Next Article in Journal
Influence of Excitation Parameters on Finishing Characteristics in Magnetorheological Finishing for 6063 Aluminum Alloy
Previous Article in Journal
Micro-Milling of Additively Manufactured Al-Si-Mg Aluminum Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Use of Periodic Mesoporous Organosilica–Benzene Adsorbent for CO2 Capture to Reduce the Greenhouse Effect

by
David Cantador-Fernandez
1,
Dolores Esquivel
2,3,
José Ramón Jiménez
4,*,† and
José María Fernández-Rodríguez
1,3,*,†
1
Departamento de Química Inorgánica e Ingeniería Química, Campus de Rabanales, Edificio Marie Curie, Universidad de Córdoba, 14071 Córdoba, Spain
2
Departamento de Química Orgánica, Universidad de Córdoba, 14001 Córdoba, Spain
3
Instituto para la Energía y el Medioambiente (IQUEMA), Universidad de Córdoba, 14071 Córdoba, Spain
4
Departamento de Ingeniería Rural, Escuela Politécnica Superior de Belmez, Universidad de Córdoba, Ed. Leonardo Da Vinci, Campus de Rabanales, Ctra. N-IV, km-396, 14001 Córdoba, Spain
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2024, 17(11), 2669; https://doi.org/10.3390/ma17112669
Submission received: 8 May 2024 / Revised: 24 May 2024 / Accepted: 30 May 2024 / Published: 1 June 2024

Abstract

:
The CO2 adsorption of a phenylene-bridged ordered mesoporous organosilica (PMO–benzene) was analyzed. The maximum capture capacity was 638.2 mg·g−1 (0 °C and 34 atm). Approximately 0.43 g would be enough to reduce the amount of atmospheric CO2 in 1 m3 to pre-industrial levels. The CO2 adsorption data were analyzed using several isotherm models, including Langmuir, Freundlich, Sips, Toth, Dubinin–Radushkevich, and Temkin models. This study confirmed the capability of this material for use in reversible CO2 capture with a minimal loss of capacity (around 1%) after 10 capture cycles. Various techniques were employed to characterize this material. The findings from this study can help mitigate the greenhouse effect caused by CO2.

Graphical Abstract

1. Introduction

Excessive increases in CO2 concentration over a short period of time lead to an amplification of the greenhouse effect and a sharp rise in the average temperature of the planet. Before the industrial revolution, CO2 levels were 280 ppm [1]. As a result of human activities, a CO2 level of 421 ppm was reached in 2022 [2,3,4]. To address this situation, many international organizations and governments have worked to reach common agreements. One of the most important agreements was the Paris Agreement (2016) [5]. The capture of CO2 [6,7] could allow its use as a revalorized by-product, such as fuels, chemicals, and building materials, adding an economic incentive for CO2 capture and the green economy, while also reducing the environmental footprint [8,9]. Different adsorbents such as the diamine-based hybrid-slurry system [10], porous carbons [11,12], silica [13,14,15], MOFs [16,17,18], metal oxides [19,20,21,22], hydrotalcites [23,24], and periodic mesoporous organosilicas (PMOs) have been studied recently for CO2 capture.
PMO materials have a considerable interest for multiple applications [25] such as catalysis [26,27,28,29,30,31], metal adsorbents [32,33], organic pollutant adsorbents [34,35,36], gas adsorbents [37,38,39], chemosensors [40,41], optical devices [42], and the immobilization of enzymes or therapy in cancer cells [43,44].
Inagaki et al. [45], Melde et al. [46], and Asefa et al. [47] published the synthesis of a new homogeneously distributed mesoporous organic–inorganic material using bis(trimethoxysilyl)ethane, bis(triethoxysilyl)ethane, and bis(triethoxysilyl)ethene as precursors, respectively. This was the first synthesized PMO material. Subsequently, in 2002, Kuroki et al. [48] synthesized a new PMO material containing a benzene functional group for the first time using 1,3,5-Tris(triethoxysilyl)benzene and 1,3-Bis(triethoxysilyl)benzene as precursors and Cetylpyridinium chloride (CPCl) or cetyltrimethylammonium bromide (CTABr) as surfactants.
Selecting specific organosilica types and synthesis conditions enables precise control over the size, shape, uniformity, and periodicity of the pore structures [49]. However, the shape of the adsorbent particle determines the specific surface area.
Several authors have addressed CO2 adsorption using PMO materials [50,51,52,53,54,55,56,57] at low pressure and temperatures of 0 °C and 25 °C, with different results.
In this study, a PMO with phenylene bridges was chosen as the CO2 adsorbent material (Figure 1). Various techniques were utilized to characterize this material, and its CO2 capture capacity at low temperatures (0 °C, 10 °C, 20 °C, and 35 °C) and high pressure (35 atm) was evaluated. The CO2 adsorption data were fitted by the Langmuir, Freundlich, Sips, Toth, Dubinin–Radushkevich, and Temkin models.
The capability of this material for use in reversible CO2 capture with a minimal loss of capacity after 10 capture cycles was addressed.
The obtained results can aid in the reduction in the CO2 greenhouse effect.

2. Materials and Methods

2.1. PMO–Benzene

Periodic mesoporous organosilica (PMO–benzene) was synthesized via acid-catalyzed hydrolysis and the condensation of bis(triethoxysilyl)benzene following the method outlined by Burleigh [58,59]. In this process, 6 g of the surfactant BrijS10-Brij76 (Polyoxyethylene (10) stearyl ether) was dissolved in a solution of 19.6 mL of HCl and 279 mL of H2O while being stirred at 50 °C for 24 h. Subsequently, 17.2 mL of 1,4-bis(triethoxysilyl)benzene was added, and the mixture was continuously stirred at 50 °C for another 24 h. The solution was then kept at 90 °C under static conditions for an additional 24 h. The resulting precipitate was collected through vacuum filtration and washed with H2O. To remove the surfactant, the synthesized material was stirred in a HCl solution (1:50 ratio) at 80 °C for 12 h. Finally, the product was filtered, washed with ethanol, and dried under vacuum for 10 h. The molar ratios of Brij76, H2O, HCl, and organosilane used in the reaction were 0.11:222:3.20:0.56.

2.2. Material Characterisation

Different techniques were used for the characterization of the PMO–benzene samples. XRD analysis was conducted using a Bruker D8 Discover A25 instrument (Karlsruhe, Germany), utilizing the ICDD 2003 database [60]. N2 adsorption isotherms were used to evaluate SBET and porosity, conducted with the Autosorb iQ2 from Quantachrome Instruments (Boynton Beach, FL, USA). Prior to this, the samples underwent degassing for 24 h at 70 °C. The data were processed with AsiQwin software (version 3.0). The surface area was calculated using the Brunauer–Emmett–Teller (BET) method. Density functional theory (DFT) was used to test the porosity. The pore volumes were calculated on the basis of the single-point method. A Mastersizer S laser diffractometer (Malvern, UK) was used to determine the particle size. Particle aggregates were separated during an ultrasonic bath (10 min). Particle structure was analyzed using a transmission electron microscope (Talos F200i TEM; Thermo Fisher Scientific) (Waltham, MA, USA). A Sievert PCTPro-2000 instrument (Setaram) (Caluire-et-Cuire, Francia) was used to obtain the CO2 isotherms. The gases CO2 4.5 (99.995%) and He 5.0 (99.999%) were used.

2.3. Adsorption Isotherms

The CO2 isotherms for the PMO–benzene, measured at different temperatures (0 °C, 10 °C, 20 °C, and 35 °C), were fitted to several models [23,61]: Langmuir [62,63], Freundlich [64], Sips [65], Toth [66], Dubinin–Radushkevich (D-R) [67,68], and Temkin [69,70]. A detailed characterization was described by Cantador et al. [23]. A brief description is provided below.
Langmuir characterizes monolayer adsorption on a uniform surface [71] with identical adsorption sites, indicating that the energy remains constant regardless of the amount adsorbed [72]. Since adsorption takes place in a single layer, the surface area can be determined. In 1995, Tóth proposed [66] a correction factor χL to enhance the accuracy of monolayer adsorption calculations.
Freundlich, on the other hand, describes adsorption on heterogeneous surfaces [73,74], enabling the assessment of the surface’s heterogeneity and the adsorption intensity.
The Sips model integrates aspects of both the Langmuir and Freundlich models by depicting Freundlich-type adsorption at low pressures and resembling the Langmuir model at high pressures, thereby unifying the two approaches [75]. Analogously to the Langmuir model, a correction factor, χS, was applied.
Toth’s model modifies Langmuir’s equation to accurately describe adsorption isotherms that include both monolayer and multilayer formations up to the maximum relative pressure.
The D-R model is employed to describe the CO2 adsorption based on potential energy during pore filling, which in turn determines adsorption capacity.
The Temkin model explains the heat generated in the pore filling process on the solid surface, establishing a correlation with the quantity of gas adsorbed.
MATLAB software version R2015a was used for calculation and fitting.

3. Results and Discussion

3.1. Characterization

High-quality PMO–benzene was characterized. As illustrated in Figure 2, the XRD pattern of the sample displays its most prominent peak (100) at d = 53.4 Å. Additionally, two smaller and broader peaks were observed at approximately d = 27.3 Å (110) and d = 21 Å (200). These peaks confirm that the sample is a mesoporous material with a 2D hexagonal (P6mm) structure, characteristic of the organosilica PMO–benzene, consistent with the literature [36,76].
Figure 3 presents the particle size distribution, revealing a bimodal pattern. The primary peak is narrow, ranging from 2.5 to 35 μm and centered at 13.2 μm, suggesting uniformity in particle size within the sample. A secondary peak appears between 0.3 and 2.5 μm, with a center at 1.45 μm, indicating the presence of a smaller particle group.
Figure 4 displays the N2 adsorption–desorption isotherm, which is Type IV according to the IUPAC. The first portion of the isotherm corresponds to monolayer–multilayer adsorption, with a well-defined adsorption limit at pressures approaching p·p0−1 = 1 [77,78]. The hysteresis in the multilayer region is indicative of capillary condensation within the mesopores. This hysteresis is classified as type H1, which is associated with a pore system that features a relatively regular arrangement and a narrow distribution of uniform mesopores [78,79], consistent with the XRD-derived crystal structure. Most pores were small mesopores, ranging from 2.4 to 4.1 nm, with a predominant pore diameter of 3.47 nm, and the total pore volume was 0.68 cm3·g−1 (Table 1). The BET method revealed a high surface area of 928 m2·g−1, which explains the material’s significant adsorption capacity. Additionally, several micropores were identified, contributing 139 m2·g−1 to the total surface area, representing 15% of the total.
A TEM was used on the PMO–benzene samples. Figure 5A shows the ordered pore structure of the material from two perspectives: longitudinal and transverse to the pore network. Figure 5B shows channels with a porous diameter around 2.867 nm, i.e., a pore size which agrees with the pore-size distribution (2.4–4.1 nm). The value of 1.01 nm corresponds to the thickness of the channel wall.

3.2. CO2 Adsorption

Figure 6 illustrates the CO2 adsorption capacity of PMO–benzene at various temperatures (0 °C, 10 °C, 20 °C, and 35 °C) and pressures ranging from 0 to 35 atm. The isotherm at 0 °C displayed a significant resemblance to the N2 isotherms, suggesting similar adsorption behavior. Monolayer formation was observed around 13.3 atm (p·p0−1 = 0.38), calculated using MATLAB R2015a, with multilayer formation extending to p·p0−1 = 1. For the samples at 10 °C and 20 °C, multilayer formation was noted at 20.6 atm (p·p0−1 = 0.46) and 30.7 atm (p·p0−1 = 0.54), respectively. At 35 °C, saturation pressure was not achieved, and monolayer filling was incomplete at 35 atm.
The CO2 adsorption capacity decreased as the temperature increased. The highest adsorption capacity was observed at 0 °C (638.2 mg·g−1), consistent with previous characterization results.
In order to analyze the decrease in adsorption and to explore the potential for reusing the material as a CO2 adsorbent, a study was conducted involving 10 consecutive adsorption–desorption cycles at 0 °C, 10 °C, and 35 °C (Figure 7). The linear trend and maximum adsorption for each cycle are shown in Figure 8. After 10 cycles, there was a small loss in adsorption in all cases, with loss amounts between 1% and 1.5%. The working temperature did not seem to affect the adsorption capacity loss. The results indicate the material’s potential for reuse without significant loss of performance.
Various models were used to fit the adsorption data.
All the isotherms fit the Freundlich model well (Figure 9). For the Langmuir model, only the isotherms at 0 and 10 °C obtained an R2 value below 0.99 due to multilayer formation and significant adsorption in the final stage (Table 2). The samples fit the Freundlich model slightly better than the Langmuir model. The Freundlich model provided a better fit, indicating surface heterogeneity (n), with results suggesting a somewhat heterogeneous surface.
According to Giles et al.’s [80] classification, the isotherms exhibited characteristics between L-type, indicative of strong intermolecular attractions and a gradual reduction in available adsorption sites, and C-type, where the number of adsorption sites remains constant. L-type curves typically align with Langmuir adsorption behavior, while C-type curves suggest linear adsorption. This shape is associated with weak intermolecular forces and a gradual occupation of adsorption sites. The monolayer filling values (qm) were generally overestimated, but using the Tóth correction yielded accurate monolayer filling predictions (qmc). This adjustment enabled the calculation of specific surface area using the Langmuir model. At 0 °C, where the adsorption curve reached a p·p0−1 of 1, the specific surface area (SL) was 904.1 m2·g−1, closely matching the BET method’s value of 928 m2·g−1 (Table 1). The Freundlich intensity factor (nf), which measures the interaction strength between the adsorbent and adsorbate at low pressures, was above 1 but close to it, indicating favorable, albeit weak, interactions [66]. This result agrees with Giles et al.’s [80] interpretation. The Langmuir equilibrium constant (KL) and Freundlich nf values (Table 2) increased with temperature, except at 35 °C, consistent with findings from other CO2 capture materials in the literature [23,24,61]. The Freundlich adsorption capacity per unit concentration (Kf) increased as the temperature decreased, directly correlating with higher CO2 adsorption capacity.
These models effectively showed the adsorption trends (qmc and Kf), which rose as the temperature decreased. The Freundlich model provided a more accurate prediction of the maximum adsorption capacity (Figure 9) compared to the Langmuir model, which only accounts for monolayer adsorption.
The Sips and Toth models (three-parameter) improved the fit over the Langmuir and Freundlich models for samples between 0 and 20 °C, reflecting in higher R2 values (Table 3). These models used relative pressures (p·p0−1) of 0.38, 0.46, and 0.54 as starting points for multilayer formation for samples at 0 °C, 10 °C, and 20 °C, respectively, obtained from the adsorption curve’s derivative. For the 35 °C sample, where no multilayer formation occurred, the maximum p·p0−1 value was used for fitting (Figure 9), though the parameter values are not listed in Table 3. The temperature-dependent trends for qs (Sips) and qT (Toth) differed from qmc (Langmuir) due to the models’ varying interpretations of the adsorption curve. The Langmuir model treated the curve as a monolayer over the entire pressure range, whereas the Sips and Toth models incorporated the pre-parameter for multilayer initiation, which varied with temperature. The Sips and Toth models predicted the CO2 adsorption capacity equally well, as indicated by the R2 values. The increasing qs and qt values with temperature are related to the increase in the relative pressure (p·p0−1) at which the multilayer is formed. This could be related to the increase in curvature at low p·p0−1 values and is in accordance with the nf value of the Freundlich model (the higher the initial curvature, the higher the nf). The heterogeneity factors from the Sips (nS) and Toth (nT) models agreed with the Freundlich model’s heterogeneity factor (n).
This research emphasizes the importance of using three-parameter models alongside classical two-parameter models to achieve better isotherm fitting, particularly when multilayer formation occurs (Figure 9).
For the D-R model, all isotherms had R2 values greater than 0.9. These adsorption capacities (qD) were slightly higher than those determined by the Langmuir model (qmc). The adsorption constant (β) was used to determine the free energy (E), which was smaller than 8 kJ⋅mol−1 for all samples (Table 4), suggesting the predominantly physical nature of adsorption. A slight increase in E was observed with increasing temperature.
A relationship between the D-R adsorption (E) and Temkin maximum binding energy (KTk) was observed, with both parameters increasing upon increasing temperature. The same trend was observed for the Temkin constant (bTk). These findings align with the expectation of low adsorption energy [81]. The physical nature of adsorption and the weak intermolecular attraction between adsorbent and adsorbate, as indicated by the fits, support the reversibility of CO2 capture. This is further evidenced by the minimal loss of adsorption after 10 cycles (Figure 7 and Figure 8). These properties suggest that the material is suitable for reversible multicycle CO2 capture by changing pressure conditions.
Table 5 provides a summary of significant studies on CO2 capture using mesoporous materials. Most of these studies focus on single temperature and pressure conditions and do not include adsorption–desorption cycling analyses. This research, however, examines the adsorption performance of PMO–benzene across various operating conditions.
Sim et al. [50] and Sim et al. [51] studied PMO–benzene samples functionalized with N-[3-(trimethoxysilyl)propyl] ethylene-diamine and PEO, respectively. The best result amounted to 133.32 mg·g−1 and was obtained for PMO–benzene (25 °C and 1 atm) modified with N-[3-(trimethoxysilyl)propyl] ethylene-diamine [50].
Other studies on PMOs at standard pressure and 0 °C have been conducted by Kirren et al. [52] and Wei et al. [53] (modified organosilica PMO–ethane), Liu et al. [54] (periodic mesoporous organosilica with a basic urea-derived framework), Rekha et al. [55] (PMO–cyclophosphazene), and Xu et al. [56] (polyethylenimine75-MCM-41) at standard pressure and 75 °C. In contrast, Lourenço et al. [57] worked at 10 atm. The maximum adsorption was 133 mg·g−1 for polyethylenimine75-MCM-41 at 75 °C and 1 atm [56].
The findings of this study indicate that CO2 adsorption at 34 atm pressure and temperatures ranging from 0 °C to 35 °C surpasses previously reported values for PMOs in the literature.
Table 5 compares these results with those for other mesoporous materials, such as those studied by Chowdhury et al. [20], Bhagiyalakshmi et al. [82], Wang et al. [83], Niu et al. [84], Liu et al. [21] and Li et al. [85]. The best result was 391.6 mg·g−1 (25 °C and 10 atm). In the current study, better results were obtained when working between temperatures of 0 °C and 10 °C and at a pressure of 34 atm.
In addition, the results from this study have been evaluated against various MOF-type materials as detailed in Table 5, including those reported by Sheng-Han et al. [86], Xiaoliang et al. [87], Bourrelly et al. [88], Zhang et al. [89], Zhou et al. [90], Zhao et al. [91], Millward et al. [92], and Furukawa et al. [93]. Notably, only three studies—Zhang et al. [89] (1007.6 mg·g−1), Millward et al. [92] (1493 mg·g−1), and Furukawa et al. [93] (2400 mg·g−1)—reported higher CO2 capture capacities at elevated pressures and 25 °C than the highest value from our study (827.8 mg·g−1 at 0 °C and 34 atm). However, these studies did not investigate adsorption–desorption cycles extensively (with Furukawa [93] conducting only a two-cycle study), which could impact the material’s structure and reduce its adsorption capacity. In contrast, our study involved multiple adsorption–desorption cycles, revealing a minor capacity loss of 1% to 1.5% after 10 cycles, underscoring its potential industrial application for purifying CO2-rich environments.
Moreover, the adsorption capacity was compared with adsorption capacities in previous studies. The obtained results for PMO–benzene at 34 atm and 0 °C (638.2 mg·g−1) was significantly higher than those of hydrotalcite Mg-Al (142.02 mg·g−1) [24] and organohydrotalcite (176.66 mg·g−1) [23], but was lower than the best result obtained with PMO–Ethane [61], which was 827.8 mg·g−1 at 0 °C and 34 atm.
Lastly, it is noteworthy to quantify the amount of adsorbent required to reduce current atmospheric CO2 levels (421 ppm (mL·m−3) [4]) to pre-industrial levels (280 ppm (mL·m−3) [1]). Given CO2’s density at 0 °C and 1 atm (1.976 mg·cm−3), the required amount would be 276.85 mg·m−3. Therefore, with a maximum capture capacity of 638.2 mg·g−1 at 0 °C, 0.43 g of PMO–benzene would suffice to lower the CO2 concentration in 1 m3 of air to pre-industrial levels. To apply this to a large volume, such as Wembley Stadium (1,139,100 m3), approximately 489.8 kg of PMO–benzene would be needed to achieve this reduction.

4. Conclusions

In this study, the adsorbent PMO–benzene was tested at high CO2 gas pressures (up to 35 atm) and low temperatures (0 °C, 10 °C, 20 °C, and 35 °C).
  • The pore size was between 2.4 and 4.1 nm, while the total pore volume was 0.68 cm3·g−1 and SBET was 928 m2·g−1.
  • All isotherms fitted the Freundlich model well. nf > 1 was very close to 1, indicating favorable and weak interactions.
  • The Sips and Toth models improved the results obtained by other equations at temperatures between 0 °C and 20 °C, where multilayer formation occurred.
  • The D-R and Temkin models showed a physical nature of adsorption (E < 8 kJ·mol−1).
  • PMO–benzene featured the maximum adsorption (638.2 mg·g−1) at 0 °C and 34 atm.
  • These results highlight that 0.43 g of PMO–benzene would be enough to reduce the CO2 level in 1 m3 of air to pre-industrial levels; 489.8 kg of PMO–benzene would be required to reduce the CO2 concentration of the volume of Wembley soccer stadium to pre-industrial levels.
  • The maximum loss of adsorption capacity was 1.45% for the sample at 0 °C, after 10 adsorption–desorption cycles. Consequently, this material could be used in capture processes using changes in the pressure conditions.
  • PMO–benzene could contribute to the development of CO2 capture and use (CCU) technology.

Author Contributions

Conceptualization, J.R.J. and J.M.F.-R.; formal analysis, D.C.-F. and D.E.; data curation, D.C.-F. and D.E.; writing—original draft preparation, D.C.-F.; writing—review and editing, J.R.J. and J.M.F.-R.; funding acquisition, J.R.J. and J.M.F.-R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Government through the PRECAST research project (Ref. PID2019-111029RB-I00).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the authors on request.

Acknowledgments

The authors wish to thank the IE57164 project (FEDER 2011) supported by the Andalusian regional government.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Bui, M.; Adjiman, C.S.; Bardow, A.; Anthony, E.J.; Boston, A.; Brown, S.; Fennell, P.S.; Fuss, S.; Galindo, A.; Hackett, L.A.; et al. Carbon Capture and Storage (CCS): The Way Forward. Energy Environ. Sci. 2018, 11, 1062–1176. [Google Scholar] [CrossRef]
  2. U.S. Department of Commerce, National Oceanic and Atmospheric Administration. Available online: https://www.noaa.gov/news-release/carbon-dioxide-now-more-than-50-higher-than-pre-industrial-levels#:~:text=PriortotheIndustrialRevolutionatmosphereforthousandsofyears (accessed on 24 October 2022).
  3. Friedlingstein, P.; Jones, M.W.; O’Sullivan, M.; Andrew, R.M.; Bakker, D.C.E.; Hauck, J.; Le Quéré, C.; Peters, G.P.; Peters, W.; Pongratz, J.; et al. Global Carbon Budget 2021. Earth Syst. Sci. Data 2022, 14, 1917–2005. [Google Scholar] [CrossRef]
  4. Scripps Institution of Oceanography, UC San Diego. The Keeling Curve. Available online: https://keelingcurve.ucsd.edu (accessed on 24 October 2022).
  5. United Nations Paris Agreement. Available online: https://unfccc.int/sites/default/files/english_paris_agreement.pdf (accessed on 12 November 2020).
  6. IPCC. Global Warming of 1.5 °C. An IPCC Special Report on the Impacts of Global Warming of 1.5 °C above Pre-Industrial Levels and Related Global Greenhouse Gas Emission Pathways, in the Context of Strengthening the Global Response to the Threat of Climate Change; Masson-Delmotte, V., Zhai, P., Pörtner, H.-O., Roberts, D.C., Skea, J., Shukla, P.R., Pirani, A., Moufouma-Okia, W., Péan, C., Pidcock, R., et al., Eds.; IPCC: Geneva, Switzerland, 2018; ISBN 978-92-9169-151-7. [Google Scholar]
  7. Zhang, W.; Yang, Y.; Li, Y.; Li, F.; Luo, M. Recent Progress on Integrated CO2 Capture and Electrochemical Upgrading. Mater. Today Catal. 2023, 2, 100006. [Google Scholar] [CrossRef]
  8. Vega, L.F.; Bahamon, D.; Alkhatib, I.I.I. Perspectives on Advancing Sustainable CO2 Conversion Processes: Trinomial Technology, Environment, and Economy. ACS Sustain. Chem. Eng. 2024, 12, 5357–5382. [Google Scholar] [CrossRef]
  9. Baena-Moreno, F.M.; Rodríguez-Galán, M.; Vega, F.; Alonso-Fariñas, B.; Vilches Arenas, L.F.; Navarrete, B. Carbon Capture and Utilization Technologies: A Literature Review and Recent Advances. Energy Sources Part A Recover. Util. Environ. Eff. 2019, 41, 1403–1433. [Google Scholar] [CrossRef]
  10. Salih, H.A.; Alkhatib, I.I.I.; Zahra, M.A.; Vega, L.F. Diamine Based Hybrid-Slurry System for Carbon Capture. J. CO2 Util. 2023, 68, 102383. [Google Scholar] [CrossRef]
  11. Cox, M.; Mokaya, R. Ultra-High Surface Area Mesoporous Carbons for Colossal Pre Combustion CO2 Capture and Storage as Materials for Hydrogen Purification. Sustain. Energy Fuels 2017, 1, 1414–1424. [Google Scholar] [CrossRef]
  12. Zhu, B.; Huang, J.; Lu, J.; Zhao, D.; Lu, L.; Jin, S.; Zhou, Q. Worm-like Hierarchical Porous Carbon Derived from Bio-Renewable Lignin with High CO2 Capture Capacity. Int. J. Electrochem. Sci. 2017, 12, 11102–11107. [Google Scholar] [CrossRef]
  13. Chen, C.; Feng, N.; Guo, Q.; Li, Z.; Li, X.; Ding, J.; Wang, L.; Wan, H.; Guan, G. Template-Directed Fabrication of MIL-101(Cr)/Mesoporous Silica Composite: Layer-Packed Structure and Enhanced Performance for CO2 Capture. J. Colloid Interface Sci. 2018, 513, 891–902. [Google Scholar] [CrossRef]
  14. Kishor, R.; Ghoshal, A.K. APTES Grafted Ordered Mesoporous Silica KIT-6 for CO2 Adsorption. Chem. Eng. J. 2015, 262, 882–890. [Google Scholar] [CrossRef]
  15. Licciulli, A.; Notaro, M.; De Santis, S.; Terreni, C.; Kunjalukkal Padmanabhan, S. CO2 Capture on Amine Impregnated Mesoporous Alumina-Silica Mixed Oxide Spheres. Fuel Process. Technol. 2017, 166, 202–208. [Google Scholar] [CrossRef]
  16. Wang, X.; Li, H.; Hou, X.J. Amine-Functionalized Metal Organic Framework as a Highly Selective Adsorbent for CO2 over CO. J. Phys. Chem. C 2012, 116, 19814–19821. [Google Scholar] [CrossRef]
  17. Ding, M.; Flaig, R.W.; Jiang, H.-L.; Yaghi, O.M. Carbon Capture and Conversion Using Metal–Organic Frameworks and MOF-Based Materials. Chem. Soc. Rev. 2019, 48, 2783–2828. [Google Scholar] [CrossRef] [PubMed]
  18. Nandi, S.; De Luna, P.; Daff, T.D.; Rother, J.; Liu, M.; Buchanan, W.; Hawari, A.I.; Woo, T.K.; Vaidhyanathan, R. A Single-Ligand Ultra-Microporous MOF for Precombustion CO2 Capture and Hydrogen Purification. Sci. Adv. 2015, 1, e1500421. [Google Scholar] [CrossRef] [PubMed]
  19. Ho, K.; Jin, S.; Zhong, M.; Vu, A.T.; Lee, C.H. Sorption Capacity and Stability of Mesoporous Magnesium Oxide in Post-Combustion CO2 Capture. Mater. Chem. Phys. 2017, 198, 154–161. [Google Scholar] [CrossRef]
  20. Chowdhury, S.; Parshetti, G.K.; Balasubramanian, R. Post-Combustion CO2 Capture Using Mesoporous TiO2/Graphene Oxide Nanocomposites. Chem. Eng. J. 2015, 263, 374–384. [Google Scholar] [CrossRef]
  21. Liu, G.; Tatsuda, K.; Yoneyama, Y.; Tsubaki, N. Synthesis of Mesoporous Cerium Compound for CO2 Capture. E3S Web Conf. 2017, 22, 00106. [Google Scholar] [CrossRef]
  22. Quesada Carballo, L.; Perez Perez, M.d.R.; Cantador Fernández, D.; Caballero Amores, A.; Fernández Rodríguez, J.M. Optimum Particle Size of Treated Calcites for CO2 Capture in a Power Plant. Materials 2019, 12, 1284. [Google Scholar] [CrossRef]
  23. Cantador Fernandez, D.; Suescum Morales, D.; Jiménez, J.R.; Fernández-Rodriguez, J.M. CO2 Adsorption by Organohydrotalcites at Low Temperatures and High Pressure. Chem. Eng. J. 2022, 431, 134324. [Google Scholar] [CrossRef]
  24. Suescum-Morales, D.; Cantador-Fernández, D.; Jiménez, J.R.; Fernández, J.M. Mitigation of CO2 Emissions by Hydrotalcites of Mg3Al-CO3 at 0 °C and High Pressure. Appl. Clay Sci. 2021, 202, 105950. [Google Scholar] [CrossRef]
  25. Karimi, B.; Ganji, N.; Pourshiani, O.; Thiel, W.R. Periodic Mesoporous Organosilicas (PMOs): From Synthesis Strategies to Applications. Prog. Mater. Sci. 2022, 125, 100896. [Google Scholar] [CrossRef]
  26. Yang, Q.; Liu, J.; Zhang, L.; Li, C. Functionalized Periodic Mesoporous Organosilicas for Catalysis. J. Mater. Chem. 2009, 19, 1945–1955. [Google Scholar] [CrossRef]
  27. Esquivel, D.; De Canck, E.; Jimenez-Sanchidrian, C.; Van Der Voort, P.; Romero-Salguero, F.J. Periodic Mesoporous Organosilicas as Catalysts for Organic Reactions. Curr. Org. Chem. 2014, 18, 1280–1295. [Google Scholar] [CrossRef]
  28. Huang, Y.; Yuan, P.; Wu, Z.; Yuan, X. Preparation of Surface-Silylated and Benzene-Bridged Ti-Containing Mesoporous Silica for Cyclohexene Epoxidation. J. Porous Mater. 2016, 23, 895–903. [Google Scholar] [CrossRef]
  29. Horiuchi, Y.; Do Van, D.; Yonezawa, Y.; Saito, M.; Dohshi, S.; Kim, T.-H.; Matsuoka, M. Synthesis and Bifunctional Catalysis of Metal Nanoparticle-Loaded Periodic Mesoporous Organosilicas Modified with Amino Groups. RSC Adv. 2015, 5, 72653–72658. [Google Scholar] [CrossRef]
  30. López, M.; Esquivel, D.; Jiménez-Sanchidrián, C.; Romero-Salguero, F.; Van Der Voort, P. A “one-step” Sulfonic Acid PMO as a Recyclable Acid Catalyst. J. Catal. 2015, 326, 139–148. [Google Scholar] [CrossRef]
  31. Jiang, Y.; Liu, X.; Chen, Y.; Zhou, L.; He, Y.; Ma, L.; Gao, J. Pickering Emulsion Stabilized by Lipase-Containing Periodic Mesoporous Organosilica Particles: A Robust Biocatalyst System for Biodiesel Production. Bioresour. Technol. 2014, 153, 278–283. [Google Scholar] [CrossRef] [PubMed]
  32. Imamoglu, M.; Pérez-Quintanilla, D.; Sierra, I. Bifunctional Periodic Mesoporous Organosilicas with Sulfide Bridges as Effective Sorbents for Hg(II) Extraction from Environmental and Drinking Waters. Microporous Mesoporous Mater. 2016, 229, 90–97. [Google Scholar] [CrossRef]
  33. Esquivel, D.; Ouwehand, J.; Meledina, M.; Turner, S.; Van Tendeloo, G.; Romero-Salguero, F.J.; Clercq, J.D.; Voort, P. Van Der Thiol-Ethylene Bridged PMO: A High Capacity Regenerable Mercury Adsorbent via Intrapore Mercury Thiolate Crystal Formation. J. Hazard. Mater. 2017, 339, 368–377. [Google Scholar] [CrossRef]
  34. Ganiyu, S.O.; Bispo, C.; Bion, N.; Ferreira, P.; Batonneau-Gener, I. Periodic Mesoporous Organosilicas as Adsorbents for the Organic Pollutants Removal in Aqueous Phase. Microporous Mesoporous Mater. 2014, 200, 117–123. [Google Scholar] [CrossRef]
  35. Deka, J.R.; Liu, C.L.; Wang, T.H.; Chang, W.C.; Kao, H.M. Synthesis of Highly Phosphonic Acid Functionalized Benzene-Bridged Periodic Mesoporous Organosilicas for Use as Efficient Dye Adsorbents. J. Hazard. Mater. 2014, 278, 539–550. [Google Scholar] [CrossRef] [PubMed]
  36. Otero, R.; Esquivel, D.; Ulibarri, M.A.; Jiménez-Sanchidrián, C.; Romero-Salguero, F.J.; Fernández, J.M. Adsorption of the Herbicide S-Metolachlor on Periodic Mesoporous Organosilicas. Chem. Eng. J. 2013, 228, 205–213. [Google Scholar] [CrossRef]
  37. Lourenço, M.A.O.; Silva, R.M.; Silva, R.F.; Pinna, N.; Pronier, S.; Pires, J.; Gomes, J.R.B.; Pinto, M.L.; Ferreira, P. Turning Periodic Mesoporous Organosilicas Selective to CO2/CH4 Separation: Deposition of Aluminium Oxide by Atomic Layer Deposition. J. Mater. Chem. A 2015, 3, 22860–22867. [Google Scholar] [CrossRef]
  38. Wu, L.; Yu, Z.; Ye, Y.; Yang, Y.; Zeng, H.; Huang, J.; Huang, Y.; Zhang, Z.; Xiang, S. Sulfonated Periodic-Mesoporous-Organosilicas Column for Selective Separation of C2H2/CH4 Mixtures. J. Solid State Chem. 2018, 264, 113–118. [Google Scholar] [CrossRef]
  39. Sanchez, C.; Jeremias, F.; Ernst, S.J.; Henninger, S.K. Synthesis, Functionalization and Evaluation of Ethylene-Bridged PMOs as Adsorbents for Sorption Dehumidification and Cooling Systems. Microporous Mesoporous Mater. 2017, 244, 151–157. [Google Scholar] [CrossRef]
  40. Lu, D.; Chen, H.; Yan, X.; Wang, L.; Zhang, J. Ratiometric Hg2+ Sensor Based on Periodic Mesoporous Organosilica Nanoparticles and Förster Resonance Energy Transfer. J. Photochem. Photobiol. A Chem. 2015, 299, 125–130. [Google Scholar] [CrossRef]
  41. Qiu, X.; Han, S.; Hu, Y.; Gao, M.; Wang, H. Periodic Mesoporous Organosilicas for Ultra-High Selective Copper(Ii) Detection and Sensing Mechanism. J. Mater. Chem. A 2014, 2, 1493–1501. [Google Scholar] [CrossRef]
  42. Grösch, L.; Lee, Y.J.; Hoffmann, F.; Fröba, M. Light-Harvesting Three-Chromophore Systems Based on Biphenyl-Bridged Periodic Mesoporous Organosilica. Chem. Eur. J. 2015, 21, 331–346. [Google Scholar] [CrossRef] [PubMed]
  43. Zhou, Z.; Taylor, R.N.K.; Kullmann, S.; Bao, H.; Hartmann, M. Mesoporous Organosilicas With Large Cage-Like Pores for High Efficiency Immobilization of Enzymes. Adv. Mater. 2011, 23, 2627–2632. [Google Scholar] [CrossRef]
  44. Aggad, D.; Jimenez, C.M.; Dib, S.; Croissant, J.G.; Lichon, L.; Laurencin, D.; Richeter, S.; Maynadier, M.; Alsaiari, S.K.; Boufatit, M.; et al. Gemcitabine Delivery and Photodynamic Therapy in Cancer Cells via Porphyrin-Ethylene-Based Periodic Mesoporous Organosilica Nanoparticles. ChemNanoMat 2018, 4, 46–51. [Google Scholar] [CrossRef]
  45. Inagaki, S.; Guan, S.; Fukushima, Y.; Ohsuna, T.; Terasaki, O. Novel Mesoporous Materials with a Uniform Distribution of Organic Groups and Inorganic Oxide in Their Frameworks. J. Am. Chem. Soc. 1999, 121, 9611–9614. [Google Scholar] [CrossRef]
  46. Melde, B.J.; Holland, B.T.; Blanford, C.F.; Stein, A. Mesoporous Sieves with Unified Hybrid Inorganic/Organic Frameworks. Chem. Mater. 1999, 11, 3302–3308. [Google Scholar] [CrossRef]
  47. Asefa, T.; MacLachlan, M.J.; Coombs, N.; Ozin, G.A. Periodic Mesoporous Organosilicas with Organic Groups inside the Channel Walls. Nature 1999, 402, 867–871. [Google Scholar] [CrossRef]
  48. Kuroki, M.; Asefa, T.; Whitnal, W.; Kruk, M.; Yoshina-Ishii, C.; Jaroniec, M.; Ozin, G.A. Synthesis and Properties of 1,3,5-Benzene Periodic Mesoporous Organosilica (PMO): Novel Aromatic PMO with Three Point Attachments and Unique Thermal Transformations. J. Am. Chem. Soc. 2002, 124, 13886–13895. [Google Scholar] [CrossRef] [PubMed]
  49. Esquivel Merino, M.D. Síntesis, Caracterización y Aplicaciones de Materiales Periódicos Mesoporosos Organosilícicos. Ph.D. Thesis, Universidad de Córdoba, Córdoba, Spain, 2011. Available online: https://helvia.uco.es/xmlui/bitstream/handle/10396/5222/9788469447673 (accessed on 24 October 2022).
  50. Sim, K.; Lee, N.; Kim, J.; Cho, E.B.; Gunathilake, C.; Jaroniec, M. CO2 Adsorption on Amine-Functionalized Periodic Mesoporous Benzenesilicas. ACS Appl. Mater. Interfaces 2015, 7, 6792–6802. [Google Scholar] [CrossRef] [PubMed]
  51. Sim, S.; Cho, E.-B. Multifunctional Periodic Mesoporous Benzene-Silicas for Evaluation of CO2 Adsorption at Standard Temperature and Pressure. Microporous Mesoporous Mater. 2019, 293, 109816. [Google Scholar] [CrossRef]
  52. Kirren, P.; Barka, L.; Rahmani, S.; Bondon, N.; Donzel, N.; Trens, P.; Bessière, A.; Raehm, L.; Charnay, C.; Durand, J.O. Periodic Mesoporous Organosilica Nanoparticles for CO2 Adsorption at Standard Temperature and Pressure. Molecules 2022, 27, 4245. [Google Scholar] [CrossRef] [PubMed]
  53. Wei, Y.; Li, X.; Zhang, R.; Liu, Y.; Wang, W.; Ling, Y.; El-Toni, A.M.; Zhao, D. Periodic Mesoporous Organosilica Nanocubes with Ultrahigh Surface Areas for Efficient CO2 Adsorption. Sci. Rep. 2016, 6, 20769. [Google Scholar] [CrossRef]
  54. Liu, M.; Lu, X.; Shi, L.; Wang, F.; Sun, J. Periodic Mesoporous Organosilica with a Basic Urea-Derived Framework for Enhanced Carbon Dioxide Capture and Conversion Under Mild Conditions. ChemSusChem 2017, 10, 1110–1119. [Google Scholar] [CrossRef]
  55. Rekha, P.; Sharma, V.; Mohanty, P. Synthesis of Cyclophosphazene Bridged Mesoporous Organosilicas for CO2 Capture and Cr (VI) Removal. Microporous Mesoporous Mater. 2016, 219, 93–102. [Google Scholar] [CrossRef]
  56. Xu, X.; Song, C.; Andresen, J.M.; Miller, B.G.; Scaroni, A.W. Novel Polyethylenimine-Modified Mesoporous Molecular Sieve of MCM-41 Type as High-Capacity Adsorbent for CO2 Capture. Energy Fuels 2002, 16, 1463–1469. [Google Scholar] [CrossRef]
  57. Lourenço, M.A.O.; Siquet, C.; Sardo, M.; Mafra, L.; Pires, J.; Jorge, M.; Pinto, M.L.; Ferreira, P.; Gomes, J.R.B. Interaction of CO2 and CH4 with Functionalized Periodic Mesoporous Phenylene-Silica: Periodic DFT Calculations and Gas Adsorption Measurements. J. Phys. Chem. C 2016, 120, 3863–3875. [Google Scholar] [CrossRef]
  58. Burleigh, M.C.; Markowitz, M.A.; Jayasundera, S.; Spector, M.S.; Thomas, C.W.; Gaber, B.P. Mechanical and Hydrothermal Stabilities of Aged Periodic Mesoporous Organosilicas. J. Phys. Chem. B 2003, 107, 12628–12634. [Google Scholar] [CrossRef]
  59. Burleigh, M.C.; Jayasundera, S.; Thomas, C.W.; Spector, M.S.; Markowitz, M.A.; Gaber, B.P. A Versatile Synthetic Approach to Periodic Mesoporous Organosilicas. Colloid Polym. Sci. 2004, 282, 728–733. [Google Scholar] [CrossRef]
  60. International Centre for Diffraction Data (ICDD). The Powder Diffraction; ICDD: Newtown Square, PA, USA, 2003. [Google Scholar]
  61. Cantador-Fernandez, D.; Suescum-Morales, D.; Esquivel, D.; Jiménez, J.R.; Fernández-Rodriguez, J.M. CO2 Adsorption by Ethane Periodic Mesoporous Organosilica at Low Temperatures and High Pressure. J. Environ. Chem. Eng. 2023, 11, 110582. [Google Scholar] [CrossRef]
  62. Langmuir, I. The Constitution and Fundamental Properties of Solids and Liquids. Part I. Solids. J. Am. Chem. Soc. 1916, 38, 2221–2295. [Google Scholar] [CrossRef]
  63. Langmuir, I. The Adsorption of Gases on Plane Surfaces of Glass, Mica and Platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef]
  64. Freundlich, H. Über Die Adsorption in Lösungen. Z. Phys. Chem. 1907, 57U, 385–470. [Google Scholar] [CrossRef]
  65. Sips, R. On the Structure of a Catalyst Surface. J. Chem. Phys. 1948, 16, 490–495. [Google Scholar] [CrossRef]
  66. Tóth, J. Uniform Interpretation of Gas/Solid Adsorption. Adv. Colloid Interface Sci. 1995, 55, 1–239. [Google Scholar] [CrossRef]
  67. Dubinin, M.M. Generalization of the Theory of Volume Filling of Micropores to Nonhomogeneous Microporous Structures. Carbon 1985, 23, 373–380. [Google Scholar] [CrossRef]
  68. Dubinin, M.M.; Zaverina, E.D. Sorbtsiya I Struktura Aktivnykh Uglei. 6. Strukturnye Tipy Aktivnykh Uglei. J. Phys. Chem. 1949, 23, 1129–1140. [Google Scholar]
  69. Temkin, M.; Pyzhev, V. Kinetics of Ammonia Synthesis on Promoted Iron Catalyst. Acta Phys. Chim. USSR 1940, 12, 327–356. [Google Scholar]
  70. Temkin, M.; Levich, V. Adsorption Equilibrium on Heterogeneous Surfaces. J. Phys. Chem. 1946, 20, 961–974. [Google Scholar]
  71. Liang, Y.; Liu, X.; Allen, M.R. Measuring and Modeling Surface Sorption Dynamics of Organophosphate Flame Retardants on Impervious Surfaces. Chemosphere 2018, 193, 754–762. [Google Scholar] [CrossRef] [PubMed]
  72. Limousin, G.; Gaudet, J.-P.; Charlet, L.; Szenknect, S.; Barthes, V.; Krimissa, M. Sorption Isotherms: A Review on Physical Bases, Modeling and Measurement. Appl. Geochem. 2007, 22, 249–275. [Google Scholar]
  73. Hu, Q.; Wang, Q.; Feng, C.; Zhang, Z.; Lei, Z.; Shimizu, K. Insights into Mathematical Characteristics of Adsorption Models and Physical Meaning of Corresponding Parameters. J. Mol. Liq. 2018, 254, 20–25. [Google Scholar] [CrossRef]
  74. Sparks, D.L. Environmental Soil Chemistry, 2nd ed.; Academic Press (Elsevier): Cambridge, MA, USA, 2003; ISBN 978-0-12-656446-4. [Google Scholar]
  75. Saadi, R.; Saadi, Z.; Fazaeli, R.; Fard, N.E. Monolayer and Multilayer Adsorption Isotherm Models for Sorption from Aqueous Media. Korean J. Chem. Eng. 2015, 32, 787–799. [Google Scholar] [CrossRef]
  76. Lee, H.I.; Pak, C.; Yi, S.H.; Shon, J.K.; Kim, S.S.; So, B.G.; Chang, H.; Yie, J.E.; Kwon, Y.-U.; Kim, J.M. Systematic Phase Control of Periodic Mesoporous Organosilicas Using Gemini Surfactants. J. Mater. Chem. 2005, 15, 4711–4717. [Google Scholar] [CrossRef]
  77. Haul, R.S.J.; Gregg, K.S.W. Sing: Adsorption, Surface Area and Porosity; Academic Press: London, UK, 1982; Volume 86. [Google Scholar]
  78. Sing, K.S.W.; Everett, D.H.; Haul, R.A.W.; Moscou, L.; Pierotti, R.A.; Rouquerol, J.; Siemieniewska, T. Reporting Physisorption Data for Gas/Solid Systems. In Handbook of Heterogeneous Catalysis; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2008; pp. 1217–1230. ISBN 9783527610044. [Google Scholar]
  79. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  80. Giles, C.H.; MacEwan, T.H.; Nakhwa, S.N.; Smith, D. 786. Studies in Adsorption. Part XI. A System of Classification of Solution Adsorption Isotherms, and Its Use in Diagnosis of Adsorption Mechanisms and in Measurement of Specific Surface Areas of Solids. J. Chem. Soc. 1960, 3973–3993. [Google Scholar] [CrossRef]
  81. Esfandiari, B.; Monajjemi, M. Physical Adsorption between Mono and Diatomic Gases inside of Carbon Nanotube with Respect to Potential Energy. J. Phys. Theor. Chem. 2013, 10, 31–42. [Google Scholar]
  82. Bhagiyalakshmi, M.; Lee, J.Y.; Jang, H.T. Synthesis of Mesoporous Magnesium Oxide: Its Application to CO2 Chemisorption. Int. J. Greenh. Gas Control 2010, 4, 51–56. [Google Scholar] [CrossRef]
  83. Wang, Y.; Du, T.; Song, Y.; Che, S.; Fang, X.; Zhou, L. Amine-Functionalized Mesoporous ZSM-5 Zeolite Adsorbents for Carbon Dioxide Capture. Solid State Sci. 2017, 73, 27–35. [Google Scholar] [CrossRef]
  84. Niu, M.; Yang, H.; Zhang, X.; Wang, Y.; Tang, A. Amine-Impregnated Mesoporous Silica Nanotube as an Emerging Nanocomposite for CO2 Capture. ACS Appl. Mater. Interfaces 2016, 8, 17312–17320. [Google Scholar] [CrossRef] [PubMed]
  85. Li, K.; Jiang, J.; Tian, S.; Yan, F.; Chen, X. Polyethyleneimine–Nano Silica Composites: A Low-Cost and Promising Adsorbent for CO2 Capture. J. Mater. Chem. A 2015, 3, 2166–2175. [Google Scholar] [CrossRef]
  86. Lo, S.-H.; Chien, C.-H.; Lai, Y.-L.; Yang, C.-C.; Lee, J.J.; Raja, D.S.; Lin, C.-H. A Mesoporous Aluminium Metal–Organic Framework with 3 Nm Open Pores. J. Mater. Chem. A 2013, 1, 324–329. [Google Scholar] [CrossRef]
  87. Si, X.; Zhang, J.; Li, F.; Jiao, C.; Wang, S.; Liu, S.; Li, Z.; Zhou, H.; Sun, L.; Xu, F. Adjustable Structure Transition and Improved Gases (H2, CO2) Adsorption Property of Metal–Organic Framework MIL-53 by Encapsulation of BNHx. Dalt. Trans. 2012, 41, 3119–3122. [Google Scholar] [CrossRef]
  88. Bourrelly, S.; Llewellyn, P.L.; Serre, C.; Millange, F.; Loiseau, T.; Férey, G. Different Adsorption Behaviors of Methane and Carbon Dioxide in the Isotypic Nanoporous Metal Terephthalates MIL-53 and MIL-47. J. Am. Chem. Soc. 2005, 127, 13519–13521. [Google Scholar] [CrossRef]
  89. Zhang, Z.; Huang, S.; Xian, S.; Xi, H.; Li, Z. Adsorption Equilibrium and Kinetics of CO2 on Chromium Terephthalate MIL-101. Energy Fuels 2011, 25, 835–842. [Google Scholar] [CrossRef]
  90. Zhou, Z.; Mei, L.; Ma, C.; Xu, F.; Xiao, J.; Xia, Q.; Li, Z. A Novel Bimetallic MIL-101(Cr, Mg) with High CO2 Adsorption Capacity and CO2/N2 Selectivity. Chem. Eng. Sci. 2016, 147, 109–117. [Google Scholar] [CrossRef]
  91. Zhao, Z.; Li, Z.; Lin, Y.S. Adsorption and Diffusion of Carbon Dioxide on Metal−Organic Framework (MOF-5). Ind. Eng. Chem. Res. 2009, 48, 10015–10020. [Google Scholar] [CrossRef]
  92. Millward, A.R.; Yaghi, O.M. Metal-Organic Frameworks with Exceptionally High Capacity for Storage of Carbon Dioxide at Room Temperature. J. Am. Chem. Soc. 2005, 127, 17998–17999. [Google Scholar] [CrossRef] [PubMed]
  93. Furukawa, H.; Ko, N.; Go, Y.B.; Aratani, N.; Choi, S.B.; Choi, E.; Yazaydin, A.O.; Snurr, R.Q.; O’Keeffe, M.; Kim, J.; et al. Ultrahigh Porosity in Metal-Organic Frameworks. Science 2010, 329, 424–428. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Representation of PMO–benzene structure.
Figure 1. Representation of PMO–benzene structure.
Materials 17 02669 g001
Figure 2. XRD patterns of PMO–benzene. The goniometer speed was 0.02°⋅s−1.
Figure 2. XRD patterns of PMO–benzene. The goniometer speed was 0.02°⋅s−1.
Materials 17 02669 g002
Figure 3. Particle size for PMO–benzene.
Figure 3. Particle size for PMO–benzene.
Materials 17 02669 g003
Figure 4. Nitrogen adsorption–desorption isotherms and pore size distribution of PMO–benzene. In the inserted figure: (Blue line) Pore volume, (Red line) Cumulative pore volume.
Figure 4. Nitrogen adsorption–desorption isotherms and pore size distribution of PMO–benzene. In the inserted figure: (Blue line) Pore volume, (Red line) Cumulative pore volume.
Materials 17 02669 g004
Figure 5. Images of PMO–benzene: (A) TEM and (B) high-resolution TEM.
Figure 5. Images of PMO–benzene: (A) TEM and (B) high-resolution TEM.
Materials 17 02669 g005
Figure 6. CO2 isotherms at 0 °C, 10 °C, 20 °C and 35 °C.
Figure 6. CO2 isotherms at 0 °C, 10 °C, 20 °C and 35 °C.
Materials 17 02669 g006
Figure 7. CO2 adsorption cycles at 0 °C, 10 °C, and 35 °C.
Figure 7. CO2 adsorption cycles at 0 °C, 10 °C, and 35 °C.
Materials 17 02669 g007
Figure 8. Loss of CO2 adsorption capacity after 10 cycles at 0 °C, 10 °C, and 35 °C.
Figure 8. Loss of CO2 adsorption capacity after 10 cycles at 0 °C, 10 °C, and 35 °C.
Materials 17 02669 g008
Figure 9. Fit curves of various models at 0 °C, 10 °C, 20 °C, and 35 °C for PMO–benzene.
Figure 9. Fit curves of various models at 0 °C, 10 °C, 20 °C, and 35 °C for PMO–benzene.
Materials 17 02669 g009
Table 1. Pore structure for PMO–benzene.
Table 1. Pore structure for PMO–benzene.
SBET
(m2·g−1)
Smc 1 (m2·g−1)Vp 2
(cm3·g−1)
Dp 3
(nm)
PMO–benzene9281390.683.47
1 Micropore surface; 2 single-point pore volume; 3 average pore size diameter.
Table 2. Langmuir and Freundlich fitting results.
Table 2. Langmuir and Freundlich fitting results.
LangmuirFreundlich
qmXLqmcKLSLR2KfnnfR2
(mg·g−1) (mg·g−1)(atm−1)(m2·g−1) (mg·g−1·atm(−1/n)) (1/n)
PMO–benzene 0 °C474.301.220388.664.538904.090.975574.1980.6631.5080.991
PMO–benzene 10 °C429.931.218353.014.590821.180.982505.8800.6431.5560.998
PMO–benzene 20 °C392.351.215322.884.648751.080.990479.1240.6411.5590.997
PMO–benzene 35 °C357.141.238288.574.208671.270.997468.5820.6921.4450.996
Table 3. Sips and Toth model fitting results.
Table 3. Sips and Toth model fitting results.
SipsToth
qSKSnSR2qTKTnTR2
(mg·g−1)(atm−1) (mg·g−1)(atm−1)
PMO–benzene 0 °C375.892.6310.7800.999349.584.8370.6900.996
PMO–benzene 10 °C384.791.9810.7801.000357.724.3030.7600.999
PMO–benzene 20 °C426.790.7090.7601.000436.590.9990.3101.000
PMO–benzene 35 °C--------
Table 4. Dubinin–R and Temkin fitting results.
Table 4. Dubinin–R and Temkin fitting results.
Dubinin–RaduskevichTemkin
qDβER2KTkBbTkR2
(mg·g−1)(mol2·kJ−2)(kJ·mol−1) (atm−1) (kJ·mol−1)
PMO–benzene 0 °C440.540.0353.7800.89727.711143.0960.0160.847
PMO–benzene 10 °C391.480.0323.9730.92630.895120.8950.0190.900
PMO–benzene 20 °C344.760.0284.2080.96137.432100.1810.0240.949
PMO–benzene 35 °C290.230.0254.4780.97542.53881.6390.0310.959
Table 5. CO2 capture for different materials.
Table 5. CO2 capture for different materials.
AdsorbentT Isotherm (°C)Pressure (atm)Capacity Adsorption (mg·g−1)Ref.
PMO–benzene25122[50]
PMO–benzene modified251133.32
PMO–benzene (A-LB)25177.44[51]
PMO–benzene (A-LBEO)25169.52
PMO–Ethane Np py0 and 25168.2 and 40.5[52]
PMO–Ethane Np Etbipy0 and 25173.04 and 41.8
PMO–Ethane Np iPrbipy0 and 25199.44 and 45.7
PMO-–Ethane 0162.48[53]
PMO-UDF0≈152.8[54]
CPMOs0196.36[55]
MCM-41-modified751133[56]
NH2-Ph-PMO25≈10114.4[57]
TiO2/Graphene25182.72[20]
MgO 25179.2[82]
ZSM-5 Mesoporous40139.6[83]
MSiNTs-PEI50 850.6121[84]
CeO2 Mesoporous2510391.6[21]
PEI50751138.16[85]
MOF-Al 01124.52[86]
MIL-53 (BNHx)01198[87]
MIL-47 (V)3120506[88]
MIL-10125301007.6[89]
MIL-101 (Cr, Mg)251145.2[90]
MOF-522.85192.4[91]
MOF-742542457[92]
MOF-17725421493
MOF-20025502400[93]
HT-MgAl-CO30≈35142.02[24]
Organohydrotalcite TDD035176.66[23]
PMO–benzene0≈34638.2This work
10≈34465.2
20≈34346.7
35≈34266.0
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cantador-Fernandez, D.; Esquivel, D.; Jiménez, J.R.; Fernández-Rodríguez, J.M. Use of Periodic Mesoporous Organosilica–Benzene Adsorbent for CO2 Capture to Reduce the Greenhouse Effect. Materials 2024, 17, 2669. https://doi.org/10.3390/ma17112669

AMA Style

Cantador-Fernandez D, Esquivel D, Jiménez JR, Fernández-Rodríguez JM. Use of Periodic Mesoporous Organosilica–Benzene Adsorbent for CO2 Capture to Reduce the Greenhouse Effect. Materials. 2024; 17(11):2669. https://doi.org/10.3390/ma17112669

Chicago/Turabian Style

Cantador-Fernandez, David, Dolores Esquivel, José Ramón Jiménez, and José María Fernández-Rodríguez. 2024. "Use of Periodic Mesoporous Organosilica–Benzene Adsorbent for CO2 Capture to Reduce the Greenhouse Effect" Materials 17, no. 11: 2669. https://doi.org/10.3390/ma17112669

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