Next Article in Journal
Individual ZnO–Ag Hybrid Nanorods for Synergistic Fluorescence Enhancement Towards Highly Sensitive and Miniaturized Biodetection
Previous Article in Journal
Encapsulation of Sulforaphane from Cruciferous Vegetables in mPEG-PLGA Nanoparticles Enhances Cadmium’s Inhibitory Effect on HepG2 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 and SO2 Capture by Cryptophane-111 Porous Liquid: Insights from Molecular Dynamics Simulations and Computational Chemistry

by
Pablo Collado
1,2,
Manuel M. Piñeiro
1,3,* and
Martín Pérez-Rodríguez
2,3,*
1
Departamento de Física Aplicada, Universidade de Vigo, E36310 Vigo, Spain
2
Instituto de Química Física Blas Cabrera (IQF), (CSIC), C/Serrano 119, 28006 Madrid, Spain
3
Modelización y Simulacón de Materiales Nanoestructurados, Universidade de Vigo, Unidad Asociada al CSIC por el IQF, Campus Lagoas-Marcosende, 36310 Vigo, Spain
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(8), 616; https://doi.org/10.3390/nano15080616
Submission received: 10 March 2025 / Revised: 4 April 2025 / Accepted: 15 April 2025 / Published: 17 April 2025
(This article belongs to the Section Theory and Simulation of Nanostructures)

Abstract

:
A computational study of the encapsulation of a gaseous mixture of CO2 and SO2 in a Type II porous liquid is performed under different conditions. The system is composed of cryptophane-111 molecules dispersed in dichloromethane, and it is described using classic molecular dynamics at atomistic resolution. Gaseous CO2 tends to almost fully occupy cryptophane-111’s cavities during the first phases of simulation, and, afterwards, it is surpassed by SO2’s tendency for occupation. Calculations are performed at five different temperatures in the range of 273 K–310 K, finding a positive correlation between SO2 adsorption and temperature. An evaluation of the radial distribution function of SO2 and CO2 with respect to cryptophane-111 is employed to quantify the number of captured molecules, and an energy study using Density Functional Theory methods is also performed to evaluate the relative stability of the two gases inside the porous liquid.

Graphical Abstract

1. Introduction

Global industrialization has led to a rapid increase in the atmospheric greenhouse gas concentration and the release of other environmentally harmful compounds that have contributed to a set of global transformations [1,2,3,4]. With the aim of countering these emissions, different methods of capture have been developed over the years, especially for CO2 emissions [5]. Porous liquids (referred to as PLs hereafter for brevity) are an emerging material with high potential in gas capture/separation areas [6].
PLs were first described by O’Reilly et al. [7] as fluid media with permanent porosity. Firstly classified into three groups, namely, Type I, Type II, and Type III [8], recently, a new group, Type IV, was proposed by Bennet et al. [9].
Type I PLs are composed of liquids, where each individual molecule presents a permanent and rigid cavity, so they cannot be automatically filled, making them hard to synthesize because of the need for this neat state. They present high melting points and a tendency to crystallize outside certain ideal range conditions [10]. Adding flexible alkyl chains reduces their porosity, as well as their melting point [11]. Type II PLs are composed of rigid cage molecules and a sterically hindered solvent, which cannot enter the cage cavities. Type III PLs consist of Metal Organic Frameworks (MOFs) suspended in a sterically hindered solvent. The latter are easier to synthesize and eliminate the need for low melting points or high solubilities because of them being in suspension. However, they may present phase separation and generate precipitation due to the different natures of the pores and their distribution on the MOF, lowering the overall porosity. To avoid this, some studies have considered the use of nanoparticles or nanocrystals despite their large size and the consequent stability issues [12]. Finally, Type IV PLs are defined as MOF species with intrinsic porosity isolated in liquid form, meaning that they do not need a solvent for their suspension. So far, few studies have been carried out on PLs, and even less on gas absorption. Several studies have suggested the possibility of developing a renewable route to adsorb CO2 under mild conditions [13,14,15,16,17,18,19,20,21,22]. However, there is still limited knowledge about the adsorption of other greenhouse gases in porous liquids. As examples, we can cite the work of Oltean et al. [23], where the CH4 interaction with cryptophane-111 (C-111) in the presence of H2 is described, or the contribution by Dou et al. [24], where SO2 absorption/desorption is analyzed. In this study, using molecular dynamics (MD) and electronic Density Functional Theory (DFT) calculations, we analyze the process of the encapsulation of a SO2 and CO2 mixture in a Type II PL at different temperatures, deepening the understanding of the competitive process behind the encapsulation and the thermodynamics. The PL consists of C-111 as the porous molecule and dichloromethane (DCM) as the solvent. In previous studies, we also analyzed this PL when studying the encapsulation of H2, SO2, and CO2 [25,26]. C-111 is the smallest molecule from the cryptophane family, and it was first analyzed as a possible gas collector by Fogarty et al. [27]. Since then, it has been intensively studied for its remarkable complexation with Xe [27,28,29].
Cryptophanes are aromatic molecules, characterized by three symmetric folds of cyclotribencylene in a crown conformation, defining an inner cavity with a stable geometry. Because of stereoisomerism, they present a great variety of structures and properties, one of which is the possibility of enclathrating a guest molecule in their cavity, and the nature of their bonded parts and functional groups define the size and flexibility of the cage geometry [30]. The narrow inner cavity of C-111 makes it a suitable candidate molecule to study the encapsulation/separation of small gas molecules. Because of that, we selected SO2 and CO2 as guest molecules in view of their size, which is comparable to that of Xe, making their adsorption feasible. As for the solvent, DCM was selected due to the solubility of C-111. The inner volume of the cavity of C-111 molecules ranges from 32 Å3 to 72 Å3 in the most expanded state, and, despite DCM presenting a molecular volume within that interval, it cannot pass through the entrance of the cavity, as discussed by Buffeteau et al. [31]. CO2 is known for being quite soluble in DCM [32,33], and SO2 is commonly commercially presented dissolved in DCM, so their mutual solubility is not an issue.
The choice of this particular Type II porous liquid was motivated by various reasons, including the previous results and performance demonstrated by C-111 in our work [25,26], as well as the accessible simulation CPU times for this type of PL. Because of the environmental and overall purposes of this kind of material, we worked with atmospheric pressure and a range of different temperatures close to standard room temperature, avoiding the possible degradation of C-111 and the phase transitions of the gases. Additionally, our research group has performed molecular simulations to characterize gas diffusion through nanoporous structured solid materials, such as hydrates or clathrates [34,35,36,37,38,39], and the process of gas capture on PLs presents many similarities to these inclusion solids. The capture of the mixture of SO2 and CO2 has been recently studied using different approaches [40,41,42,43,44,45]. Considering an alternative for this study of other PLs, different issues have been taken into account: Type III PLs are bound for precipitation occurrence, possibly complicating the search for a viable option; Type I PLs present very complex molecular structures, making molecular simulations less accessible; and references available for Type IV PLs are still very scarce. Because of this, Type II PLs represent probably the most accessible option to currently approach this problem. In conclusion, based on the results of our previous works, we intend to achieve a better understanding of this kind of encapsulation with the use of comparative and qualitative explanations. With these studies, we aim to contribute to the theoretical understanding of potential alternatives to atmospheric greenhouse gas separation and capture, exploring the viability and efficiency of PLs in this context.

2. Computational Methods

We used the C-111, CO2, SO2, and DCM molecular topologies described in our previous works [25,26]. C-111’s atomic coordinates were retrieved from Joseph et al. [28], who studied the encapsulation of Xenon, and the corresponding topology was built using the Automated Topology Builder (ATB) v.3.0 repository [46]. The molecular model for DCM was also obtained through ATB from the work by Stroel et al. [47]. SO2 is a widely studied molecule, so examples of its topology were readily available at ATB, and we selected the “MOLID 370654” model. ATB uses a knowledge-based approach, in combination with QM calculations, to assign force-field parameters, and its results are known to be reliable [48]. A submitted molecule is initially optimized at the HF/STO-3G (or AM1 or PM3) level of theory, after which an initial draft output is available. Molecules with <50 atoms are then re-optimized at the B3LYP/6-31G* level of theory with the PCM implicit solvent model for water. The QM electrostatic potential (ESP) is then calculated from the B3LYP-optimized structure, which the ATB uses to obtain ESP-fitted charges. Finally, the QM Hessian is calculated for molecules with <40 atoms to improve the assignment of bond and angle terms.
Many CO2 molecular models developed for molecular simulation calculations are available in the literature. We decided to use the TraPPE [49] molecular model because of its widely reported excellent performance [50]. In previous studies, we used this CO2 molecular model in the study of solid–liquid equilibrium [51] and the description of CO2 hydrates [52] or CO2 organic clathrates [35], in all cases yielding excellent performance in systems including solid phases, essential to the present study.
All MD simulations were performed using GROMACS v.2021.5 software [53,54,55,56,57,58,59,60]. A typical calculation consisted of 100 ns of simulation with a 1 fs time step, except for when indicated otherwise. An initial energy minimization was performed using the conjugated gradient method to avoid molecular overlapping. Electrostatic long-range corrections were handled using the Particle Mesh Ewald (PME) technique [61], a 4th order of calculation, and a Fourier spacing of 0.1 nm. The dispersive van der Waals cut-off radius was set to 1 nm, and long-range dispersive corrections were applied to energy and pressure. A Nose–Hoover [62,63] thermostat was used to keep the temperature constant, with a 2 ps coupling constant, and a Parrinello–Rahman [64,65] barostat was used to set the pressure value, with a 4 ps coupling constant. All simulations were initiated by creating a bubble of the gas mixture in the simulation box, generated through the random insertion of C-111 and DCM molecules. Their final configurations were analyzed using GROMACS v.2021.5 tools and VMD v.1.9.4a51open-source software [66,67].
Fifteen molecular simulation setups were studied. The simulation boxes contained 14 C-111 molecules, 28 CO2 and SO2 molecules, and 600 DCM molecules.
Because DCM’s volume is smaller than the maximum cavity volume of C-111, special care was taken during the initial setup simulation to avoid the accidental insertion of a DCM molecule inside a C-111 cavity. During the simulation, DCM itself could not spontaneously access the cavity of C-111 because the cavity threshold was smaller than the average DCM molecule diameter.
After the energy minimization step, NpT molecular dynamics was performed at 1 bar and temperatures of 273 K, 283 K, 293 K, 300 K, and 310 K, maximizing the working temperature range between the boiling points of DCM and SO2. We built three different replica boxes for each setup and calculated the radial distribution functions for every temperature.
For the computation of the complexation energies, we employed four different DFT methods, as implemented in Gaussian 16, Revision C.01 [68], over individual molecules. We initially considered two well-known and reliable hybrid methods, suitable for the molecules tested: PBEh1PBE [69] and B3LYP [70,71,72,73]. Furthermore, due to the expected interaction type between the porous liquid and the captured gases, we used two additional functionals incorporating empirical dispersion terms: wB97XD [74,75] and the empirically corrected B3LYP, B3LYP-D3 [76]. All DFT calculations were performed using the 6-311+G(d) basis set, whose results are considered accurate for this kind of study [77], at an affordable computational cost. Indeed, a larger basis set did not provide significantly better results, being notably more expensive in terms of computational resources.

3. Results and Discussion

3.1. Simulation Analysis

Using the specifications mentioned before, simulation boxes were built with C-111 molecules, with CO2 and SO2 gaseous molecules in a DCM solution. We placed the CO2 and SO2 mixture in a bubble, trying to mimic the flow of a flue stream through the PL, as shown in Figure 1.
We built three replicas of the system at each of the five different temperatures. After the simulations finished, we analyzed the production trajectory using Gromacs tools and VMD. We obtained the C-111 occupancy rates of CO2 and SO2 at each temperature, observing the behavior of both gases with C-111 throughout the simulation and examining how and how many interactions occur with C-111. We observed a greater selectivity towards CO2, as expected from our previous studies [25,26]. However, this selectivity towards SO2 decreased as the temperature increased. In Figure 2, we can observe an increase in SO2 adsorption rates, contrary to the CO2 trend. SO2 adsorption was more than two times larger at the highest temperature in the range tested, while CO2 suffered a continuous decrease. By carefully examining the simulation trajectories, we observed different events. At 310 K, CO2 encapsulated in a CO2/C-111 complex was substituted by SO2. CO2 was captured at 12.12 ns of the simulation and was pushed at 98.4 ns. This event resembles one in our previous work [25], where a CO2 molecule pushed another CO2 molecule. This event occurred at 300 K, meaning that this oddity could be improved at higher temperatures. However, SO2 molecules, when adsorbed, were never observed to be replaced by the SO2 or CO2 molecules present. In Figure 3, we can observe an example of these events, with the approximation of a CO2 molecule to an encapsulated SO2. This CO2 seemed to slightly attract SO2, with no other effect observed other than a little reorientation. In some cases, CO2 and SO2 molecules were observed to approach the C-111 cavity with the preferential orientation to obtain access, and, in these cases, SO2 was always the one captured. We also noticed an increase in the SO2 adsorption rates in the final simulation stages, as more CO2 was being desorbed. This suggests that SO2 could be favored thermodynamically over CO2, but its size makes it less likely to access the pore first.
The observed result can be interpreted as follows: As the temperatures increases, the SO2 kinetic barrier is lowered enough to begin compensating for the kinetic preference of smaller molecules such as CO2, prioritizing a higher proportion of SO2 being captured despite there being more CO2 in the pore. With the aim of checking this assumption, we performed a series of radial distribution function calculations, comparing both gases individually with C-111 as a reference. In Figure 4, the computed RDFs for adsorption are shown, with a logarithmic scale on the Y axis for clarity. As a reference, we used the center of mass of both molecules to represent the RDFs. The physical barrier (C-111 cage) was placed at 0.23 nm, as shown in the plot minima. However, for CO2, we obtained non-zero values at this distance. As discussed, adsorbed CO2 can be displaced by another CO2 or SO2 molecule. Because of this, a CO2 molecule can partially enter the C-111 cage while another CO2 molecule is already in there, explaining the presence of these non-negligible values. This phenomenon does not occur with SO2, which is not observed to leave the pore once captured. In Figure 5, we can observe this tendency of CO2 to be encapsulated since the first stages of the simulation and how SO2 starts to appear inside C-111 right after the initial 50 ns of the simulation have passed. We can also observe the increased tendency for the encapsulation of SO2 in the 310 K image while the profile of entering later than CO2 is still maintained. We performed Cumulative Sum RDF calculations, searching for a quantitative method to follow the capture process through the simulation. These Cumulative Sum RDFs represent the increase in pore occupation by the indicated molecule against time, with y = 1 being the full occupancy of every C-111 molecule present in the simulation box. These results are shown in Figure 6. In Figure 7, we can observe the difference represented in the Cumulative Sum RDFs in terms of occupation numbers vs. time, showing how there is no linear correlation between occupation and time. In these figures, a greater difference in capture for CO2 at the lowest temperature is evidenced, reaching the maximum occupation in the middle stages of the simulation. For SO2, we observe low occupancy rates at the beginning of the simulation; however, at 310 K, this occupation duplicates, thus reducing CO2 capture.
Taking into consideration the possibility of C-111 precipitation, which occurs in Type III PLs, we decided to search for aggregates that could influence the capture process. Several clusters of C-111 formed occasionally during the simulations. We then expected some kind of interaction between the C-111 molecules, but the overall absorption rates did not seem to be affected because neither absorption nor desorption occurred just before, after, or during this cluster formation, and redispersion occurred, as shown in Figure 8.

3.2. DFT Calculations and Stability

In order to analyze the relative stability of the adsorbed species, we performed additional DFT simulations. Through these, we obtained the energies of C-111 and the adsorbed gases individually, as well as the energies of the resultant encapsulated complex. Afterwards, we subtracted the energies of the component molecules from the corresponding complexes, Δ E C - 111 [ g u e s t ] = E C - 111 [ g u e s t ] ( E C - 111 + E g u e s t ) , to directly obtain the interaction energies. We did not independently consider the deformation energies of the molecules, going from the free state to the complex. Other significant contributions, such as zero-point, thermal, or solvation energies, were also not taken into account. The reason for this is that, in the present study, our aims were mainly to check the validity of the conclusions extracted from the MD simulations and to establish the binding preferences for C-111, but not to obtain a precise value of those energies. However, we were interested in the influence that the dispersive interactions may have on the relative stability; therefore, we considered methods without (PBE and B3LYP) and with dispersion corrections (wB97XD and B3LYP-D3) to check this point, as mentioned in Section 2.
The DFT results confirm what the MD simulations suggested, with SO2 being more stable than CO2 within the C-111 cavity in terms of energy, as shown in Figure 9. Each method represents both pairs as thermodynamically stable, except for C-111[CO2] in B3LYP, which yields a very low positive value. B3LYP shows the highest difference in terms of energy among the four methods used. The use of dispersion greatly increases the stabilization energy of both complexes, consistently keeping C-111[SO2] as the most stable. In B3LYP-D3, a significant difference between energies is observed in comparison with the wB97XD method, which perhaps can be attributed to the increased complexity of the B3LYP-D3 method [78]. We find many similarities in terms of energy in pairs (for all models, either including dispersion or not). Moreover, in the case of the global profile, the trend is coincident for all methods: Δ E C - 111 [ S O 2 ] is more stable for every method than Δ E C - 111 [ C O 2 ] . If we perform the subtraction of Δ E C - 111 [ S O 2 ] minus Δ E C - 111 [ C O 2 ] , we obtain almost identical differences of −0.09 eV, −0.11 eV, and −0.12 eV for PBE, B3LYP (non-dispersive), and wB97XD. However, the B3LYP-D3 method gives a result of −0.40 eV. All of these results point to C-111[SO2] as being the most stable complex, supporting the conclusions drawn from molecular dynamics, with a thermodynamical preference of C-111 for encapsulating SO2 rather than CO2, with the latter being favored by kinetics. This kinetic effect is supported by the relative size of both molecules, with CO2 being smaller and less sterically hindered to access the pore. The chemical difference between both molecules has to also be pointed out. With both consisting of a central molecule bonded to two oxygen molecules, CO2 presents a linear geometry ( O C O ^ = 180°), while SO2 has an angular one ( O S O ^ = 119°). Because of this, CO2 has no dipole moment but only a quadrupolar moment, while SO2 presents a significant dipole moment of 1633 D [79], close to the water value. In addition, due to the angular geometry, SO2 is effectively wider than it would be as a linear molecule. The differences in the overall geometry and dipole moment of CO2 and SO2 explain the easier encapsulation of CO2 and the greater stability of C-111[SO2] complexes.

3.3. Future Approaches

We discussed the viability of CO2 and SO2 encapsulation in Type II porous liquids, as well as feasible methods to improve the selectivity and efficiency of SO2 capture vs. CO2. However, further tests must be carried out to identify an efficient way to desorb these gases from C-111, achieving a renewable source of CO2 and SO2 recovery. We expect an easier desorption of CO2, as it has been shown to be able to leave the C-111 cavity by being pushed by another molecule. However, SO2 seems to be firmly absorbed and harder to recover, so more trials should be carried out. This is also reaffirmed with the DFT values obtained, with SO2 being more stable. Changing the solvent for a more sustainable one is also a desirable objective and could probably improve our results of the desorption process, generating a better understanding of how others factors work for this kind of material. Another solvent may allow one to expand the working temperature range without losing the interest of maintaining it near room temperature. The proposal of a feasible, renewable, and non-pollutant system for CO2 and SO2 capture is the final objective of all of our studies.

4. Conclusions

This study is a contribution to the theoretical modeling and characterization of porous liquids. It analyzed the feasibility of using a Type II porous liquid for the encapsulation of greenhouse gases through adsorption. The selectivity of the PL was greater towards CO2 than SO2. However, we observed an increasing trend of SO2 adsorption as the temperature increased. Although the CO2 occupation rates were higher, we observed that SO2 absorption was favored in the later stages of the simulation. DFT calculations showed a greater stability of the C-111/SO2 complex vs. the CO2 one, suggesting a thermodynamical preference for absorption by SO2 and a kinetic preference for CO2, in consonance with the MD simulation results. Possible reasons for this preference were also discussed, pointing out the relevance of the chemical properties of the gases.
These results highlight porous liquids as a potential alternative for the development of innovative strategies for greenhouse gas capture and separation due to their intrinsic porosity and high selectivity towards some of these gases. The possibility of using them as a renewable option for the capture of these gases opens an alternative path to mitigate their harmful environmental impact.

Author Contributions

Conceptualization, M.M.P. and M.P.-R.; methodology, M.M.P. and M.P.-R.; formal analysis, P.C.; investigation, P.C.; data curation, P.C.; writing—original draft preparation, P.C.; writing—review and editing, all authors; supervision, M.M.P. and M.P.-R.; funding acquisition, M.M.P. and M.P.-R. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge support from the following grants: Ref. PID2021-125081NB-I00, financed by Spanish Ministerio de Ciencia e Innovación, and Ref. CNS2022-135881, financed by MCIN/AEI/10.13039/501100011033 and the European Union NextGenerationEU/PRTR.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors acknowledge support from the following grants: Ref. PID2021-125081NB-I00, financed by Spanish Ministerio de Ciencia e Innovación, and Ref. CNS2022-135881, financed by MCIN/AEI/10.13039/501100011033 and the European Union NextGenerationEU/PRTR. The authors also acknowledge the computational resources provided by Centro de Supercomputación de Galicia (CESGA, www.cesga.es).

Conflicts of Interest

The authors declare no conflicts of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

The following abbreviations are used throughout this manuscript:
ATBAutomatic Topology Builder
DCMdichloromethane
C-111cryptophane-111
MOFMetal Organic Framework
PDBProtein Data Bank
PLporous liquid

References

  1. Barnes, J.; Dove, M.; Lahsen, M.; Mathews, A.; McElwee, P.; McIntosh, R.; Moore, F.; O’Reilly, J.; Orlove, B.; Puri, R.; et al. Contribution of anthropology to the study of climate change. Nat. Clim. Change 2013, 3, 541–544. [Google Scholar] [CrossRef]
  2. Anukwonke, C.C.; Tambe, E.B.; Nwafor, D.C.; Malik, K.T. Climate Change and Interconnected Risks to Sustainable Development. In Climate Change: The Social and Scientific Construct; Bandh, S.A., Ed.; Springer International Publishing: Cham, Switzerland, 2022; pp. 71–86. [Google Scholar] [CrossRef]
  3. Lawrence, J.; Blackett, P.; Cradock-Henry, N.A. Cascading climate change impacts and implications. Clim. Risk Manag. 2020, 29, 100234. [Google Scholar] [CrossRef]
  4. Räsänen, A.; Juhola, S.; Nygren, A.; Käkönen, M.; Kallio, M.; Monge, A.; Kanninen, M. Climate change, multiple stressors and human vulnerability: A systematic review. Reg. Environ. Change 2016, 16, 2291–2302. [Google Scholar] [CrossRef]
  5. Shukla, P.; Skea, J.; Slade, R.; Khourdajie, A.A.; van Diemen, R.; McCollum, D.; Pathak, M.; Some, S.; Vyas, P.; Fradera, R.; et al. IPCC, 2022: Climate Change 2022: Mitigation of Climate Change. Contribution of Working Group III to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; IPCC: Geneva, Switzerland, 2022. [Google Scholar]
  6. Wang, D.; Xin, Y.; Yao, D.; Li, X.; Ning, H.; Zhang, H.; Wang, Y.; Ju, X.; He, Z.; Yang, Z.; et al. Shining Light on Porous Liquids: From Fundamentals to Syntheses, Applications and Future Challenges. Adv. Funct. Mater. 2022, 32, 2104162. [Google Scholar] [CrossRef]
  7. O’Reilly, N.; Giri, N.; James, S.L. Porous liquids. Chem.—Eur. J. 2007, 13, 3020–3025. [Google Scholar] [CrossRef] [PubMed]
  8. James, S.L. The dam bursts for porous liquids. Adv. Mater. 2016, 28, 5712–5716. [Google Scholar] [CrossRef]
  9. Bennett, T.D.; Coudert, F.X.; James, S.L.; Cooper, A.I. The changing state of porous materials. Nat. Mater. 2021, 20, 1179–1187. [Google Scholar] [CrossRef] [PubMed]
  10. Jie, K.; Zhou, Y.; Ryan, H.P.; Dai, S.; Nitschke, J.R. Porous liquids: Engineering permanent porosity into liquids. Adv. Mater. 2021, 33, 2170136. [Google Scholar] [CrossRef]
  11. Melaugh, G.; Giri, N.; Davidson, C.E.; James, S.L.; Del Pópolo, M.G. Designing and understanding permanent microporosity in liquids. Phys. Chem. Chem. Phys. 2014, 16, 9422–9431. [Google Scholar] [CrossRef]
  12. Liu, H.; Liu, B.; Lin, L.C.; Chen, G.; Wu, Y.; Wang, J.; Gao, X.; Lv, Y.; Pan, Y.; Zhang, X.; et al. A hybrid absorption-adsorption method to efficiently capture carbon. Nat. Commun. 2014, 5, 5147. [Google Scholar] [CrossRef]
  13. Avila, J.; Červinka, C.; Dugas, P.Y.; Pádua, A.A.H.; Costa Gomes, M. Porous ionic liquids: Structure, stability, and gas absorption mechanisms. Adv. Mater. Interfaces 2021, 8, 2001982. [Google Scholar] [CrossRef]
  14. Li, P.; Chen, H.; Schott, J.A.; Li, B.; Zheng, Y.; Mahurin, S.M.; Jiang, D.E.; Cui, G.; Hu, X.; Wang, Y.; et al. Porous liquid zeolites: Hydrogen bonding-stabilized H-ZSM-5 in branched ionic liquids. Nanoscale 2019, 11, 1515–1519. [Google Scholar] [CrossRef]
  15. Avila, J.; Lepre, L.F.; Santini, C.; Tiano, M.; Denis-Quanquin, S.; Szeto, K.C.; Padua, A.; Costa Gomes, M. High-performance porous ionic liquids for low pressure CO2 capture. Angew. Chem. 2021, 60, 12876–12882. [Google Scholar] [CrossRef] [PubMed]
  16. Li, P.; Wang, D.; Zhang, L.; Liu, C.; Wu, F.; Wang, Y.; Wang, Z.; Zhao, Z.; Wu, W.; Liang, Y.; et al. An in situ coupling strategy toward porous carbon liquid with permanent porosity. Small 2021, 17, e2006687. [Google Scholar] [CrossRef] [PubMed]
  17. Ahmad, M.Z.; Fuoco, A. Porous liquids—Future for CO2 Capture and Separation? Curr. Res. Green Sustain. Chem. 2021, 4, 100070. [Google Scholar] [CrossRef]
  18. Brand, M.C.; Rankin, N.; Cooper, A.I.; Greenaway, R.L. Photoresponsive Type III Porous Liquids. Chem.—Eur. J. 2023, 29, e202202848. [Google Scholar] [CrossRef]
  19. Sarmad, S.; Nikjoo, D.; Mikkola, J.P. Innovative CO2 capture technologies: Exploring the potential of porous liquids containing deep eutectic solvents and hypercrosslinked polymers. Sep. Purif. Technol. 2025, 352, 128189. [Google Scholar] [CrossRef]
  20. Cheng, B.; Ma, J.; Ju, X.; Wei, F.; Cheng, W.; Li, P.; Liu, X. Metal-organic framework-based porous liquids via surface functionalization engineering for selective gas adsorption. Sep. Purif. Technol. 2024, 350, 127994. [Google Scholar] [CrossRef]
  21. Ji, D.; Xue, C.; Wen, Y.; Zhao, Y.; Gong, W.; Li, Y. Preparation and CO2 adsorption behavior of ZIF-67-based porous liquids: A molecular dynamics study. Sep. Purif. Technol. 2024, 347, 127564. [Google Scholar] [CrossRef]
  22. Li, Y.; Yuan, S. The Role of the Nanoconfinement Effect in the Adsorption of Carbon Dioxide by Porous Liquids. ACS Sustain. Chem. Eng. 2024, 12, 10455–10465. [Google Scholar] [CrossRef]
  23. Oltean, M. Theoretical Study of Chemical Reactivity: From Nano-Objects to Follow the Entropy via Metadynamics Simulations. Ph.D. Dissertation, Université Joseph-Fourier-Grenoble I, Saint-Martin-d’Hères, France, 2010. [Google Scholar]
  24. Dou, S.; Liu, K.; Feng, Y.; Wang, X.; Zhang, B.; Yang, C.; Sun, M.; Han, J.; Duan, E. A new strategy to effectively capture and recover SO2 based on the functional porous liquids. J. Clean. Prod. 2024, 467, 143006. [Google Scholar] [CrossRef]
  25. Collado, P.; Piñeiro, M.M.; Pérez-Rodríguez, M. Molecular Simulation of CO2 and H2 Encapsulation in a Nanoscale Porous Liquid. Nanomaterials 2023, 13, 409. [Google Scholar] [CrossRef]
  26. Collado, P.; Piñeiro, M.M.; Pérez-Rodríguez, M. Molecular Simulation of SO2 Separation and Storage Using a Cryptophane-Based Porous Liquid. Int. J. Mol. Sci. 2024, 25, 2718. [Google Scholar] [CrossRef]
  27. Fogarty, H.A.; Berthault, P.; Brotin, T.; Huber, G.; Desvaux, H.; Dutasta, J.P. A cryptophane core optimized for xenon encapsulation. J. Am. Chem. Soc. 2007, 129, 10332–10333. [Google Scholar] [CrossRef] [PubMed]
  28. Joseph, A.I.; Lapidus, S.H.; Kane, C.M.; Holman, K.T. Extreme confinement of xenon by cryptophane-111 in the solid state. Angew. Chem. 2015, 127, 1491–1495. [Google Scholar] [CrossRef]
  29. Mecozzi, S.; Rebek, J., Jr. The 55 % solution: A formula for molecular recognition in the liquid state. Chemistry 1998, 4, 1016–1022. [Google Scholar] [CrossRef]
  30. El-Ayle, G.; Holman, K.T. Cryptophanes; Elsevier: Amsterdam, The Netherlands, 2017; pp. 199–249. [Google Scholar] [CrossRef]
  31. Buffeteau, T.; Pitrat, D.; Daugey, N.; Calin, N.; Jean, M.; Vanthuyne, N.; Ducasse, L.; Wien, F.; Brotin, T. Chiroptical properties of cryptophane-111. Phys. Chem. Chem. Phys. 2017, 19, 18303–18310. [Google Scholar] [CrossRef] [PubMed]
  32. Shirono, K.; Morimatsu, T.; Takemura, F. Gas Solubilities (CO2, O2, Ar, N2, H2, and He) in Liquid Chlorinated Methanes. J. Chem. Eng. Data 2008, 53, 1867–1871. [Google Scholar] [CrossRef]
  33. Vonderheiden, F.H.; Eldridge, J.W. The System Carbon Dioxide-Methylene Chloride. Solubility, Vapor Pressure, Liquid Density, and Activity Coefficients. J. Chem. Eng. Data 1963, 8, 20–21. [Google Scholar] [CrossRef]
  34. Pérez-Rodríguez, M.; Otero-Fernández, J.; Comesaña, A.; Fernández-Fernández, A.M.; Piñeiro, M.M. Simulation of Capture and Release Processes of Hydrogen by β-Hydroquinone Clathrate. ACS Omega 2018, 3, 18771–18782. [Google Scholar] [CrossRef]
  35. Torré, J.P.; Gornitzka, H.; Coupan, R.; Dicharry, C.; Pérez-Rodríguez, M.; Comesaña, A.; Piñeiro, M.M. Insights into the Crystal Structure and Clathration Selectivity of Organic Clathrates Formed with Hydroquinone and (CO2 + CH4) Gas Mixtures. J. Phys. Chem. C 2019, 123, 14582–14590. [Google Scholar] [CrossRef]
  36. Fernández-Fernández, A.M.; Conde, M.M.; Pérez-Sánchez, G.; Pérez-Rodríguez, M.; Piñeiro, M.M. Molecular simulation of methane hydrate growth confined into a silica pore. J. Mol. Liq. 2022, 362, 119698. [Google Scholar] [CrossRef]
  37. Méndez-Morales, T.; Montes-Campos, H.; Pérez-Rodríguez, M.; Piñeiro, M.M. Evaluation of hydrogen storage ability of hydroquinone clathrates using molecular simulations. J. Mol. Liq. 2022, 360, 119487. [Google Scholar] [CrossRef]
  38. Rodríguez-García, B.; Piñeiro, M.M.; Pérez-Rodríguez, M. Diffusion and dynamics of noble gases in hydroquinone clathrate channels. J. Chem. Phys. 2023, 158, 044503. [Google Scholar] [CrossRef]
  39. Rodríguez García, B.; Piñeiro, M.M.; Pérez-Rodríguez, M. Influence of Lennard-Jones Parameters in the Temperature Dependence of Real Gases Diffusion through Nanochannels. Nanomaterials 2023, 13, 1534. [Google Scholar] [CrossRef] [PubMed]
  40. Jiang, L.; Xiao, Q.; Tao, H.; Wang, S.; Lin, W.; Wang, X.; Li, H.; Shi, G.; Wang, C. Highly Selective Absorption of Low-Concentration SO2 over CO2 in Flue Gas through Cation-Tunable Protic Ionic Liquids. Energy Fuels 2023, 37, 1180–1186. [Google Scholar] [CrossRef]
  41. Fan, Y.L.; Zhang, H.P.; Yin, M.J.; Krishna, R.; Feng, X.F.; Wang, L.; Luo, M.B.; Luo, F. High Adsorption Capacity and Selectivity of SO2 over CO2 in a Metal–Organic Framework. Inorg. Chem. 2021, 60, 4–8. [Google Scholar] [CrossRef] [PubMed]
  42. Taylor, S.F.R.; McClung, M.; McReynolds, C.; Daly, H.; Greer, A.J.; Jacquemin, J.; Hardacre, C. Understanding the Competitive Gas Absorption of CO2 and SO2 in Superbase Ionic Liquids. Ind. Eng. Chem. Res. 2018, 57, 17033–17042. [Google Scholar] [CrossRef]
  43. Choi, W.Y.; Lee, D.; Jang, K.; Yoo, Y.; Park, J. A novel process for selective absorption of CO2/SO2 mixture gas with a single absorbent derived from seawater-based industrial wastewater. Desalination 2023, 560, 116661. [Google Scholar] [CrossRef]
  44. Zhang, X.; Feng, X.; Li, H.; Peng, J.; Wu, Y.; Hu, X. Cyano-Containing Protic Ionic Liquids for Highly Selective Absorption of SO2 from CO2: Experimental Study and Theoretical Analysis. Ind. Eng. Chem. Res. 2016, 55, 11012–11021. [Google Scholar] [CrossRef]
  45. Qinlan, L.; Feng, B.; Zhuhan, L.; Zhou, Q.; Zhang, Y.; Li, N. Experimental study on simultaneous absorption and desorption of CO2/SO2/NOx using aqueous MDEA + DMSO solutions. Energy Fuels 2018, 32, 3647–3659. [Google Scholar] [CrossRef]
  46. Malde, A.K.; Zuo, L.; Breeze, M.; Stroet, M.; Poger, D.; Nair, P.C.; Oostenbrink, C.; Mark, A.E. An Automated force field Topology Builder (ATB) and repository: Version 1.0. J. Chem. Theory Comput. 2011, 7, 4026–4037. [Google Scholar] [CrossRef] [PubMed]
  47. Stroet, M.; Caron, B.; Visscher, K.M.; Geerke, D.P.; Malde, A.K.; Mark, A.E. Automated topology builder version 3.0: Prediction of solvation free enthalpies in water and hexane. J. Chem. Theory Comput. 2018, 14, 5834–5845. [Google Scholar] [CrossRef] [PubMed]
  48. Koziara, K.B.; Stroet, M.; Malde, A.K.; Mark, A.E. Testing and validation of the Automated Topology Builder (ATB) version 2.0: Prediction of hydration free enthalpies. J. Comput.-Aided Mol. Des. 2014, 28, 221–233. [Google Scholar] [CrossRef]
  49. Potoff, J.J.; Siepmann, J.I. Vapor–liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogen. AIChE J. 2001, 47, 1676–1682. [Google Scholar] [CrossRef]
  50. Hamani, A.W.S.; Bazile, J.P.; Hoang, H.; Luc, H.T.; Daridon, J.L.; Galliero, G. Thermophysical properties of simple molecular liquid mixtures: On the limitations of some force fields. J. Mol. Liq. 2020, 303, 112663. [Google Scholar] [CrossRef]
  51. Pérez-Sánchez, G.; González-Salgado, D.; Piñeiro, M.M.; Vega, C. Fluid-solid equilibrium of carbon dioxide as obtained from computer simulations of several popular potential models: The role of the quadrupole. J. Chem. Phys. 2013, 138, 084506. [Google Scholar] [CrossRef]
  52. Míguez, J.M.; Conde, M.M.; Torré, J.P.; Blas, F.J.; Piñeiro, M.M.; Vega, C. Molecular dynamics simulation of CO2 hydrates: Prediction of three phase coexistence line. J. Chem. Phys. 2015, 142, 124505. [Google Scholar] [CrossRef]
  53. Bekker, H.; Berendsen, H.; Dijkstra, E.; Achterop, S.; Vondrumen, R.; Vanderspoel, D.; Sijbers, A.; Keegstra, H.; Renardus, M. Gromacs: A parallel computer for molecular dynamics simulations. In Physics Computing 92; DeGroot, R., Nadrchal, J., Eds.; World Scientific Publishing: Singapore, 1993; pp. 252–256. [Google Scholar]
  54. Berendsen, H.; van der Spoel, D.; van Drunen, R. GROMACS: A message-passing parallel molecular dynamics implementation. Comput. Phys. Commun. 1995, 91, 43–56. [Google Scholar] [CrossRef]
  55. Lindahl, E.; Hess, B.; Spoel, D. GROMACS 3.0: A package for molecular simulation and trajectory analysis. J. Mol. Model. 2001, 7, 306–317. [Google Scholar] [CrossRef]
  56. Van Der Spoel, D.; Lindahl, E.; Hess, B.; Groenhof, G.; Mark, A.E.; Berendsen, H.J.C. GROMACS: Fast, flexible, and free. J. Comput. Chem. 2005, 26, 1701–1718. [Google Scholar] [CrossRef]
  57. Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. GROMACS 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation. J. Chem. Theory Comput. 2008, 4, 435–447. [Google Scholar] [CrossRef]
  58. Pronk, S.; Páll, S.; Schulz, R.; Larsson, P.; Bjelkmar, P.; Apostolov, R.; Shirts, M.R.; Smith, J.C.; Kasson, P.M.; van der Spoel, D.; et al. GROMACS 4.5: A high-throughput and highly parallel open source molecular simulation toolkit. Bioinformatics 2013, 29, 845–854. [Google Scholar] [CrossRef] [PubMed]
  59. Páll, S.; Abraham, M.J.; Kutzner, C.; Hess, B.; Lindahl, E. Tackling Exascale Software Challenges in Molecular Dynamics Simulations with GROMACS. In Solving Software Challenges for Exascale: International Conference on Exascale Applications and Software; Markidis, S., Laure, E., Eds.; Springer International Publishing: Cham, Switzerland, 2015; pp. 3–27. [Google Scholar] [CrossRef]
  60. Abraham, M.J.; Murtola, T.; Schulz, R.; Páll, S.; Smith, J.C.; Hess, B.; Lindahl, E. GROMACS: High performance molecular simulations through multi-level parallelism from laptops to supercomputers. SoftwareX 2015, 1–2, 19–25. [Google Scholar] [CrossRef]
  61. Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An N log(N) method for Ewald sums in large systems. J. Chem. Phys. 1993, 98, 10089–10092. [Google Scholar] [CrossRef]
  62. Nosé, S. A molecular dynamics method for simulations in the canonical ensemble. Mol. Phys. 1984, 52, 255–268. [Google Scholar] [CrossRef]
  63. Hoover, W.G. Canonical dynamics: Equilibrium phase-space distributions. Phys. Rev. A 1985, 31, 1695. [Google Scholar] [CrossRef]
  64. Parrinello, M.; Rahman, A. Crystal Structure and Pair Potentials: A Molecular-Dynamics Study. Phys. Rev. Lett. 1980, 45, 1196–1199. [Google Scholar] [CrossRef]
  65. Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys 2007, 126, 014101. [Google Scholar] [CrossRef]
  66. Humphrey, W.; Dalke, A.; Schulten, K. VMD—Visual Molecular Dynamics. J. Mol. Graph. 1996, 14, 33–38. [Google Scholar] [CrossRef]
  67. Stone, J.; Gullingsrud, J.; Grayson, P.; Schulten, K. A System for Interactive Molecular Dynamics Simulation. In Proceedings of the 2001 ACM Symposium on Interactive 3D Graphics; Hughes, J.F., Séquin, C.H., Eds.; ACM: New York, NY, USA, 2001; pp. 191–194. [Google Scholar] [CrossRef]
  68. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian~16 Revision C.01; Gaussian Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  69. Ernzerhof, M.; Perdew, J.P. Generalized gradient approximation to the angle- and system-averaged exchange hole. J. Chem. Phys. 1998, 109, 3313–3320. [Google Scholar] [CrossRef]
  70. Vosko, S.H.; Wilk, L.; Nusair, M. Accurate spin-dependent electron liquid correlation energies for local spin density calculations: A critical analysis. Can. J. Phys. 1980, 58, 1200–1211. [Google Scholar] [CrossRef]
  71. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Phys. Chem. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  72. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef]
  73. Stephens, P.J.; Devlin, F.J.; Chabalowski, C.F.; Frisch, M.J. Ab Initio Calculation of Vibrational Absorption and Circular Dichroism Spectra Using Density Functional Force Fields. J. Phys. Chem. 1994, 98, 11623–11627. [Google Scholar] [CrossRef]
  74. Chai, J.D.; Head-Gordon, M. Systematic optimization of long-range corrected hybrid density functionals. J. Chem. Phys. 2008, 128, 084106. [Google Scholar] [CrossRef]
  75. Chai, J.D.; Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom–atom dispersion corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. [Google Scholar] [CrossRef] [PubMed]
  76. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef]
  77. Bálint, D.; Jäntschi, L. Comparison of Molecular Geometry Optimization Methods Based on Molecular Descriptors. Mathematics 2021, 9, 2855. [Google Scholar] [CrossRef]
  78. Cukras, J.; Sadlej, J. The influence of the dispersion corrections on the performance of DFT method in modeling HNgY noble gas molecules and their complexes. Chem. Phys. Lett. 2018, 691, 319–324. [Google Scholar] [CrossRef]
  79. Patel, D.; Margolese, D.; Dykea, T.R. Electric dipole moment of SO2 in ground and excited vibrational states. J. Chem. Phys. 1979, 70, 2740–2747. [Google Scholar] [CrossRef]
Figure 1. Snapshot of the initial configuration, showing the porous liquid (C-111 in DCM solvent) and gas bubble (CO2 and SO2) in the bottom-left part of the figure. SO2 is represented by yellow and red; CO2 is represented by blue and red.
Figure 1. Snapshot of the initial configuration, showing the porous liquid (C-111 in DCM solvent) and gas bubble (CO2 and SO2) in the bottom-left part of the figure. SO2 is represented by yellow and red; CO2 is represented by blue and red.
Nanomaterials 15 00616 g001
Figure 2. Number of occupied C-111 pores vs. T for CO2 (green-blue bars) and SO2 (yellow bars). As a guide, a dashed line is included for CO2, and a solid one is included for SO2.
Figure 2. Number of occupied C-111 pores vs. T for CO2 (green-blue bars) and SO2 (yellow bars). As a guide, a dashed line is included for CO2, and a solid one is included for SO2.
Nanomaterials 15 00616 g002
Figure 3. Snapshots of (left side) CO2 inside C-111 with an outsider CO2 and (right side) SO2 inside C-111 interacting with an outsider CO2. Elements are colored as follows: S, yellow; O, red; C, blue; and H, white.
Figure 3. Snapshots of (left side) CO2 inside C-111 with an outsider CO2 and (right side) SO2 inside C-111 interacting with an outsider CO2. Elements are colored as follows: S, yellow; O, red; C, blue; and H, white.
Nanomaterials 15 00616 g003
Figure 4. RDF of (A) CO2 vs. C-111 at 273 K and 310 K; (B) SO2 vs. C-111 at 273 K and 310 K.
Figure 4. RDF of (A) CO2 vs. C-111 at 273 K and 310 K; (B) SO2 vs. C-111 at 273 K and 310 K.
Nanomaterials 15 00616 g004aNanomaterials 15 00616 g004b
Figure 5. Adsorbed molecule RDF for (A) CO2 vs. C-111 at 273 K and 310 K; (B) SO2 vs. C-111 at 273 K and 310 K.
Figure 5. Adsorbed molecule RDF for (A) CO2 vs. C-111 at 273 K and 310 K; (B) SO2 vs. C-111 at 273 K and 310 K.
Nanomaterials 15 00616 g005
Figure 6. Mulative Sum RDF of (A) CO2 vs. C-111 at 273 K and 310 K; (B) SO2 vs. C-111 at 273 K and 310 K.
Figure 6. Mulative Sum RDF of (A) CO2 vs. C-111 at 273 K and 310 K; (B) SO2 vs. C-111 at 273 K and 310 K.
Nanomaterials 15 00616 g006
Figure 7. Number of occupied C-111 pores (out of 14) vs. time for CO2 (green-blue bars) and SO2 (yellow bars) in the base of the Cumulative RDF Sum of both temperatures.
Figure 7. Number of occupied C-111 pores (out of 14) vs. time for CO2 (green-blue bars) and SO2 (yellow bars) in the base of the Cumulative RDF Sum of both temperatures.
Nanomaterials 15 00616 g007
Figure 8. C-111 clusters during simulation.
Figure 8. C-111 clusters during simulation.
Nanomaterials 15 00616 g008
Figure 9. Binding energy in eV of the gas-C-111 complexes, Δ E C - 111 [ S O 2 ] (yellow bars) and Δ E C - 111 [ C O 2 ] (green-blue bars), obtained from different DFT methods.
Figure 9. Binding energy in eV of the gas-C-111 complexes, Δ E C - 111 [ S O 2 ] (yellow bars) and Δ E C - 111 [ C O 2 ] (green-blue bars), obtained from different DFT methods.
Nanomaterials 15 00616 g009
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

Collado, P.; Piñeiro, M.M.; Pérez-Rodríguez, M. CO2 and SO2 Capture by Cryptophane-111 Porous Liquid: Insights from Molecular Dynamics Simulations and Computational Chemistry. Nanomaterials 2025, 15, 616. https://doi.org/10.3390/nano15080616

AMA Style

Collado P, Piñeiro MM, Pérez-Rodríguez M. CO2 and SO2 Capture by Cryptophane-111 Porous Liquid: Insights from Molecular Dynamics Simulations and Computational Chemistry. Nanomaterials. 2025; 15(8):616. https://doi.org/10.3390/nano15080616

Chicago/Turabian Style

Collado, Pablo, Manuel M. Piñeiro, and Martín Pérez-Rodríguez. 2025. "CO2 and SO2 Capture by Cryptophane-111 Porous Liquid: Insights from Molecular Dynamics Simulations and Computational Chemistry" Nanomaterials 15, no. 8: 616. https://doi.org/10.3390/nano15080616

APA Style

Collado, P., Piñeiro, M. M., & Pérez-Rodríguez, M. (2025). CO2 and SO2 Capture by Cryptophane-111 Porous Liquid: Insights from Molecular Dynamics Simulations and Computational Chemistry. Nanomaterials, 15(8), 616. https://doi.org/10.3390/nano15080616

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