Next Article in Journal
Polymer/Enzyme Composite Materials—Versatile Catalysts with Multiple Applications
Next Article in Special Issue
Oxalamide Based Fe(II)-MOFs as Potential Electrode Modifiers for Glucose Detection
Previous Article in Journal
Head vs. Tail Squaramide–Naphthalimide Conjugates: Self-Assembly and Anion Binding Behaviour
Previous Article in Special Issue
Enhancing of CO Uptake in Metal-Organic Frameworks by Linker Functionalization: A Multi-Scale Theoretical Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing NO Uptake in Metal-Organic Frameworks via Linker Functionalization. A Multi-Scale Theoretical Study

by
Charalampos G. Livas
1,
Emmanuel Tylianakis
1,2 and
George E. Froudakis
1,*
1
Department of Chemistry, University of Crete, Voutes Campus, GR-71003 Heraklion, Crete, Greece
2
Department of Materials Science and Technology, University of Crete, Voutes Campus, GR-71003 Heraklion, Crete, Greece
*
Author to whom correspondence should be addressed.
Chemistry 2022, 4(4), 1300-1311; https://doi.org/10.3390/chemistry4040086
Submission received: 12 September 2022 / Revised: 4 October 2022 / Accepted: 13 October 2022 / Published: 18 October 2022

Abstract

:
In the present work, ab initio calculations and Monte Carlo simulations were combined to investigate the effect of linker functionalization on nitric oxide (NO)’s storage ability of metal–organic frameworks (MOFs). The binding energy (BE) of nitric oxide with a set of forty-two strategically selected, functionalized benzenes was investigated using Density Functional Theory calculations at the RI-DSD-BLYP/def2-TZVPP level of theory. It was found that most of the functional groups (FGs) increased the interaction strength compared to benzene. Phenyl hydrogen sulfate (–OSO3H) was the most promising among the set of ligands, with an enhancement of 150%. The organic linker of IRMOF-8 was modified with the three top-performing functional groups (–OSO3H, –OPO3H2, –SO3H). Their ability for NO adsorption was investigated using Grand Canonical Monte Carlo (GCMC) simulations at an ambient temperature and a wide pressure range. The results showed great enhancement in NO uptake constituting the above-mentioned FGs, suggesting them to be promising modification candidates in a plethora of porous materials.

1. Introduction

Nitric oxide (NO) or nitrogen monoxide is a colorless gas and one of the principal oxides of nitrogen (NOx). The latter is a family of highly reactive gases that are largely released into the atmosphere as a result of the combustion of fossil fuels [1]. The majority (43%) of NO emissions originate from road transport and are also the dominant component (95%) of all gaseous NOx effluents [2]. Consequently, it is the focus of NOx removal efforts. Both the Environmental Protection Agency (EPA) of the United States of America and the European Environment Agency (EEA) regard NO as an important air pollutant among other gases such as NH3 and SO2 [3]. Its toxicity influences the environment and human health [4]. Air pollution and chronic exposure to hazardous gases can cause respiratory and cardiovascular diseases such as lung cancer and heart failure. Almost all of the global population (99%) are living in areas with poor air quality, a fact that leads to more than four million deaths every year, according to the World Health Organization (WHO) [5]. There are estimates that the global economic cost of tackling air pollution and its health effects has increased to USD 8.1 trillion which is equivalent to 6.1 percent of global GDP [6]. Moreover, NOx can react with water, oxygen, and other chemicals to produce acid rain [7]. It has negative biological consequences, especially in aquatic areas, and with dry deposition, it causes metal corrosion and can negatively impact buildings such as monuments [8].
Among the solutions being investigated to reduce NO emissions into the atmosphere is the discovery of new materials with high adsorption capacity for NO. Metal–organic frameworks (MOFs) have aroused a lot of attention as promising gas-adsorption materials. MOFs are organic–inorganic hybrid materials consisting of metal ions or clusters joined by organic linkers, as discovered by O. Yaghi in 1995 [9]. Most MOFs have an open porous framework, which gives them essential features, including low density, large surface area, and high porosity. These characteristics result in the adsorption of gases in greater quantities and at lower pressures. MOF characteristics may be fine-tuned for each application because of the diversity and tunability of their structure. Gas storage, ion exchange, molecular separation, and heterogeneous catalysis are possible uses [10,11,12,13,14,15].
Important techniques have been used to improve the nitric oxide storage capacity of MOFs. To modify the nature of the pore surface and increase the interaction between NO and the framework, the organic linkers can be functionalized, and the metal nodes serve as binding sites for the unitary structure. Aside from the benefits outlined above, certain MOFs feature open-metal sites (OMSs), which might be used wisely for specific applications [16]. Xiao et al. conducted experimental measurements for NO adsorption on HKUST-1. The latter gave an adsorption capacity of 9.0 mmol·g−1 at 196 K and 1 bar. At the same pressure and at 298 K the adsorption capacity was ~3 mmol·g−1 [17]. In addition, McKinlay et al., synthesized and tested Fe-based MIL-88A, MIL-88B, MIL-88B-NO2, and MIL-88B-2OH for NO adsorption capacity. MIL-88A performed the best, with an adsorption capacity of 2.5 mmol·g−1 [18].
The aim of this study is to investigate and propose new MOF materials with improved nitric oxide adsorption capabilities. Following this pathway, IRMOF-8 was incorporated with strategically selected functional groups (FGs) in order to enhance the interaction between the gas and the material, thus increasing the adsorption capacity for nitric oxide. First, we calculate the binding energy between functionalized benzene rings and the NO molecule, and then, the top-performing functional groups are introduced on the phenyl ring of the organic linker of IRMOF-8 to investigate the impact on the NO adsorption profile.
Our group has applied the aforementioned strategy in several studies that focus on different gases, and the results are indicative of the effectiveness of linker functionalization [19,20,21,22,23]. For instance, Raptis et al. [20] showed that the introduction of numerous FGs increases the binding energy (BE) between the functionalized benzene ring and nitrogen dioxide by 170% compared to the unfunctionalized one. This increase leads to a 16-fold enhancement in gravimetric adsorption (mmol·g−1) at 1.2 bar and 298 K. In addition, the work of Frysali et al. [21] revealed that the presence of a sulfate anion in the phenyl ring reinforces the interaction between the latter and the carbon dioxide (−22.6 kJ/mol) almost 2-fold.
Considering that aromatic rings are the building blocks of many MOFs’ organic linkers, we investigated the interaction of a NO gas molecule with a series of 42 carefully selected substituted benzene rings. The generation of electron redistribution plots and electrostatic potential maps of the dimers and of the functionalized benzenes, respectively, was required to shed light on the nature of the interactions. Finally, in order to assess the effect of linker functionalization on the nitric oxide uptake profile, we conducted Grand Canonical Monte Carlo (GCMC) simulations for IRMOF-8 modified with the three top-performing FGs. The total gravimetric and volumetric uptake isotherms were obtained at 298 K and for a wide pressure range.

2. Methodology

Full geometry optimizations were employed using the ORCA program package [24,25] at the DSD-BLYP/def2-TZVPP level of theory [26,27,28] to investigate the binding strength between NO and a large set of 42 functionalized benzene molecules. Resolution of identity (RI) approximation was used along with the auxiliary basis set to speed up the calculation time [29]. The Basis Set Superposition Error (BSSE) was accounted for and corrected using the Counterpoise method as proposed by Boys and Bernardi [30].
In order to obtain a better understanding of the interactions between the dimers, we constructed electrostatic potential maps of the functionalized benzene monomers and electron density redistribution plots of the dimers. The visualization was carried out using the gOpenMol graphics program [31,32].
To shed light on the effect of linker functionalization on the adsorption capacities of NO in MOFs, IRMOF-8 was chosen and was modified by the three functional groups that showed the greatest interaction with NO. The parent and modified organic linkers can be seen in Figure 1. Monte Carlo simulations were employed in the Grand Canonical ensemble at 298 K for a pressure range up to 100 bar using the RASPA software package [33].
For the description of host–sorbate interactions, the Lennard-Jones [34] + Coulomb potentials were used. Equation (1) represents the aforementioned potential.
V i j = 4 ε i j [ ( σ i j r i j ) 12 ( σ i j r i j ) 6 ] + q i q j 4 π ε 0 r i j
Here, ε0 is the vacuum permittivity constant, qi and qj are the corresponding partial charges for atoms i and j, rij is the interatomic distance between interacting atoms i and j, and eij and σij are the LJ potential well depth and the repulsion distance between atoms i and j, respectively.
Nitric oxide was treated using the COMPASS model [35], and during the simulations, the bond length was 1.152 Å and was not allowed to vary. For the van der Waals interactions, the potential parameters used for the nitrogen and oxygen atom were ε/kb = 79.5 K, σ = 3.0 Å, and ε/kb = 96.9 Κ, σ = 2.8 Å, respectively. The corresponding parameters for each MOF framework were obtained via the UFF force field [36]. Lorentz–Berthelot mixing rules [37] were employed to describe the interactions between the MOF framework and the gas molecule. To better describe the intermolecular interactions between NO molecules and the functional group atoms, we fitted the classical potential parameters to the quantum chemical data. More information about the fitting procedure can be found in the Supplementary Material. The cutoff value for the LJ potential was set to 12.8 Å. Electrostatic interactions were also considered. Point charges of 0.0288e and −0.0288e were appointed to the nitrogen and oxygen atoms, respectively, for the electrostatic interactions. The partial charges of the MOF atoms were calculated as implemented in ORCA using the CHELPG method [38]. The resulting excess charges were split in order to keep the molecular structure in the periodic box neutral.

3. Results and Discussion

In this study, the NO interaction with 42 functionalized benzene molecules was theoretically investigated. In Figure 2 and Figure 3, the geometries and the corresponding binding energies of the most promising systems can be seen. A detailed representation of the results for the remaining systems under study can be found in Figures S1 and S2 of the supporting information. The energetically most favorable structures, seen in Figure 3, can be divided in two categories: NO located above the benzene ring (Group 1) or towards the FG (Group 2). The binding energies calculated at the RI-DSD-BLYP/def2-TZVPP level, reported in Figure 2, range from 10.1 to 19.3 kJ/mol for NO–C6H5OOH and NO–C6H5OSO3H, respectively. The weakest interaction, as seen in Figure S1, was with C6H5CN, with a binding energy of 6.6 kJ/mol.
The functional group that showed the highest interaction energy was phenyl hydrogen sulfate with a corresponding binding energy of 19.4 kJ/mol, i.e., almost 3 times stronger binding than that of the unfunctionalized benzene ring (7.5 kJ/mol). As seen in Figure 3. RI-DSD-BLYP/def2-TZVPP optimized geometries for the 10 most energetically favorable dimers the NO molecule is above the phenyl ring, so it belongs to group 1 and is close to the FG. The nitrogen atom of the NO molecule is coordinated towards the hydrogen atom of the –OSO3H functional group at a distance of 2.04 Å, which is the smallest among all the systems under study.
The functional groups that constitute the top-three-performing FGs, along with OSO3H, are –SO3H and –OPO3H2, which have binding energies of 14.0 kJ/mol and 13.6 kJ/mol, respectively. In contrast with the best one, –OSO3H, the NO molecule in the aforementioned systems, is away from the benzene ring and is oriented next to the functional group, thus falling within group 2. We can see that the nitrogen atom of the NO molecule is close to the hydrogen atom of each of the FGs. Specifically, the corresponding distances are 2.10 Å for –SO3H and 2.13 Å for –OPO3H2. These distances are the second and third smallest ones observed among all the dimers, as seen in Figure 4. We can also notice that the distance between the aforementioned N and H atoms increases as the binding energy decreases in an almost linear correlation.
Figure S1 shows the optimized geometries for all the systems under study. We can observe that nitric oxide orients itself above the benzene ring in parallel mode in most cases. It is also slightly displaced towards the FG (Group 1). This suggests that, in addition to the predicted interaction between the dipole moment of NO and the π system of the benzene monomer, there is an extra interaction between NO and the atoms of the polar groups. There are cases where the NO molecule is placed next to the functional group and away from the benzene ring (Group 2). In these dimers, the NO molecule is closer to the FG compared with the cases where it is above the ring, as seen by comparing Figure 5a,b. Thus, NO is taking advantage of the additional interaction between the nitrogen atom of NO and the hydrogen atom of the FG.
Aiming to shed light on the nature of interactions, electrostatic potential maps were produced for all the functionalized benzene rings studied in this work. By doing so, we are capable of visualizing the charge distribution of the monomers and explaining the optimized geometries. In Figure 5, the electrostatic potential maps of the top-10-performing substituted benzene rings can be seen, and they can be seen in Figure S3 for all the benzene monomers under study. In both, we can see the corresponding map for the NO molecule, which has a high-potential region in the middle and two low-potential regions at both ends of the molecule. This, combined with the observation that the middle of all the benzene ring molecules have a low-potential value, plays an important role in the preference of the gas molecule being located above the benzene ring. In addition, the nitrogen atom of the NO molecule has a higher density of electrons than the oxygen atom. This explains the orientation of the N atom towards the hydrogen atom of the functional group which, as seen below, has a low electron density.
The next step in our study is to validate the functionalization’s effect in increasing NO adsorption in MOFs. To this end, IRMOF-8 was selected and modified with the top-three-performing functional groups. As seen in Figure 1, the organic linker of IRMOF-8 is 2,6-naphthalene dicarboxylate and it was functionalized. The gravimetric and volumetric uptake isotherms, which are shown in Figure 6 and Figure 7, were obtained for pressure up to 100 bar and at a temperature of 298 K using Grand Canonical Monte Carlo simulations.
In terms of the gravimetric uptake, we see a great enhancement, especially in the region of pressure lower than 40 bar. At ambient pressure, i.e., 1 bar, we see a 20-fold increase for the best-performing one, IRMOF8-OPO3H2. Concerning the uptake profile for pressure above 40 bar, we see that the parent material behaves better than the modified ones. This is explained by the fact that the gravimetric uptake is the amount of gas adsorbed divided by the weight of the system. By adding the functional groups to the framework, we increase the denominator. At larger pressures where the key factor is the pressure itself and not the binding energy between the framework and the gas molecule, we see similar quantities of gas adsorbed. So, the nominator for the different materials is similar but the denominator is smaller for the parent material, i.e., IRMOF-8.
Concerning the volumetric uptake, we produce analogous results, as demonstrated in Figure 7. At the low loading limit, the performance of the modified materials is improved and at a pressure above 40 bar, the difference between the modified IRMOF-8 materials and the parent one is decreasing. The parent material cannot surpass the volumetric uptake of the modified ones because the denominator of this indicator, i.e., cm3(STP)/cm3 is the same for all systems. By comparing the latter results with the highest nitric oxide uptake under these thermodynamic conditions found in the literature [16], one can conclude that the functionalization of MOFs is an important technique for the improvement of gas uptake. More specifically, Xiao et al. [17] found a 3 mmol/g storage capacity for HKUST-1 at 298K and 1 bar, which constitutes one third of the adsorption of IRMOF8-OPO3H2 in the same thermodynamic conditions.
We notice that the IRMOF-8 functionalized with –OPO3H2 has greater adsorption capacity than the ones modified with –OSO3H and –SO3H. This is in contrast with the binding energies of the three dimers of C6H5-X (X = –OSO3H, –SO3H, and –OPO3H2) as the order is exactly the opposite. We can attribute this to the fact that the binding energy of the global minimum is an important indicator of the adsorption capabilities of the material, but it is not the only one that plays a crucial role. The different binding sites and the corresponding binding energies have a strong impact on the adsorption capacity. Figure 8 is a schematic representation of the different local minima obtained after analyzing the data of the DFT study, with a more detailed depiction shown in Table S1 and Figure S7. We can see that C6H5-OPO3H2 has, on average, greater binding energies than C6H5-OSO3H and C6H5-SO3H. So, in addition to C6H5-OPO3H2 having lower binding energy for the first NO molecule, it has better binding sites for the other approaching NO molecules. This translates to greater adsorption capacities in GCMC simulations. The same applies to C6H5-SO3H and C6H5-OSO3H, and that is why we observe lower adsorption for the latter.

4. Conclusions

In this study, 42 functionalized benzene rings were examined by means of QM calculations concerning their interaction strength with the NO molecule. Almost all the functional groups show enhanced interaction when compared to the parent ring with the –OSO3H group, showing a binding energy of −19.3 kJ/mol. The nature of the interactions was further analyzed using electrostatic potential maps for the modified rings and redistribution electron density maps for the optimized dimers with NO molecules. The impact of functionalizing the benzene ring with the best-performing functional groups, i.e., –OSO3H, –SO3H, and –OPO3H2, was examined by performing Grand Canonical Monte Carlo simulations on the accordingly modified IRMOF-8 structures. The simulations were conducted at 298 K and up to 100 bar. All the functionalized materials show increased uptake under ambient conditions in both volumetric and gravimetric terms. Overall, this study establishes a systematic and transferrable database that may be utilized for additional experimental and computational investigations on the augmentation of the NO adsorption capabilities of porous materials.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/chemistry4040086/s1, Figure S1: Binding energies (kJ/mol) of the NO···C6H5-X dimers, calculated at the DSD-BLYP/def2-TZVPP level of theory. Basis Set Superposition Error (BSSE) was taken into consideration by the full counterpoise method; Figure S2: More energy-favorable configurations for the systems in this work; Figure S3: Electrostatic potential maps of all substituted monomers C6H5-X. Calculated using the DSD-BLYP/def2-TZVPP method with ORCA 4.2 and visualized with gOpenMol. The varying intensities are ranging from −0.03 to +0.03 Hartree·e−1. Red: Electron-poor regions-high potential value, Blue: electron-rich regions-low potential; Figure S4: Fitting of the (ε/kb, σ) parameters of the UFF potential on the QM data obtained from the DFT scan of NO over benzene; Figure S5: Fitting of the (ε/kb, σ) parameters of the UFF potential on the QM data obtained from the ab-DFT scan of NO over (a) C6H5-OSO3H (b) C6H5-SO3H (c) C6H5- OPO3H2; Figure S6: Gravimetric (a) and Volumetric (b) Nitric Oxide uptake isotherms for IRMOF-8 and IRMOF-8-n (n: –OSO3H, –SO3H, –OPO3H2); Figure S7: The geometries of the different local minima (RI-DSD-BLYP/def2-TZVPP); Table S1: The binding energies of the different local minima calculated at the RI-DSD-BLYP/def2-TZVPP level of theory (without BSSE correction); [24,25,30,31,32,36,37].

Author Contributions

Conceptualization, C.G.L. and G.E.F.; methodology, C.G.L., E.T. and G.E.F.; software, C.G.L. and E.T.; validation, C.G.L. and E.T.; formal analysis, C.G.L.; investigation, C.G.L.; resources, G.E.F.; data curation, C.G.L.; writing—original draft preparation, C.G.L.; writing—review and editing, C.G.L. and G.E.F.; visualization, C.G.L.; supervision, G.E.F. All authors have read and agreed to the published version of the manuscript.

Funding

European Regional Development Fund of the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship, and Innovation, under the call RESEARCH—CREATE—INNOVATE: T2EΔK-01976.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hill, S.C.; Douglas Smoot, L. Modeling of nitrogen oxides formation and destruction in combustion systems. Prog. Energy Combust. Sci. 2000, 26, 417–458. [Google Scholar] [CrossRef]
  2. European Environment Agency. Nitrogen Oxides (NOx) Emissions. Available online: https://www.eea.europa.eu/data-and-maps/indicators/eea-32-nitrogen-oxides-nox-emissions-1/assessment.2010-08-19.0140149032-3 (accessed on 12 July 2022).
  3. European Environment Agency. Air Pollution Sources. Available online: https://www.eea.europa.eu/themes/air/air-pollution-sources-1 (accessed on 6 June 2022).
  4. Mannucci, P.M.; Harari, S.; Martinelli, I.; Franchini, M. Effects on health of air pollution: A narrative review. Intern. Emerg. Med. 2015, 10, 657–662. [Google Scholar] [CrossRef]
  5. Air Pollution. Available online: https://www.who.int/data/gho/data/themes/theme-details/GHO/air-pollution (accessed on 6 June 2022).
  6. World Bank. The Global Health Cost of PM2.5 Air Pollution: A Case for Action Beyond 2021; World Bank: Washington, DC, USA, 2022; ISBN 978-1-4648-1816-5. [Google Scholar]
  7. US EPA. What Is Acid Rain? Available online: https://www.epa.gov/acidrain/what-acid-rain (accessed on 6 June 2022).
  8. US EPA. Effects of Acid Rain. Available online: https://www.epa.gov/acidrain/effects-acid-rain (accessed on 6 June 2022).
  9. Yaghi, O.M.; Li, G.; Li, H. Selective binding and removal of guests in a microporous metal–organic framework. Nature 1995, 378, 703–706. [Google Scholar] [CrossRef]
  10. Batten, S.R.; Robson, R. Interpenetrating Nets: Ordered, Periodic Entanglement. Angew. Chem. Int. Ed. Engl. 1998, 37, 1460–1494. [Google Scholar] [CrossRef]
  11. Férey, G. Some suggested perspectives for multifunctional hybrid porous solids. Dalton Trans. 2009, 23, 4400–4415. [Google Scholar] [CrossRef] [PubMed]
  12. Fischer, R.A.; Wöll, C. Functionalized Coordination Space in Metal–Organic Frameworks. Angew. Chem. Int. Ed. 2008, 47, 8164–8168. [Google Scholar] [CrossRef]
  13. Kitagawa, S.; Matsuda, R. Chemistry of coordination space of porous coordination polymers. Coord. Chem. Rev. 2007, 251, 2490–2509. [Google Scholar] [CrossRef]
  14. Kepert, C.J. Advanced functional properties in nanoporous coordination framework materials. Chem. Commun. 2006, 7, 695–700. [Google Scholar] [CrossRef]
  15. Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastré, J. Metal–organic frameworks—Prospective industrial applications. J. Mater. Chem. 2006, 16, 626–636. [Google Scholar] [CrossRef]
  16. Lin, J.; Ho, W.; Qin, X.; Leung, C.; Au, V.K.; Lee, S. Metal–Organic Frameworks for NOx Adsorption and Their Applications in Separation, Sensing, Catalysis, and Biology. Small 2022, 18, 2105484. [Google Scholar] [CrossRef]
  17. Xiao, B.; Wheatley, P.S.; Zhao, X.; Fletcher, A.J.; Fox, S.; Rossi, A.G.; Megson, I.L.; Bordiga, S.; Regli, L.; Thomas, K.M.; et al. High-Capacity Hydrogen and Nitric Oxide Adsorption and Storage in a Metal−Organic Framework. J. Am. Chem. Soc. 2007, 129, 1203–1209. [Google Scholar] [CrossRef]
  18. McKinlay, A.C.; Eubank, J.F.; Wuttke, S.; Xiao, B.; Wheatley, P.S.; Bazin, P.; Lavalley, J.-C.; Daturi, M.; Vimont, A.; De Weireld, G.; et al. Nitric Oxide Adsorption and Delivery in Flexible MIL-88(Fe) Metal–Organic Frameworks. Chem. Mater. 2013, 25, 1592–1599. [Google Scholar] [CrossRef]
  19. Livas, C.G.; Tylianakis, E.; Froudakis, G.E. Enhancing of CO Uptake in Metal–Organic Frameworks by Linker Functionalization: A Multi-Scale Theoretical Study. Chemistry 2022, 4, 43. [Google Scholar] [CrossRef]
  20. Raptis, D.; Livas, C.; Stavroglou, G.; Giappa, R.M.; Tylianakis, E.; Stergiannakos, T.; Froudakis, G.E. Surface Modification Strategy for Enhanced NO2 Capture in Metal–Organic Frameworks. Molecules 2022, 27, 3448. [Google Scholar] [CrossRef]
  21. Frysali, M.G.; Klontzas, E.; Tylianakis, E.; Froudakis, G.E. Tuning the interaction strength and the adsorption of CO2 in metal organic frameworks by functionalization of the organic linkers. Microporous Mesoporous Mater. 2016, 227, 144–151. [Google Scholar] [CrossRef]
  22. Giappa, R.M.; Tylianakis, E.; Di Gennaro, M.; Gkagkas, K.; Froudakis, G.E. A combination of multi-scale calculations with machine learning for investigating hydrogen storage in metal organic frameworks. Int. J. Hydrogen Energy 2021, 46, 27612–27621. [Google Scholar] [CrossRef]
  23. Stergiannakos, T.; Klontzas, E.; Tylianakis, E.; Froudakis, G.E. Enhancement of CO2 Adsorption in Magnesium Alkoxide IRMOF-10. J. Phys. Chem. C 2015, 119, 22001–22007. [Google Scholar] [CrossRef]
  24. Neese, F. Software update: The ORCA program system, version 4.0. WIREs Comput. Mol. Sci. 2018, 8, e1327. [Google Scholar] [CrossRef]
  25. Neese, F. The ORCA program system. WIREs Comput. Mol. Sci. 2012, 2, 73–78. [Google Scholar] [CrossRef]
  26. Kozuch, S.; Gruzman, D.; Martin, J.M.L. DSD-BLYP: A General Purpose Double Hybrid Density Functional Including Spin Component Scaling and Dispersion Correction. J. Phys. Chem. C 2010, 114, 20801–20808. [Google Scholar] [CrossRef]
  27. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef] [PubMed]
  28. Balanced Basis Sets of Split Valence, Triple Zeta Valence and Quadruple Zeta Valence Quality for H to Rn: Design and Assessment of Accuracy—Physical Chemistry Chemical Physics (RSC Publishing). Available online: https://pubs.rsc.org/en/content/articlelanding/2005/cp/b508541a (accessed on 7 June 2022).
  29. Skylaris, C.-K.; Gagliardi, L.; Handy, N.C.; Ioannou, A.G.; Spencer, S.; Willetts, A. On the resolution of identity Coulomb energy approximation in density functional theory. J. Mol. Struct. THEOCHEM 2000, 501–502, 229–239. [Google Scholar] [CrossRef] [Green Version]
  30. Boys, S.F.; Bernardi, F. The calculation of small molecular interactions by the differences of separate total energies. Some procedures with reduced errors. Mol. Phys. 1970, 19, 553–566. [Google Scholar] [CrossRef]
  31. Laaksonen, L. A graphics program for the analysis and display of molecular dynamics trajectories. J. Mol. Graph. 1992, 10, 33–34, 24. [Google Scholar] [CrossRef]
  32. Bergman, D.L.; Laaksonen, L.; Laaksonen, A. Visualization of solvation structures in liquid mixtures. J. Mol. Graph. Model 1997, 15, 301–306, 328–333. [Google Scholar] [CrossRef]
  33. Dubbeldam, D.; Calero, S.; Ellis, D.E.; Snurr, R.Q. RASPA: Molecular simulation software for adsorption and diffusion in flexible nanoporous materials. Mol. Simul. 2016, 42, 81–101. [Google Scholar] [CrossRef] [Green Version]
  34. Jones, J.E.; Chapman, S. On the determination of molecular fields.—II. From the equation of state of a gas. Proc. R. Soc. Lond. A 1924, 106, 463–477. [Google Scholar] [CrossRef] [Green Version]
  35. Yang, J.; Ren, Y.; Tian, A.; Sun, H. COMPASS Force Field for 14 Inorganic Molecules, He, Ne, Ar, Kr, Xe, H2, O2, N2, NO, CO, CO2, NO2, CS2, and SO2, in Liquid Phases. J. Phys. Chem. B 2000, 104, 4951–4957. [Google Scholar] [CrossRef]
  36. Rappe, A.K.; Casewit, C.J.; Colwell, K.S.; Goddard, W.A.; Skiff, W.M. UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024–10035. [Google Scholar] [CrossRef]
  37. Lorentz, H.A. Ueber die Anwendung des Satzes vom Virial in der kinetischen Theorie der Gase. Ann. Phys. 1881, 248, 127–136. [Google Scholar] [CrossRef]
  38. Breneman, C.M.; Wiberg, K.B. Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysis. J. Comput. Chem. 1990, 11, 361–373. [Google Scholar] [CrossRef]
Figure 1. The original and functionalized linkers of the IRMOF-8 structures studied using GCMC simulations: (a) the parent IRMOF-8 linker; the (b) –OSO3H, (c) –SO3H, and (d) –OPO3H2 functionalized linkers. Carbon, hydrogen, oxygen, sulfur, and phosphorus atoms are depicted as gray, white, red, yellow, and orange spheres, respectively.
Figure 1. The original and functionalized linkers of the IRMOF-8 structures studied using GCMC simulations: (a) the parent IRMOF-8 linker; the (b) –OSO3H, (c) –SO3H, and (d) –OPO3H2 functionalized linkers. Carbon, hydrogen, oxygen, sulfur, and phosphorus atoms are depicted as gray, white, red, yellow, and orange spheres, respectively.
Chemistry 04 00086 g001
Figure 2. The 10 substituted benzenes (blue) with the highest NO binding energy and the unsubstituted benzene (red) for comparison, calculated at the RI-DSD-BLYP/def2-TZVPP level of theory and corrected for BSSE.
Figure 2. The 10 substituted benzenes (blue) with the highest NO binding energy and the unsubstituted benzene (red) for comparison, calculated at the RI-DSD-BLYP/def2-TZVPP level of theory and corrected for BSSE.
Chemistry 04 00086 g002
Figure 3. RI-DSD-BLYP/def2-TZVPP optimized geometries for the 10 most energetically favorable dimers.
Figure 3. RI-DSD-BLYP/def2-TZVPP optimized geometries for the 10 most energetically favorable dimers.
Chemistry 04 00086 g003
Figure 4. Plot of the distance between the nitrogen atom of the NO molecule and the hydrogen atom of the functional group as a function of the binding energy for (a) Group 1 and (b) Group 2.
Figure 4. Plot of the distance between the nitrogen atom of the NO molecule and the hydrogen atom of the functional group as a function of the binding energy for (a) Group 1 and (b) Group 2.
Chemistry 04 00086 g004aChemistry 04 00086 g004b
Figure 5. The electrostatic potential maps for the 10 most energetically favorable dimers. Calculated using ORCA 4.2 and visualized using gOpenMol [27,28]. The varying intensities range from −0.03 to +0.03 Hartree·e−1. Red: electron-poor regions—high-potential value; blue: electron-rich regions—low-potential.
Figure 5. The electrostatic potential maps for the 10 most energetically favorable dimers. Calculated using ORCA 4.2 and visualized using gOpenMol [27,28]. The varying intensities range from −0.03 to +0.03 Hartree·e−1. Red: electron-poor regions—high-potential value; blue: electron-rich regions—low-potential.
Chemistry 04 00086 g005
Figure 6. Absolute Gravimetric isotherms for IRMOF8 and IRMOF8-n (n: –OSO3H, –SO3H, and –OPO3H2,) at T = 298 K and pressure range up to (a) 100 bar and (b) 0.10 bar.
Figure 6. Absolute Gravimetric isotherms for IRMOF8 and IRMOF8-n (n: –OSO3H, –SO3H, and –OPO3H2,) at T = 298 K and pressure range up to (a) 100 bar and (b) 0.10 bar.
Chemistry 04 00086 g006
Figure 7. Absolute volumetric isotherms for IRMOF8 and IRMOF8-n (n: –OSO3H, –SO3H, and –OPO3H2,) at T = 298 K and pressure ranges up to (a) 100 bar (b) 0.10 bar.
Figure 7. Absolute volumetric isotherms for IRMOF8 and IRMOF8-n (n: –OSO3H, –SO3H, and –OPO3H2,) at T = 298 K and pressure ranges up to (a) 100 bar (b) 0.10 bar.
Chemistry 04 00086 g007
Figure 8. Schematic representation of the different binding sites of the three top-performing functional groups (–OSO3H, –SO3H, –OPO3H2). The numbers in the oval shapes are the corresponding binding energies in kJ/mol calculated at the RI-DSD-BLYP/def2-TZVPP level of theory (without BSSE correction).
Figure 8. Schematic representation of the different binding sites of the three top-performing functional groups (–OSO3H, –SO3H, –OPO3H2). The numbers in the oval shapes are the corresponding binding energies in kJ/mol calculated at the RI-DSD-BLYP/def2-TZVPP level of theory (without BSSE correction).
Chemistry 04 00086 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Livas, C.G.; Tylianakis, E.; Froudakis, G.E. Enhancing NO Uptake in Metal-Organic Frameworks via Linker Functionalization. A Multi-Scale Theoretical Study. Chemistry 2022, 4, 1300-1311. https://doi.org/10.3390/chemistry4040086

AMA Style

Livas CG, Tylianakis E, Froudakis GE. Enhancing NO Uptake in Metal-Organic Frameworks via Linker Functionalization. A Multi-Scale Theoretical Study. Chemistry. 2022; 4(4):1300-1311. https://doi.org/10.3390/chemistry4040086

Chicago/Turabian Style

Livas, Charalampos G., Emmanuel Tylianakis, and George E. Froudakis. 2022. "Enhancing NO Uptake in Metal-Organic Frameworks via Linker Functionalization. A Multi-Scale Theoretical Study" Chemistry 4, no. 4: 1300-1311. https://doi.org/10.3390/chemistry4040086

Article Metrics

Back to TopTop