-functionalized Multiwalled Carbon Nanotubes. Ind. Eng. Chem. Res. 2014, 53, 11677–11683. [Google Scholar] [CrossRef]
  55. Avrami, M. Kinetics of Phase Change. I General Theory. J. Chem. Phys. 1939, 7, 1103–1112. [Google Scholar] [CrossRef]
  56. Num, G. Granulation, Phase Change, and Microstructure Kinetics of Phase Change. III. J. Chem. Phys. 1941, 9, 177–184. [Google Scholar] [CrossRef]
  57. Benedict, J.B.; Coppens, P. Kinetics of the Single-Crystal to Single-Crystal Two-Photon Photodimerization of Alpha-Trans-Cinnamic Acid to Alpha-Truxillic Acid. J. Phys. Chem. A 2009, 113, 3116–3120. [Google Scholar] [CrossRef] [PubMed]
  58. Paul, S.; Ghoshal, A.K.; Mandal, B. Absorption of Carbon Dioxide into Aqueous Solutions of 2-Piperidineethanol: Kinetics Analysis. Ind. Eng. Chem. Res. 2009, 48, 1414–1419. [Google Scholar] [CrossRef]
  59. Wang, H.M.; Meng-Hui, L.I. Kinetics of Absorption of Carbon Dioxide into Aqueous Solutions of 2-Amino-2-Methyl-l-Propanol+Diethanolamine. J. Chem. Eng. Jpn. 2004, 37, 267–278. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of MMS and AMMS.
Figure 1. XRD patterns of MMS and AMMS.
Molecules 27 03429 g001
Figure 2. SEM images of MMS and AMMS-T.
Figure 2. SEM images of MMS and AMMS-T.
Molecules 27 03429 g002
Figure 3. The nitrogen adsorption isotherms of MMS and AMMS-T.
Figure 3. The nitrogen adsorption isotherms of MMS and AMMS-T.
Molecules 27 03429 g003
Figure 4. TGA results of AMMS.
Figure 4. TGA results of AMMS.
Molecules 27 03429 g004
Figure 5. FTIR spectrums of MMS and AMMS.
Figure 5. FTIR spectrums of MMS and AMMS.
Molecules 27 03429 g005
Figure 6. CO2 Adsorption curves of MMS and AMMS-T at 303 K, 323 K, and 343 K.
Figure 6. CO2 Adsorption curves of MMS and AMMS-T at 303 K, 323 K, and 343 K.
Molecules 27 03429 g006
Figure 7. The results of the Avrami kinetics models and the experimental CO2 uptakes of AMMS-T.
Figure 7. The results of the Avrami kinetics models and the experimental CO2 uptakes of AMMS-T.
Molecules 27 03429 g007
Figure 8. Arrhenius plots for the kinetic constant ka obtained for the Avrami model.
Figure 8. Arrhenius plots for the kinetic constant ka obtained for the Avrami model.
Molecules 27 03429 g008
Figure 9. Results of TPD and desorption experiment at 383 K.
Figure 9. Results of TPD and desorption experiment at 383 K.
Molecules 27 03429 g009
Figure 10. Recycle adsorption/desorption runs of AMMS-T (adsorption/desorption at 343/383 K).
Figure 10. Recycle adsorption/desorption runs of AMMS-T (adsorption/desorption at 343/383 K).
Molecules 27 03429 g010
Table 1. Chemicals used in the experiment.
Table 1. Chemicals used in the experiment.
Chemical Namemol. wt.PuritiesCAS-No.Sources
Tetrapropylammonium hydroxide solution203.3625.0%4499-86-9Macklin Biochemical Co., Ltd. (Shanghai, China)
Aluminium isopropoxide204.25≥98.0%555-31-7BASF Biotechnology Co., Ltd. (Hangzhou, China)
Tetraethylorthosilicate208.3398.0%78-10-4Aladdin Biochemical Technology Co., Ltd. (Shanghai, China)
Polyethylene-polypropylene glycol~5800 9003-11-6Macklin Biochemical Co., Ltd. (Shanghai, China)
Hydrochloric acid36.4635.0%7647-01-0Fuzhou Yihua Chemical Co., Ltd. (Fuzhou, China)
n-Butanol74.12≥99.5%71-36-3Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China)
Ethylenediamine60.1≥99.0%107-15-3Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China)
Diethylenetriamine103.1799.0%111-40-0Aladdin Biochemical Technology Co., Ltd. (Shanghai, China)
Tetraethylenepentamine189.3098.0%112-57-2Aladdin Biochemical Technology Co., Ltd. (Shanghai, China)
Pentaethylenehexamine232.3898.0%4067-16-7Macklin Biochemical Co., Ltd. (Shanghai, China)
Ethylene imine polymer60099.0%9002-98-6Aladdin Biochemical Technology Co., Ltd. (Shanghai, China)
Ethanol46.07≥99.7%64-17-5Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China)
Carbon dioxide44.0≥99.999%124-38-9Fuzhou Yuanhua Chemical Co., Ltd. (Fuzhou, China)
Nitrogen28.099.999%7727-37-9Fuzhou Yuanhua Chemical Co., Ltd. (Fuzhou, China)
Table 2. Nitrogen adsorption/desorption characterization details for MMS and AMMS.
Table 2. Nitrogen adsorption/desorption characterization details for MMS and AMMS.
SampleSurface Area (m2·g−1)Total Pore Volume (cm3·g−1)Micropore VolumeMesopore VolumeAverage Pore Diameter (nm)
MMS4981.2610.1701.0913.5
AMMS-E4041.4450.0801.3653.5
AMMS-D3141.3510.1401.2111.2
AMMS-T1210.7700.0800.6901.6
AMMS-P880.5000.0500.4501.4
AMMS-PEI9.70.08400.0843.3
Table 3. The CO2 adsorption capacity of MMS and AMMS.
Table 3. The CO2 adsorption capacity of MMS and AMMS.
AdsorbentAmine-Modified Materialqa (mmol-CO2/g-Adsorbent)
303 K323 K343 K
MMS * 0.990.450.33
AMMS-E *EDA0.790.470.37
AMMS-DDETA0.700.670.51
AMMS-TTEPA2.232.903.32
AMMS-PPEHA1.411.952.65
AMMS-PEIPEI0.741.111.75
* C0 = 100 vol.% CO2.
Table 4. Comparison of CO2 adsorption capacity of AMMS-T (present study) with the literature.
Table 4. Comparison of CO2 adsorption capacity of AMMS-T (present study) with the literature.
SupportAmine TypeTemp.
K
CO2 Partial Pressure (bar)CO2 Adsorption
(mmol-CO2/g-ads)
Ref.
HMSPEI31812.40[39]
MS-3040 (Microspherical Silica)PEI3580.953.26[40]
SBA-15TEPA3330.152.15[41]
Zn/CoZIFPEI29811.82[42]
SFM-0.83-100-5.2PEI3480.152.48[43]
SilicaPEI3530.152.86[44]
Mesoporous PCN-777PEI2980.251.41[45]
MOFPEI29812.84[46]
MCM550 (Mesoporous Monolithic)PEI3480.121.89[47]
AMMS-TTEPA3430.103.32This work
Table 5. The approximate values of the model parameters obtained by Avrami’s model and the corresponding correlation coefficient R2.
Table 5. The approximate values of the model parameters obtained by Avrami’s model and the corresponding correlation coefficient R2.
T (K)C0 (vol.%)Avrami Modelqa (mmol/g)
Kaqe (mmol/g)nR2
303100.3092.092.0000.9962.235
303150.4531.4861.6810.9951.576
303200.5921.1560.5430.9741.133
303400.8580.8511.0240.9821.189
323100.3482.6851.8130.9882.900
323150.4701.5851.7310.9911.775
323200.5681.7580.6530.9891.744
323400.8462.2911.8800.9882.667
343100.2323.1601.9820.9993.322
343150.3262.0741.5490.9952.368
343200.5471.9471.7500.9962.322
343400.8222.8221.8820.9953.078
Table 6. Related parameters calculated from CO2 adsorption isotherms fitted to the Arrhenius Equation.
Table 6. Related parameters calculated from CO2 adsorption isotherms fitted to the Arrhenius Equation.
CO2 Concentration (vol.%)A (min−1)Ea (KJ/mol)R2
100.031696−5.818060.43709
150.03105−6.765340.63287
200.300659−1.669470.99993
400.597542−0.897880.94752
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chen, Y.; Wu, J.; Wang, X.; Liu, M.; Liu, Y. Synthesis, Characterization and Application of Amine-Functionalized Hierarchically Micro-Mesoporous Silicon Composites for CO2 Capture in Flue Gas. Molecules 2022, 27, 3429. https://doi.org/10.3390/molecules27113429

AMA Style

Chen Y, Wu J, Wang X, Liu M, Liu Y. Synthesis, Characterization and Application of Amine-Functionalized Hierarchically Micro-Mesoporous Silicon Composites for CO2 Capture in Flue Gas. Molecules. 2022; 27(11):3429. https://doi.org/10.3390/molecules27113429

Chicago/Turabian Style

Chen, Yilan, Junjie Wu, Xin Wang, Minyi Liu, and Yamin Liu. 2022. "Synthesis, Characterization and Application of Amine-Functionalized Hierarchically Micro-Mesoporous Silicon Composites for CO2 Capture in Flue Gas" Molecules 27, no. 11: 3429. https://doi.org/10.3390/molecules27113429

Article Metrics

Back to TopTop