Next Article in Journal
The Composition Optimization of Curcumin-Loaded Double Oil–Water–Oil Emulsions and Their Stability Evaluation
Previous Article in Journal
An Overview on the Synthesis of Lamellarins and Related Compounds with Biological Interest
Previous Article in Special Issue
Iodophor-/H2O2-Mediated 2-Sulfonylation of Indoles and N-Methylpyrrole in Aqueous Phase
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Unveiling the Influence of Water Molecules for NF3 Removal by the Reaction of NF3 with OH: A DFT Study

1
Hunan Provincial Key Laboratory of Xiangnan Rare-Precious Metals Compounds Research and Application, School of Chemistry and Environmental Science, Xiangnan University, Chenzhou 423000, China
2
Key Laboratory of Electronic Functional Materials and Devices of Guangdong Province, School of Chemistry and Materials Engineering, Huizhou University, Huizhou 516007, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(17), 4033; https://doi.org/10.3390/molecules29174033
Submission received: 25 July 2024 / Revised: 22 August 2024 / Accepted: 23 August 2024 / Published: 26 August 2024
(This article belongs to the Special Issue Catalysis for Green Chemistry II)

Abstract

:
The removal of nitrogen trifluoride (NF3) is of significant importance in atmospheric chemistry, as NF3 is an important anthropogenic greenhouse gas. However, the radical species OH and O(1D) in atmospheric conditions are nonreactive towards NF3. It is necessary to explore possible ways to remove NF3 in atmosphere. Therefore, the participation of water molecules in the reaction of NF3 with OH was discussed, as water is abundant in the atmosphere and can form very stable complexes due to its ability to act as both a hydrogen bond donor and acceptor. Systemic DFT calculations carried out at the CBS-QB3 and ωB97XD/aug-cc-pVTZ level of theory suggest that water molecules could affect the NF3 + OH reaction as well. The energy barrier of the SN2 mechanism was decreased by 8.52 kcal/mol and 10.58 kcal/mol with the assistance of H2O and (H2O)2, respectively. Moreover, the presence of (H2O)2 not only reduced the energy barrier of the reaction, but also changed the product channels, i.e., formation of NF2O + (H2O)2-HF instead of NF2OH + (H2O)2-F. Therefore, the removal of NF3 by reaction with OH is possible in the presence of water molecules. The results presented in this study should provide useful information on the atmospheric chemistry of NF3.

1. Introduction

As the most extensively used perfluoro compound, nitrogen trifluoride (NF3) has attracted great interest in recent years. NF3 is commonly used in the semiconductor industry [1,2,3] and as a fluorine-supplying source in the electronic industry [4,5]. The industrial use of NF3 had been considered safe for a long time, as it did not produce carbon contamination residues. Hence, the production of NF3 as a substitute for other perfluorinated gases such as CF4 and C2F6 increased dramatically in recent years [6], resulting in a very large amount of NF3 atmospheric emissions. Unfortunately, recent studies have warned that there is a clear risk in using NF3 [7]. Firstly, NF3 is considered a new greenhouse gas, although it is not included in the list of greenhouse gases in the Kyoto protocol [8,9]. In fact, NF3 has a global warming potential (GWP) of 17,200, which is 10,800 times greater than that of CO2 when compared over a 100-year period [10,11,12]. Furthermore, NF3 and its decomposition products have been proposed to be toxic and pose a health risk [8]. Because of these concerns, great interest has focused on developing new processes to destroy or remove unreacted effluent NF3.
To date, various methods have been reported for the adsorption and decomposition of NF3 [13,14,15,16,17,18]. However, these methods are designed to deal with the tail gas in the semiconductor industry and electronic industry. On the other hand, there is a significant shortage of research on removal and decomposition processes for NF3 in the atmosphere. Gargano et al. [19] studied two important reactions involved in the decomposition of NF3, i.e., NF3 + F and N2 + F. Later, Cunha and coworkers [20] investigated other reactions involved in the decomposition of NF3 employing theoretical calculations at the CCSD(T)/cc-pVTZ level of theory, for example, NF2 + N, NF3 + NF, and the dissociation of N2F4 and N2F3. These pioneer studies provide fundamental insight into the mechanism of NF3 decomposition.
Several studies have also focused on the removal of NF3 in the atmosphere through the reaction with atmospheric oxidants. Wine and coworkers [21] studied the reaction of NF3 with O(1D), measured the rate coefficient to be k(T) = 2.0 × 10−11 exp(52/T) cm3 molecule−1s−1, and suggested that the reaction with O(1D) is an important atmospheric sink for NF3. Baasandorj and coworkers [22] also measured the rate coefficient of O(1D) with NF3, which is in good agreement with the results of Wine and coworkers. However, the reaction of reactive OH radical with NF3 was not mentioned. Dillon and coworkers [23] explored the possibility of removing NF3 by reactions with the atmospheric oxidants O(1D), OH and O3, and the results showed that the reaction rate of NF3 + OH is as slow as 2.0 × 10−29 cm3 molecule−1s−1; thus, they concluded that OH could not play an important role in atmospheric NF3 degradation. Although the reaction of OH with NF3 is extremely slow, the possibility of removing NF3 by reaction with OH should not be excluded because water molecules in the atmosphere have been shown to have a significant chemical catalytic effect on certain atmospheric reactions. Buszek and coworkers [24] reviewed the effect of water molecules on various atmospheric reactions, including radical–molecule, radical–radical, molecule–molecule and unimolecular reactions. It is surprising that, to our best knowledge, the influence of water molecules on the reaction of OH + NF3 has not been explored yet, though the reaction of OH with various molecules, such as HCOOH [25], HNO3 [26],CH3CHO [27,28,29], fluoroalcohols [30], HOCl [31], glyoxal [32,33], CH4 [34], DMSO [35], CH3OH [36,37,38], etc., have been studied extensively. Could additional water molecules accelerate the reaction of OH + NF3? If the answer is positive, how do the additional water molecules affect the reaction? Here, we decided to study the reaction of OH + NF3 with the participation of water molecules by using computational methods. These questions are critical for exploring the processes of removing NF3 in the atmosphere.

2. Results and Discussion

In principle, it is better to carry out benchmark calculations aiming to assess the accuracy of the DFT methods. Fortunately, several references have proved that the ωB97XD functional including dispersion was capable of treating various reactions [39,40]. As a result, all the discussions are based on the data obtained from CBS-QB3//ωB97XD/aug-cc-pVTZ methods. Moreover, as the reactions discussed here are in the gas phase, the electronic energy with zero-point energy was employed to discuss the thermodynamics [41,42].

2.1. The Reaction of NF3 + OH

According to a previous work [22], the reaction of NF3 + OH can be carried out through three distinct processes, i.e., the SN2 mechanism, F abstraction and H addition to the N center. However, the energy barrier is too high for H addition to N to be of consideration. As a result, only the SN2 mechanism and F abstraction were discussed here. As for the SN2 mechanism, an OH radical attacks the N, while one N-F bond is broken simultaneously, forming NF2OH and F. As shown in Figure 1, the corresponding transition state is TS1 with an energy barrier of 16.04 kcal/mol. Unfortunately, the product is endothermic by 2.50 kcal/mol, indicating this process is unfavorable thermodynamically, especially in atmospheric conditions. In the case of the F abstraction mechanism, an OH radical abstracts F from NF3 directly, leading to the production of HFO and NF2. The transition state for this process is TS2, in which the OH interacts with the leaving F. As can be seen in Figure 1, the energy barrier of TS2 (32.72 kcal/mol) is much higher than that of TS1. Moreover, the product is endothermic by as much as 10.02 kcal/mol. These results indicate the F abstraction process is unfeasible. In a word, although the SN2 mechanism is predominant in comparison to the F abstraction mechanism, the reaction of NF3 + OH is difficult to accomplish in view of thermodynamics. This is in good accordance with the extremely slow reaction rate measured experimentally [23]. As a result, the removal of NF3 through the gas-phase reaction with OH radial is of minor importance in atmospheric conditions. These results are in accordance with a previous report [23].

2.2. The Influence of Water Molecules on the Reaction of NF3 + OH

Inspired by the fact that the participation of water molecules could affect various atmospheric reactions in the gas phase [43,44,45,46,47,48,49], the influence of water molecules on the reaction of NF3 + OH is discussed here. In the condition where one H2O participates in the reaction, the NF3 + OH reaction takes place through two distinct process similar to the naked reaction shown in Figure 2. The corresponding geometry structures of all intermediates and transition states are available in Figure S1.
In contrast to the naked reaction, a pre-reactive complex formed in the entrance of the reaction due to the existence of a hydrogen bond between H2O and the reactants. For example, W1-RC1 and W1-RC2 are the pre-reactive complexes for the SN2 mechanism and F abstraction mechanism, respectively, as shown in Figure 2. Attributed to the formation of a hydrogen bond, W1-RC1 and W1-RC2 are 7.12 kcal/mol and 7.32 kcal/mol more stable than the reactants, respectively. Starting from W1-RC1, the reaction takes place via the SN2 mechanism. The corresponding transition state is W1-TS1. It should be noted that the structure of W1-TS1 is similar to that of TS1, except the breaking F atom bonded to the H2O due to the formation of an F∙∙∙H∙∙∙O hydrogen bond. The energy barrier of W1-TS1 is 14.64 kcal/mol, which is only 1.40 kcal/mol lower than that of TS1. Therefore, the influence of one H2O molecule on the SN2 mechanism is negligible in view of the kinetics. On the other hand, although the product complex W1-PC1 was located 18.84 kcal/mol below the reactants owing to the formation of a hydrogen bond, the final product, NF2OH + H2O-F, is endothermic by 1.90 kcal/mol, which is similar to that of PC1. It is reasonable to conclude that an additional H2O molecule is of no influence on the SN2 mechanism in view of the thermodynamics as well. As for the F abstraction mechanism, the corresponding transition state is W1-TS2, with a geometry structure similar to that of TS2. Unfortunately, the relative energy of W1-TS2 is as high as 26.30 kcal/mol. As a result, the reaction should overcome the energy barrier of 33.62 kcal/mol, which is even about 1 kcal/mol larger than that of F abstraction in the absence of H2O (TS2). Thus the participation of one H2O molecule is unable to accelerate the F abstraction process. In a word, when one additional H2O molecule takes part in the reaction of NF3 + OH, neither the kinetics nor thermodynamics are affected. This can be explained by the structure of the transition states. As can be seen in Figure S1, the hydrogen transfer process is not involved in either the SN2 mechanism or F abstraction mechanism. The additional H2O molecule only connects OH and F with the formation of a hydrogen bond rather than assisting the hydrogen transfer. Thus an additional, single H2O molecule only acts as a spectator rather than catalyst in the reaction of NF3 + OH. It is not unexpected that the effects of H2O molecules on various atmospheric reactions reported in the references do not appear here.

2.3. The Influence of an Additional Two H2O Molecules on the NF3 + OH Reaction

The structures of various pre-reactive complexes and transition states for the NF3 + OH reaction with an additional two H2O are shown in Figure 3. In contrast to the reaction with one participating H2O, there are three possible processes, as can be seen from the corresponding potential energy profiles (see Figure 4).
For the SN2 mechanism, the structure of the pre-reactive complex W2-RC1 is extremely similar to that of W1-RC1. However, the F-H bond in W2-RC1 is 0.3 Å shorter than that of W1-RC1, and the distance of N-O is reduced by about 0.2 Å. This could be attributable to the stabilization energy provided by the hydrogen bond of two H2O molecules, and could be proved by the energy of W2-RC1, which is 8 kcal/mol more stable than that of W1-RC1. The transition state corresponding to the broken N-F bond and N-O bond formation is W2-TS1, with an energy barrier of 12.17 kcal/mol. Moreover, W2-TS1 is located 3.06 kcal/mol below the initial reactants; thus, this transformation is accessible kinetically. It should be attributed to the direct participation of (H2O)2 in the reaction as a proton shuttle. As shown in Figure 3, the F-H in the W2-TS1 bond has shrunk to 1.89 Å, which is 0.32 Å shorter than that of W1-TS1, indicating the eliminated F has connected to the H2O molecules. This is verified by the intrinsic reaction coordinate (IRC) [50] calculation (see Figure S2). Moreover, the O-H bond in the OH radical is intended to break in the product direction of the IRC calculation, suggesting the products should change compared with the naked reaction and the reactions with one additional participating H2O. In fact, owing to the direct participation of (H2O)2, the broken H migrates along the (H2O)2 skeleton, resulting in the formation of NF2O + (H2O)2-HF, as exhibited in Figure 4, which is different from the SN2 mechanism of the NF3 + OH reaction with one additional H2O. It is interesting that the formation of NF2O + (H2O)2-HF is exothermic by 70.49 kcal/mol. In a word, when two additional H2O molecules take part in the reaction of NF3 + OH as catalyst, the formation of the products NF2O + (H2O)2-HF is favorable both thermodynamically and kinetically.
Considering the F abstraction mechanism, a pre-reactive complex, W2-RC2, was confirmed as well. Due to the hydrogen bond, W2-RC2 is 17.70 kcal/mol lower than the reactants, which is of marginal difference with the naked reaction and the reaction with one participating H2O molecule. The F abstraction was accomplished through W2-TS2, which is similar to W1-TS2 as well. However, the energy barrier of W2-TS2 is almost the same as that of TS2 and W1-TS2, inferring that (H2O)2 has marginal influence on the F abstraction mechanism. This result is not unexpected because the (H2O)2 plays the role of spectator, as can be seen from the structure of W2-TS2.
Apart from the SN2 mechanism and F abstraction mechanism, there is a new reaction process in the case of (H2O)2 participating in the NF3 + OH reaction, as depicted in Figure 4. This process initiates by the formation of W2-RC3, which is a pre-reactive complex formed by the contact between (H2O)2-OH and NF3. Starting from W2-RC3, the reaction proceeds via the transition state of W2-TS3, in which the change in O-N and F-N bonds is similar to that in W2-TS1. It is interesting that the IRC calculation of W2-TS3 bears evidence of the interaction between the substituted F and (H2O)2, leading to the formation of complex NF2O-(H2O)2-HF (see W2-PC3 in Figure 4). It is worth noting that W2-TS3 lies 0.53 kcal/mol below the reactants, suggesting this process in kinetically favorable as well.
It is well-known that the concentrations of larger complexes involving more than two molecules are very low in the troposphere [37]; as a result, only an additional one or two H2O molecules were taken into account. In summary, it is obvious that the participation of additional H2O molecules influences the reaction of NF3 with OH dramatically. Taking the SN2 mechanism, for example (see Figure 5), without the assistance of additional H2O molecules, the reaction is difficult to accomplish, as the energy barrier is high, and the products are endothermic. Fortunately, the energy barrier of the SN2 mechanism decreases by 8.5 kcal/mol and 10.6 kcal/mol in the case of H2O and (H2O)2 catalyzed reactions. Especially, the thermodynamics of the reaction change as the products change from NF2OH + F to NF2O + HF with the formation of an O-H∙∙∙F hydrogen bond.

3. Materials and Methods

Computational Methods

All reactants, products, pre-reactive complexes (RC, PC) and transition states (TS) were fully optimized using the density functional theory at the ωB97XD/aug-cc-pVTZ level of theory, as the long-range correction functional ωB97XD described the hydrogen bond well [51,52,53,54,55]. The harmonic vibrational frequencies of all optimized structures were calculated at the same level of theory to confirm the stationary point (intermediate or transition states) and for the zero-point energy (ZPE) corrections. The intrinsic reaction coordinate (IRC) [50] calculations were carried out to verify that the predicted transition states connect the designated reactants and products. In order to obtain more accurate thermodynamics data, the single-point energies of all species were calculated using the CBS-QB3 method [56,57]. The energy calculated at the CBS-QB3 level of theory was employed in the following discussion. All the DFT calculations were performed using the Gaussian 09 program [58]. The bond length comparison of selected species and all the optimized cartesian coordinates of species involved in the reactions are available in the Supporting Information (SI). The zero-point energy (ZPE) and relative energies are listed in Table 1. The energy profiles and corresponding structures for the reaction of NF3 with OH with the assistance of water molecules are illustrated in Figure 1, Figure 2, Figure 3, Figure 4 and Figure 5.

4. Conclusions

The possibility of removal of NF3 by the NF3 + OH reaction was studied at the CBS-QB3 level of theory. It was found that the NF3 + OH reaction in the absence of H2O molecules (naked reaction) is of no importance for the removal of NF3 in atmospheric conditions, as both the SN2 and F abstraction mechanisms must overcome a high energy barrier, while the products are endothermic. Although the participation of one H2O molecule has no influence on the NF3 + OH reaction, as the H2O acts as a spectator, it significantly changes when (H2O)2 takes part in the reaction as catalyst. The presence of (H2O)2 not only reduces the energy barrier of the SN2 mechanism, but also changes the products, i.e., with the formation of NF2O + (H2O)2-HF instead of NF2OH + (H2O)2-F. The reaction of NF3 + OH is favorable in the presence of (H2O)2, both kinetically and thermodynamically. The results indicate that it is possible to remove NF3 by reaction with OH radical in the presences of water molecules.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules29174033/s1, Figure S1: The optimized geometries of all species in the reaction of NF3 + OH with additional water molecule. Figure S2: IRC results for W2-TS1. Figure S3: IRC results for W2-TS3. Figure S4: Bond length comparison of selected species optimized at the level of ωB97XD/aug-cc-pVTZ (black) and MP2/aug-cc-pVTZ (red) methods. Optimized Cartesian coordinates of intermediates and transition states involved in Figure 1, Figure 2, Figure 3, Figure 4 and Figure 5.

Author Contributions

Conceptualization, X.L. (Xiaobing Lan) and Y.S.; methodology, X.L. (Xiaobing Lan) and Y.S.; validation, D.L. and X.Z.; formal analysis, J.C. and B.D.; investigation, J.L. and Y.Z.; resources, X.L. (Xiaobing Lan); data curation, J.L. and X.L. (Xueqi Lian); writing—original draft preparation, J.L. and Y.Z.; writing—review and editing, X.L. (Xiaobing Lan); visualization, Y.Z. and X.L. (Xueqi Lian); supervision, X.L. (Xiaobing Lan); project administration, X.L. (Xiaobing Lan) and Y.S.; funding acquisition, X.L. (Xiaobing Lan) and Y.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Outstanding Youth Project of Hunan Education Department (23B0777, 21B0750); and the Professorial and Doctoral Scientific Research Foundation of Huizhou University (No.2020JB046, No.2022JB009). Open Project Program of Guangdong Provincial Key Laboratory of Electronic Functional Materials and Devices, Huizhou University (No. EFMDN2021004M).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

We thank the people from the Hunan Provincial Key Laboratory of Xiangnan Rare-Precious Metal Compound Research and Application, School of Chemistry and Environmental Science, Xiangnan University.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tseng, Y.-H.; Tsui, B.-Y. Microtrenching-free two-step reactive ion etching of 4H-SiC using NF3/HBr/O2 and Cl2/O2. J. Vac. Sci. Technol. A 2014, 32, 031601. [Google Scholar] [CrossRef]
  2. Tasaka, A.; Takahashi, K.; Tanaka, K.; Shimizu, K.; Mori, K.; Tada, S.; Shimizu, W.; Abe, T.; Inaba, M.; Ogumi, Z. Plasma etching of SiC surface using NF3. J. Vac. Sci. Technol. A 2002, 20, 1254–1260. [Google Scholar] [CrossRef]
  3. Kastenmeier, B.; Oehrlein, G.; Langan, J.G.; Entley, W.R. Gas utilization in remote plasma cleaning and stripping applications. J. Vac. Sci. Technol. A 2000, 18, 2102–2107. [Google Scholar] [CrossRef]
  4. Donnelly, V.M. Review Article: Reactions of fluorine atoms with silicon, revisited, again. J. Vac. Sci. Technol. A 2017, 35, 05C202. [Google Scholar] [CrossRef]
  5. Tasaka, A. Electrochemical synthesis and application of NF3. J. Fluorine Chem. 2007, 128, 296–310. [Google Scholar] [CrossRef]
  6. Illuzzi, F.; Thewissen, H. Perfluorocompounds emission reduction by the semiconductor industry. J. Integr. Environ. Sci. 2010, 7, 201–210. [Google Scholar] [CrossRef]
  7. Prather, M.J.; Hsu, J. NF3, the greenhouse gas missing from Kyoto. Geophys. Res. Lett. 2008, 35, L12810. [Google Scholar] [CrossRef]
  8. Tsai, W.-T. Environmental and health risk analysis of nitrogen trifluoride (NF3), a toxic and potent greenhouse gas. J. Hazard. Mater. 2008, 159, 257–263. [Google Scholar] [CrossRef]
  9. Molina, L.T.; Wooldridge, P.J.; Molina, M. Atmospheric reactions and ultraviolet and infrared absorptivities of nitrogen trifluoride. Geophys. Res. Lett. 1995, 22, 1873–1876. [Google Scholar] [CrossRef]
  10. Totterdill, A.; Kovács, T.; Feng, W.; Dhomse, S.; Smith, C.J.; Gómez-Martín, J.C.; Chipperfield, M.P.; Forster, P.M.; Plane, J. Atmospheric lifetimes, infrared absorption spectra, radiative forcings and global warming potentials of NF3 and CF3 CF2 Cl (CFC-115). Atmos. Chem. Phys. 2016, 16, 11451–11463. [Google Scholar] [CrossRef]
  11. Rogelj, J.; McCollum, D.L.; O’Neill, B.C.; Riahi, K. 2020 emissions levels required to limit warming to below 2 °C. Nat. Clim. Change 2013, 3, 405–412. [Google Scholar] [CrossRef]
  12. Arnold, T.; Harth, C.M.; Mühle, J.; Manning, A.J.; Salameh, P.K.; Kim, J.; Ivy, D.J.; Steele, L.P.; Petrenko, V.V.; Severinghaus, J.P. Nitrogen trifluoride global emissions estimated from updated atmospheric measurements. Proc. Natl. Acad. Sci. USA 2013, 110, 2029–2034. [Google Scholar] [CrossRef]
  13. Ji, J.; Xiong, W.; Zhang, X.; Peng, L.; Shi, M.; Wu, Y.; Hu, X. Reversible absorption of NF3 with high solubility in Lewis acidic ionic liquids. Chem. Eng. J. 2022, 440, 135902. [Google Scholar] [CrossRef]
  14. Gao, Q.; Wang, Y.; Pan, Y.; Li, Y.; Sui, Z.; Xu, X. NF3 decomposition over V2O5, Fe2O3 and Co3O4 coated-Al2O3 reagents: The effect of promoter loadings on reactivity. J. Environ. Chem. Eng. 2020, 8, 103890. [Google Scholar] [CrossRef]
  15. Gao, Q.; Liu, Z.; Li, Y.; Sui, Z.; Liao, W.; Xu, X. NF3 decomposition without water over Cr2O3 coated-Al2O3 reagents. J. Environ. Chem. Eng. 2020, 8, 104166. [Google Scholar] [CrossRef]
  16. Xu, X.; Gao, Q.; Yin, C.; Pan, Y. NF3 decomposition in the absence of water over some metal oxides coated-Al2O3 reagents. J. Environ. Chem. Eng. 2019, 7, 103192. [Google Scholar] [CrossRef]
  17. Lai, S.Y.; Pan, W.; Ng, C.F. Catalytic hydrolysis of dichlorodifluoromethane (CFC-12) on unpromoted and sulfate promoted TiO2–ZrO2 mixed oxide catalysts. Appl. Catal. B 2000, 24, 207–217. [Google Scholar] [CrossRef]
  18. Karmakar, S.; Greene, H.L. Oxidative destruction of chlorofluorocarbons (CFC11 and CFC12) by zeolite catalysts. J. Catal. 1992, 138, 364–376. [Google Scholar] [CrossRef]
  19. Claudino, D.; Gargano, R.; Carvalho-Silva, V.H.; e Silva, G.M.; Da Cunha, W. Investigation of the abstraction and dissociation mechanism in the nitrogen trifluoride channels: Combined post-Hartree–Fock and Transition State Theory approaches. J. Phys. Chem. A 2016, 120, 5464–5473. [Google Scholar] [CrossRef]
  20. Ramalho, S.S.; da Cunha, W.F.; Albernaz, A.F.; Neto, P.H.; e Silva, G.M.; Gargano, R. An extensive investigation of reactions involved in the nitrogen trifluoride dissociation. New J. Chem. 2013, 37, 3244–3251. [Google Scholar] [CrossRef]
  21. Zhao, Z.; Laine, P.L.; Nicovich, J.M.; Wine, P.H. Reactive and nonreactive quenching of O (1D) by the potent greenhouse gases SO2F2, NF3, and SF5CF3. Proc. Natl. Acad. Sci. USA 2010, 107, 6610–6615. [Google Scholar] [CrossRef] [PubMed]
  22. Baasandorj, M.; Hall, B.; Burkholder, J. Rate coefficients for the reaction of O (1D) with the atmospherically long-lived greenhouse gases NF3, SF5 CF3, CHF3, C2F6, C4 F8, n-C5F12, and n-C6F14. Atmos. Chem. Phys. 2012, 12, 11753–11764. [Google Scholar] [CrossRef]
  23. Dillon, T.J.; Vereecken, L.; Horowitz, A.; Khamaganov, V.; Crowley, J.N.; Lelieveld, J. Removal of the potent greenhouse gas NF3 by reactions with the atmospheric oxidants O (1D), OH and O3. Phys. Chem. Chem. Phys. 2011, 13, 18600–18608. [Google Scholar] [CrossRef] [PubMed]
  24. Buszek, R.J.; Francisco, J.S.; Anglada, J.M. Water effects on atmospheric reactions. Int. Rev. Phys. Chem. 2011, 30, 335–369. [Google Scholar] [CrossRef]
  25. Anglada, J.M.; Gonzalez, J. Different catalytic effects of a single water molecule: The gas-phase reaction of formic acid with hydroxyl radical in water vapor. ChemPhysChem 2009, 10, 3034–3045. [Google Scholar] [CrossRef]
  26. Gonzalez, J.; Anglada, J.M. Gas phase reaction of nitric acid with hydroxyl radical without and with water. A theoretical investigation. J. Phys. Chem. A 2010, 114, 9151–9162. [Google Scholar] [CrossRef]
  27. Neeman, E.; González, D.; Blázquez, S.; Ballesteros, B.; Canosa, A.; Antiñolo, M.; Vereecken, L.; Albaladejo, J.; Jiménez, E. The impact of water vapor on the OH reactivity toward CH3CHO at ultra-low temperatures (21.7–135.0 K): Experiments and theory. J. Chem. Phys. 2021, 155, 034306. [Google Scholar] [CrossRef]
  28. Iuga, C.; Alvarez-Idaboy, J.R.; Reyes, L.; Vivier-Bunge, A. Can a single water molecule really catalyze the acetaldehyde+ OH reaction in tropospheric conditions? J. Phys. Chem. Lett. 2010, 1, 3112–3115. [Google Scholar] [CrossRef]
  29. Vohringer-Martinez, E.; Hansmann, B.; Hernandez, H.; Francisco, J.; Troe, J.; Abel, B. Water catalysis of a radical-molecule gas-phase reaction. Science 2007, 315, 497–501. [Google Scholar] [CrossRef]
  30. Bai, F.-Y.; Deng, M.-S.; Chen, M.-Y.; Kong, L.; Ni, S.; Zhao, Z.; Pan, X.-M. Atmospheric oxidation of fluoroalcohols initiated by ˙OH radicals in the presence of water and mineral dusts: Mechanism, kinetics, and risk assessment. Phys. Chem. Chem. Phys. 2021, 23, 13115–13127. [Google Scholar] [CrossRef]
  31. Gonzalez, J.; Anglada, J.M.; Buszek, R.J.; Francisco, J.S. Impact of water on the OH+ HOCl reaction. J. Am. Chem. Soc. 2011, 133, 3345–3353. [Google Scholar] [CrossRef] [PubMed]
  32. Long, B.; Tan, X.; Ren, D.; Zhang, W. Theoretical study on the water-catalyzed reaction of glyoxal with OH radical. J. Mol. Struct. THEOCHEM 2010, 956, 44–49. [Google Scholar] [CrossRef]
  33. Iuga, C.; Alvarez-Idaboy, J.R.; Vivier-Bunge, A. Single water-molecule catalysis in the glyoxal+ OH reaction under tropospheric conditions: Fact or fiction? A quantum chemistry and pseudo-second order computational kinetic study. Chem. Phys. Lett. 2010, 501, 11–15. [Google Scholar] [CrossRef]
  34. Allodi, M.A.; Dunn, M.E.; Livada, J.; Kirschner, K.N.; Shields, G.C. Do hydroxyl radical− water clusters, OH (H2O) n, n= 1− 5, exist in the atmosphere? J. Phys. Chem. A 2006, 110, 13283–13289. [Google Scholar] [CrossRef] [PubMed]
  35. Jørgensen, S.; Kjaergaard, H.G. Effect of Hydration on the Hydrogen Abstraction Reaction by HO in DMS and its Oxidation Products. J. Phys. Chem. A 2010, 114, 4857–4863. [Google Scholar] [CrossRef]
  36. Ali, M.A.; Balaganesh, M.; Al-Odail, F.A.; Lin, K. Effect of ammonia and water molecule on OH+ CH3OH reaction under tropospheric condition. Sci. Rep. 2021, 11, 12185. [Google Scholar] [CrossRef]
  37. Wu, J.; Gao, L.G.; Varga, Z.; Xu, X.; Ren, W.; Truhlar, D.G. Water catalysis of the reaction of methanol with OH radical in the atmosphere is negligible. Angew. Chem. Int. Ed. 2020, 132, 10918–10922. [Google Scholar] [CrossRef]
  38. Jara-Toro, R.A.; Hernández, F.J.; Taccone, R.A.; Lane, S.I.; Pino, G.A. Water Catalysis of the Reaction between Methanol and OH at 294 K and the Atmospheric Implications. Angew. Chem. Int. Ed. 2017, 56, 2166–2170. [Google Scholar] [CrossRef]
  39. Schenker, S.; Schneider, C.; Tsogoeva, S.B.; Clark, T. Assessment of popular DFT and semiempirical molecular orbital techniques for calculating relative transition state energies and kinetic product distributions in enantioselective organocatalytic reactions. J. Chem. Theory Comput. 2011, 7, 3586–3595. [Google Scholar] [CrossRef] [PubMed]
  40. Muniz-Miranda, F.; Occhi, L.; Fontanive, F.; Menziani, M.C.; Pedone, A. Quantum-Chemistry Study of the Hydrolysis Reaction Profile in Borate Networks: A Benchmark. Molecules 2024, 29, 1227. [Google Scholar] [CrossRef]
  41. Zhang, M.; Hou, H.; Wang, B. Theoretical Study on the Mechanisms and Kinetics of Atmospheric Oxidation of Tetrafluoropropyne and Its Analogues. J. Phys. Chem. A 2024, 128, 1511–1522. [Google Scholar] [CrossRef] [PubMed]
  42. Zhang, M.; Tian, Y.; Hou, H.; Wang, B. Mechanistic and kinetic study of the oxidation of trifluoroacetonitrile by hydroxyl and oxygen. Int. J. Quantum Chem. 2023, 123, e27147. [Google Scholar] [CrossRef]
  43. Zhao, X.; Liu, Z.; Zhao, R.; Xu, T. The effect of (H2O) n (n= 1–3) clusters on the reaction of HONO with HCl: A mechanistic and kinetic study. Phys. Chem. Chem. Phys. 2022, 24, 10011–10024. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, Y.; Zhao, M.; Liu, Y.; Sun, Y. The influence of a single water molecule on the reaction of BrO+ HONO. J. Mol. Graph. Model. 2022, 116, 108261. [Google Scholar] [CrossRef]
  45. Zhao, C.; Ma, X.; Wu, X.; Thomsen, D.L.; Bierbaum, V.M.; Xie, J. Single solvent molecules induce dual nucleophiles in gas-phase ion–molecule nucleophilic substitution reactions. J. Phys. Chem. Lett. 2021, 12, 7134–7139. [Google Scholar] [CrossRef]
  46. Zhang, T.; Zhai, K.; Zhang, Y.; Geng, L.; Geng, Z.; Zhou, M.; Lu, Y.; Shao, X.; Lily, M. Effect of water and ammonia on the HO+ NH3→ NH2+ H2O reaction in troposphere: Competition between single and double hydrogen atom transfer pathways. Comput. Theor. Chem. 2020, 1176, 112747. [Google Scholar] [CrossRef]
  47. Kumar, A.; Mallick, S.; Kumar, P. Effect of water on the oxidation of CO by a Criegee intermediate. Phys. Chem. Chem. Phys. 2020, 22, 21257–21266. [Google Scholar] [CrossRef]
  48. Zhang, T.; Wen, M.; Zhang, Y.; Lan, X.; Long, B.; Wang, R.; Yu, X.; Zhao, C.; Wang, W. Atmospheric chemistry of the self-reaction of HO2 radicals: Stepwise mechanism versus one-step process in the presence of (H2O) n (n= 1–3) clusters. Phys. Chem. Chem. Phys. 2019, 21, 24042–24053. [Google Scholar] [CrossRef]
  49. Chao, W.; Yin, C.; Takahashi, K. Effects of water vapor on the reaction of CH2OO with NH3. Phys. Chem. Chem. Phys. 2019, 21, 22589–22597. [Google Scholar] [CrossRef]
  50. Gonzalez, C.; Schlegel, H.B. An improved algorithm for reaction path following. J. Chem. Phys. 1989, 90, 2154–2161. [Google Scholar] [CrossRef]
  51. Alipour, M.; Fallahzadeh, P. First principles optimally tuned range-separated density functional theory for prediction of phosphorus–hydrogen spin–spin coupling constants. Phys. Chem. Chem. Phys. 2016, 18, 18431–18440. [Google Scholar] [CrossRef] [PubMed]
  52. Chai, J.-D.; Head-Gordon, M. Systematic optimization of long-range corrected hybrid density functionals. J. Chem. Phys. 2008, 128, 084106. [Google Scholar] [CrossRef] [PubMed]
  53. 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]
  54. Woon, D.E.; Dunning, T.H., Jr. Gaussian basis sets for use in correlated molecular calculations. III. The atoms aluminum through argon. J. Chem. Phys. 1993, 98, 1358–1371. [Google Scholar] [CrossRef]
  55. Kendall, R.A.; Dunning, T.H.; Harrison, R.J. Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796–6806. [Google Scholar] [CrossRef]
  56. Montgomery, J.A., Jr.; Frisch, M.J.; Ochterski, J.W.; Petersson, G.A. A complete basis set model chemistry. VI. Use of density functional geometries and frequencies. J. Chem. Phys. 1999, 110, 2822–2827. [Google Scholar] [CrossRef]
  57. Montgomery, J.; Frisch, M.J.; Ochterski, J.W.; Petersson, G.A. A complete basis set model chemistry. VII. Use of the minimum population localization method. J. Chem. Phys. 2000, 112, 6532–6542. [Google Scholar] [CrossRef]
  58. 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 09, Revision E.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
Figure 1. The energy profile of the NF3 + OH reaction.
Figure 1. The energy profile of the NF3 + OH reaction.
Molecules 29 04033 g001
Figure 2. Energy profiles of one additional H2O participating in the reaction of NF3 + OH.
Figure 2. Energy profiles of one additional H2O participating in the reaction of NF3 + OH.
Molecules 29 04033 g002
Figure 3. The structures of intermediates and transition states involved in NF3 + OH reactions with two H2O molecules participating. The distances are in Å.
Figure 3. The structures of intermediates and transition states involved in NF3 + OH reactions with two H2O molecules participating. The distances are in Å.
Molecules 29 04033 g003
Figure 4. The energy profiles of NF3 + OH reactions with an additional two H2O molecules.
Figure 4. The energy profiles of NF3 + OH reactions with an additional two H2O molecules.
Molecules 29 04033 g004
Figure 5. Comparison of naked NF3 + OH reaction and the H2O and (H2O)2 assisted reactions.
Figure 5. Comparison of naked NF3 + OH reaction and the H2O and (H2O)2 assisted reactions.
Molecules 29 04033 g005
Table 1. The ZPE and calculated relative energies (in kcal/mol) for all reactants, transition states and products.
Table 1. The ZPE and calculated relative energies (in kcal/mol) for all reactants, transition states and products.
SpeciesZPE (ωB97XD/avtz)ΔE + ZPE(ωB97XD/avtz)ΔE (CBS-QB3)
NF3+ OH + nH2O/00
TS113.7219.7716.04
TS212.6932.2432.72
W1-RC128.24−4.40−7.12
W1-RC228.21−4.32−7.32
W1-TS129.5813.877.52
W1-TS228.0027.7526.30
W1-PC131.23−18.94−18.84
W1-PC229.582.12−0.53
W2-RC144.22−9.60−15.23
W2-RC244.88−11.16−17.70
W2-RC344.86−11.28−17.87
W2-TS145.735.99−3.06
W2-TS244.5520.6915.78
W2-TS345.608.39−0.53
W2-PC145.98−66.14−72.93
W2-PC245.56−4.78−10.50
W2-PC346.30−66.01−73.08
NF2OH + F15.062.792.50
NF2 + HFO12.848.9510.02
NF2OH + H2O-F28.395.951.90
NF2 + H2O-HFO28.572.982.05
NF2O + (H2O)2-HF43.85−64.63−70.49
NF2 + (H2O)2-HFO44.99−3.90−7.92
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

Liu, J.; Zhao, Y.; Lian, X.; Li, D.; Zhang, X.; Chen, J.; Deng, B.; Lan, X.; Shao, Y. Unveiling the Influence of Water Molecules for NF3 Removal by the Reaction of NF3 with OH: A DFT Study. Molecules 2024, 29, 4033. https://doi.org/10.3390/molecules29174033

AMA Style

Liu J, Zhao Y, Lian X, Li D, Zhang X, Chen J, Deng B, Lan X, Shao Y. Unveiling the Influence of Water Molecules for NF3 Removal by the Reaction of NF3 with OH: A DFT Study. Molecules. 2024; 29(17):4033. https://doi.org/10.3390/molecules29174033

Chicago/Turabian Style

Liu, Jiaxin, Yong Zhao, Xueqi Lian, Dongdong Li, Xueling Zhang, Jun Chen, Bin Deng, Xiaobing Lan, and Youxiang Shao. 2024. "Unveiling the Influence of Water Molecules for NF3 Removal by the Reaction of NF3 with OH: A DFT Study" Molecules 29, no. 17: 4033. https://doi.org/10.3390/molecules29174033

Article Metrics

Back to TopTop