Next Article in Journal
Catalytic Performance of Bulk and Al2O3-Supported Molybdenum Oxide for the Production of Biodiesel from Oil with High Free Fatty Acids Content
Next Article in Special Issue
Zeolite Beta Doped with La, Fe, and Pd as a Hydrocarbon Trap
Previous Article in Journal
Oxygen Reduction Reaction Catalyzed by Pt3M (M = 3d Transition Metals) Supported on O-doped Graphene
Previous Article in Special Issue
Cu-Mg-Fe-O-(Ce) Complex Oxides as Catalysts of Selective Catalytic Oxidation of Ammonia to Dinitrogen (NH3-SCO)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Water Molecule on Photo-Assisted Nitrous Oxide Decomposition over Oxotitanium Porphyrin: A Theoretical Study

by
Phornphimon Maitarad
1,2,*,
Vinich Promarak
2,3,
Liyi Shi
1 and
Supawadee Namuangruk
1,4,*
1
Research Center of Nano Science and Technology, Shanghai University, Shanghai 200444, China
2
Department of Materials Science and Engineering, School of Molecular Science and Engineering, Vidyasirimedhi Institute of Science and Technology, Rayong 21210, Thailand
3
Research Network of NANOTEC-VISTEC on Nanotechnology for Energy, Vidyasirimedhi Institute of Science and Technology, Wangchan, Rayong 21210, Thailand
4
National Nanotechnology Center (NANOTEC), NSTDA, 111 Thailand Science Park, Pahonyothin Road, Klong Luang, Pathum Thani 12120, Thailand
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(2), 157; https://doi.org/10.3390/catal10020157
Submission received: 3 January 2020 / Revised: 21 January 2020 / Accepted: 21 January 2020 / Published: 1 February 2020
(This article belongs to the Special Issue Catalysis for the Removal of Gas-Phase Pollutants)

Abstract

:
Water vapor has generally been recognized as an inhibitor of catalysts in nitrous oxide (N2O) decomposition because it limits the lifetime of catalytic reactors. Oxygen produced in reactions also deactivates the catalytic performance of bulk surface catalysts. Herein, we propose a potential catalyst that is tolerant of water and oxygen in the process of N2O decomposition. By applying density functional theory calculations, we investigated the reaction mechanism of N2O decomposition into N2 and O2 catalyzed by oxotitanium(IV) porphyrin (TiO-por) with interfacially bonded water. The activation energies of reaction Path A and B are compared under thermal and photo-assisted conditions. The obtained calculation results show that the photo-assisted reaction in Path B is highly exothermic and proceeds smoothly with the low activation barrier of 27.57 kcal/mol at the rate determining step. The produced O2 is easily desorbed from the surface of the catalyst, requiring only 4.96 kcal/mol, indicating the suppression of catalyst deactivation. Therefore, TiO-por is theoretically proved to have the potential to be a desirable catalyst for N2O decomposition with photo-irradiation because of its low activation barrier, water resistance, and ease of regeneration.

Graphical Abstract

1. Introduction

Nitrous oxide (N2O) is an anthropogenic gas reported as the largest contribution to the ozone depleting gas emissions [1]. Once it is transported to the stratosphere, it destroys the ozone and causes the greenhouse effect. At present, the development of technologies for the abatement of N2O mainly focuses on (i) nonselective catalytic reduction, (ii) selective catalytic reduction, and (iii) direct catalytic decomposition. Among the up-to-date technologies, the direct decomposition of N2O (deN2O) has attracted much attention because of its efficiency and cost effectiveness [2,3,4,5]. Researches in the field of N2O catalytic decomposition mainly focus on low-temperature deN2O catalysts that can be applied to N2O abatement in medical operating rooms, nitric acid plants, three-way catalytic converters, etc. For example, N2O is used as an anesthetic gas in hospitals and it can lead to miscarriages in gestation, liver disorder, kidney trouble, and so on. Doi et al. [6,7] worked on the deN2O of the air contaminated with N2O in operating rooms by varying the alumina-supported metals (Pt, Pd, or Rh), and found that the Rh/Al2O3 is the most suitable catalyst for deN2O in operating rooms which could reach 100% decomposition at 773K.
The catalytic deN2O has been studied to a greater extent than selective catalytic reduction (SCR) over a large variety of catalysts, including noble metals [4,8,9,10,11], perovskites [12,13,14], metal oxides [15,16,17,18], and zeolites [19,20,21,22]. It is also important to note that in deN2O processes, tolerance to various substances coexisting in the exhaust gases (e.g. NOx, SO2, O2, and H2O) should be simultaneously accomplished without sacrificing low-temperature activity. For example, the effect of SO2 and/or H2O on the NOx and N2O reduction was examined over the In/Al2O3-Ru/Al2O3 dual-bed reactor. Their transient experiments showed that the N2O conversion dropped to zero in the presence of SO2 and H2O [23,24]. In addition, there have been reports of various catalysts which show that their catalytic performance of deN2O is decreased due to the active site poisoned with water vapor or oxygen molecules produced during the reaction [25,26,27,28]. Regarding to water inhibition, it was studied, theoretically, by Heyden et al. [27] who concluded that water impurity (<100 ppm) may strongly affect the kinetics of N2O decomposition, leading to an increase in apparent activation energy for 26 kcal/mol approximately. It is, obviously, important to take into account both the reaction mechanism and activation energy. In addition, those who are working on development of new catalysts for deN2O should consider the effect of water at the active sites as well as the catalytic site regeneration by oxygen desorption, which is key to the reaction.
Besides thermal deN2O, photo-assisted direct decomposition is another alternative method for N2O removal [29,30,31]. This method applies light as the energy source for initiating the reaction, and the catalysis is active even at low gas concentration and low temperature. Therefore, with a suitable catalyst, a photocatalytic under the UV and near UV lights with wavelength 200–400 nm can be applied to decompose N2O. In general, N2O is decomposed into N2 and O2 under a photocatalytic reaction [32] by
N 2 O   h v ,     c a t a l y s t   N 2 +   1 2   O 2
The well-accepted photo-catalyst for N2O decomposition is TiO2 loaded with various noble metals [33,34,35,36]. However, noble metals are expensive for practical uses. Thus, a search for low-cost and effective catalysts for photocatalytic N2O decomposition has been of great interest.
Porphyrin is a potential compound because it possesses two axial coordination sites which are suitable for anchoring to a solid substrate, such as in the metal-porphyrin assembly that features in several useful applications, including photovoltaic materials, field-responsive materials, and catalytic materials [37,38,39,40]. In our previous theoretical work [41], we proposed that low-valent titanium(II)-porphyrin (Ti-por) is a good catalyst for deN2O under thermal conditions; the activation energies in the energy profiles are comparable to other potential catalysts. However, this low-valent Ti-por can be easily oxidized to form a more stable structure, reducing its suitability. The high-valent oxotitanium(IV)-porphyrin (TiO-por) is a better choice and has been reported as a potential catalyst for chemical and biological reactions [42,43,44]. Because of the structural combination of TiO and the electron-rich porphyrin ligand, it is of interest to extend the theoretical study to the photo-assisted deN2O. In this work, we theoretically investigated the reaction mechanism of deN2O with the TiO-por catalyst by using density functional theory (DFT), and included a water molecule at the active site, and studied O2 desorption as well.

2. Results and Discussion

The presence of water in the system might provide some steric hindrance around the active site and increase the activation energy barriers of the reaction. Hence, the effect of water was intensively studied in this work. Considering water molecule at the TiO site (Figure 1), we divided the N2O decomposition over the TiO-por into four elementary steps as follows.
Step 1: Water-complex formation on TiO-por
TiO-por + H2O → Ti(OH)2-por
Step 2: First N2O decomposition
Ti(OH)2-por + 1st N2O → TiO(OH)2-por + N2
Step 3-A: Second N2O decomposition followed by water desorption (Path A)
TiO(OH)2-por + 2nd N2O → Ti(OOH)2-por + N
Ti(OOH)2-por → TiO3-por + H2O
Step 3-B: Water desorption followed by second N2O decomposition (Path B)
TiO(OH)2-por → TiO2-por + H2O
TiO2-por + 2nd N2O → TiO3-por + N2
Step 4: Catalyst regeneration; oxygen formation and desorption
TiO3-por → TiO-por + O2
As mentioned earlier, the thermal and photo-assisted reactions were simulated by using the DFT calculations of singlet and triplet states, respectively. The calculated activation energies of all steps in the singlet and triplet states are listed in Table 1. Based on the pairwise comparison, it can be clearly seen that in the overall reaction processes of N2O direct decomposition in the photo-assisted condition, TiO-por (3TiO-por) catalyst gives a more favorable reaction route with lower activation barriers. Therefore, the details of each reaction step in the photo-assisted condition will be further discussed. For the photo-absorption of the system, the absorption spectra of TiO-por and N2O are compared in Figure S1. The first absorption band of TiO-por appears in the range of 300–400 nm which is assigned as the transition among the Gouterman’s four orbitals (Figure S2). On the other hand, the absorption of N2O appears at around 100–150 nm, in FUV region. Thus, the UV light irradiation selectively promotes TiO-por to the excited state which leads to the N2O decomposition reaction. The spin density plot of the 3TiO-por (Figure S3) also shows that the UV absorption is characterized as π–π* transition, with the spin density predominantly localized at meso-nitrogen positions.

2.1. Water Dissociation over the TiO-por

The energy profile of water dissociation over the TiO-por is shown in Figure 2 with optimized structures. First, a water molecule approaches the TiO site, in which Ti is slightly above the porphyrin plane and this water molecule undergoes adsorption with the TiO-por (AD1), resulting in an energy change of −7.57 kcal/mol. The optimized AD1 structure shows that a weak hydrogen bond is generated with a distance of 2.00 Å; the structural parameters of the adsorbed water are not significantly changed.
The first step of the reaction is the concerted hydrogen transfer and Ti–O bond formation, which was preceded by the O2–H1 bond breaking and O1–H1 bond formation, as seen in the structure of TS1. The activation energy barrier of this step is 19.61 kcal/mol. The H1 of the water molecule forms a bond with the O1 with a distance of 1.13 Å and an <O1–H1–O2 angle of 138.5°. Because of this O–H bond formation, the Ti–O1 bond is elongated to 1.75 Å at an angle to the porphyrin plane. This structure results in the O2 atom of the water molecule being closer to the Ti atom. The imaginary frequency at TS1 is 1284i cm−1. The O1–H1 bond formation follows, with the O1–H1 bond distance of 0.96 Å. In addition, the Ti–O2 bond forms with a bond distance of 1.89 Å (IN1). Therefore, the new active site is generated as Ti(OH)2 (IN1), from the water associated TiO-por catalyst. The relative energy of IN1 is −3.00 kcal/mol, which is slightly exothermic, but endothermic when being relative to AD1.

2.2. First N2O Decomposition

After water dissociation and formation of Ti(OH)2 (IN1), the first N2O decomposition proceeds. The reaction profile consists of four stationary points: N2O adsorption, first transition state, N2 production, and N2 desorption, as displayed in Figure 3. The molecule weakly interacts with the OH via a hydrogen bond with an adsorption energy of −4.94 kcal/mol (AD2). Then, it forms a Ti–O3 bond (2.28 Å) at the transition state (TS2), in which the N2O molecule shows a bent structure with an <O3–N1–N2 of 150°. The O3–N2 bond is elongated from 1.18 Å to 1.46 Å, indicating that the O3–N2 bond is activated. Because of this insertion, the Ti–O1 bond is elongated from 1.85 Å to 2.03 Å. This transition state has an activation energy barrier of 27.57 kcal/mol, with an imaginary frequency of 439i cm−1. The N–O bond breaks at the next intermediate (IN2), and a Ti–O3 bond is formed with a distance of 1.93 Å. Because of this Ti–O3 bond formation, the Ti–O1 bond becomes weak and the O1 approaches the O3 atom with a distance of 1.45 Å. The product N2 molecule is generated in this step. It requires 3.11 kcal/mol to desorb the N2 molecule from the catalyst intermediate (IN2-1). Consequently, in the first N2O decomposition step, the N2O molecule weakly adsorbs over the hydroxyl-oxotitanium porphyrin (Ti(OH)2-por) and the N2O decomposition needs an activation energy of 27.57 kcal/mol to generate the N2 molecule.

2.3. Second N2O Decomposition

Two possible routes, Path A and Path B, were examined for the second N2O decomposition. Path A is the case where the second N2O is decomposed before water desorption. Alternatively, a water molecule is desorbed before the second N2O decomposition in Path B.
Path A: The second N2O decomposition (Figure 4) starts from the IN2-1 (TiO(OH)2-por) whose active site is coordinated by three oxygen atoms. All the steps of this decomposition proceed similarly to the first N2O reaction. First, the AD3A intermediate is formed when a second N2O is adsorbed on top of the IN2-1; the terminal O atom of N2O interacts with the OH on the Ti center. The calculated adsorption energy is −3.77 kcal/mol. Through the transition state (TS3A) at an imaginary frequency of 608i cm−1, N–O bond dissociation occurs to form the intermediate IN3A and a weakly interacting N2 molecule. The activation energy barrier of this step is 29.01 kcal/mol, which is the same energy as the first N2O decomposition. The produced N2 desorbs from the active site, which requires 3.29 kcal/mol, and leaves the IN3-1A intermediate, which consists of two OH groups coordinated at the Ti center. The resultant H2O has a nearly linear structure with the <H1–O1–H2 angle being 171°. As a result, the H2O molecule is formed easily (IN4A) in an exothermic process, −20.34 kcal/mol relating to the IN3-1A intermediate. Finally, the water molecule is generated and desorbed with energy of 9.61 kcal/mol (IN4-1A).
For Path B, when considering the structure of IN2-1 (Figure 5) and the alternative pathway is also possible, namely, the H2O molecule is released at the beginning, since the O1 has a hydrogen bond with the H2 atom and the distance is 2.25 Å with ∠H1–O1–H2 of about 171°. Thus, Path B starts from H2O molecule desorption, which produces the intermediate IN3B. The formation of the water molecule is an exothermic step with a relative energy of −11.59 kcal/mol. After the water desorption (IN3-1B), the active site becomes the TiO2-por intermediate. Then, the second N2O interacts with the TiO2-por intermediate with an energy change of −5.10 kcal/mol (AD4B); this coordination is slightly stronger than the N2O adsorption in Path A. To decompose the N2O molecule over the TiO2-por through the transition state (TS4B) requires an activation energy of 12.37 kcal/mol, which is much lower than the case of Path A (TS3A). This TS4B has an imaginary frequency of 648i cm−1, and the N2O molecule has a bent structure with an ∠O3–N1–N2 angle of 126°. The O3 forms a Ti–O bond and the N1–O3 bond dissociates to give intermediate IN4B. Finally, the N2 molecule desorbs from IN4B, which generates TiO3-porphyrin (IN4-1B). The calculated desorption energy of the N2 molecule is 4.45 kcal/mol.
In general, the rate-determining step of N2O direct decomposition is desorption of the produced O2 from the catalysts. The O2 desorption from the active site usually requires high energy, which could prohibit reaction when the active site is deactivated [20,26,27,45,46]. However, our present work found that the O2 is formed over the TiO3 intermediate of IN4-1A and IN4-1B in both Path A and Path B, as shown in the last step of Figure 4 and Figure 5, respectively. It is worth noting that O2 molecule (IN5 intermediate) is easily formed with activation energy of 6.45 kcal/mol and barrierless for Path A and Path B, respectively. Furthermore, this O2 formation step is an extremely exothermic process. Therefore, it is shown that the TiO3-por catalyst intermediate easily regenerates the active TiO-por catalyst, which is an advantage of this catalyst, compared with the related zeolite catalysts for N2O decomposition [45,46].

2.4. Overall Energetics of Photo-Assisted N2O Decomposition over Hydroxyl-Oxotitanium Porphyrin Catalysts

The overall reaction energy profile of the water dissociation followed by N2O decomposition on the TiO-por is summarized in Figure 6. Initially, the water molecule is adsorbed on the TiO-por catalyst at −7.57 kcal/mol (AD1), and the active site becomes Ti(OH)2 (IN1). As regarding to the first N2O decomposition, the N2O molecule is slightly adsorbed on Ti(OH)2-por active site (AD2), requiring 27.57 kcal/mol (TS2) for the N–O dissociation to produce the N2 molecule. The active site then becomes the TiO(OH)2-por (IN2-1). Two pathways are possible starting from the IN2-1 intermediate. In Path A, H2O formation follows the second N2O adsorption. The activation energy in this pathway is 29.01 kcal/mol (TS3A). In Path B, the water molecule releases prior to N2O decomposition. In Path B, the second N2O decomposition occurs with an energy barrier of 12.37 kcal/mol (TS4B). Therefore, the second N2O decomposition over the TiO2-por in Path B is expected to be more preferable than over the TiO(OH)2-por of Path A, as summarized in Scheme 1. The final step is catalyst regeneration by oxygen molecule formation, and the calculated activation energy is barrierless to form the extremely exothermic intermediate, and the oxygen molecule easily desorbs from the surface of catalyst with only 4.96 kcal/mol. In the overall reaction, the N–O dissociation of the first N2O molecule in Step 2 is the rate-determining step, which correlates well with the N2O decomposition over Cu-ZSM5 [47]. It is worth to note that this present N2O direct decomposition reaction has total reaction energy approximately at −52.58 kcal/mol.

2.5. Effect of Water Molecule on the N2O Decomposition Barriers

On the pairwise comparison of the theoretical reaction mechanism for N2O decomposition with and without introducing H2O, only a few works have addressed this issue [26,27]. It is well known that Fe-ZSM5 zeolite is one of the potential catalysts for N2O direct decomposition and more importantly, theoretical analysis addressed the issue of activation barriers for the first and second N2O decompositions and oxygen molecule formation. Thus, the present oxotitanium porphyrin catalyst for deN2O with/without water molecule can be compared with the Fe-ZSM5 zeolite [27] to elucidate the catalytic potential and the effect of water. In Table 2, a comparison has been made between the Fe-ZSM5 zeolite catalysts with and without water in the N2O decomposition, and these are represented by Z[FeO]+ and Z[Fe(OH)2]+ active sites, respectively. It is clear that the Z[Fe(OH)2]+ active sites result in higher activation barriers to decompose the N2O and form oxygen. Comparing oxotitanium porphyrin with and without water molecule represented by 3Ti(OH)2-por and 3TiO-por, [41] respectively, the N2O decomposition over the 3Ti(OH)2-por results in similar activation barriers as the 3TiO-por active site [41]. In addition, these trends are also seen in the singlet state of oxotitanium porphyrin in this work. These results imply that the hydroxyl site of oxotitanium porphyrin produced by water dissociation does not increase the activation energy barrier for deN2O, whereas the Z[Fe(OH)2]+ catalyst shows a significant increase in the activation energy barriers for the first and second N2O decompositions and the desorption energy of the oxygen molecule.
For the photo-assisted reaction, which is the main target in this work, the presence of water molecule on the active site of oxotitanium porphyrin has been intensively examined, to get a thorough view of the catalytic activity for N2O decomposition, as displayed the overall reaction pathway in Scheme 1. In the absence of water [41], the first N2O directly decomposes into N2 product and 3TiO2-por intermediate forms. The first N2O adsorption on the TiO-por active site is −5.71 kcal/mol with activation energy barrier of 35.37 kcal/mol, it is worth mentioning that these energies do not include the zero-point energy (ZPE) correction [41]. In this reaction with presence of water, water molecule adsorption on the 3TiO-por site (−7.57 kcal/mol) is stronger than the N2O molecule. The water molecule more favorably covers the catalyst surface and generates the hydroxyl active site 3Ti(OH)2-por. Over 3Ti(OH)2-por active sites, the N2O adsorption energy is −4.94 kcal/mol and the calculated energy barrier to decompose the N2O is 27.57 kcal/mol. These results mean that the N2O decomposition is not inhibited by the presence of water at the active site 3Ti(OH)2-por. On the other hand, water molecules seem to act as an assisting reagent in the N2O decomposition, as it can easily desorb after the N2O decomposition process. Therefore, the TiO-por is proposed as a candidate catalyst for photo-assisted N2O decomposition even under aqueous condition.

3. Computational Details

Nitrous oxide decomposition on the TiO-por model system was considered in the ground and excited states. The reaction pathways were calculated with the DFT method using the M06-L functional [48,49], in which full geometry optimization was performed without any geometrical restriction. To simulate the lowest lying excited state potential energy surface which represents the reaction under photo-irradiation, we adopted the unrestricted DFT (UM06L) method for triplet state calculation. The unrestricted DFT method is occasionally used for the calculations of photo-catalytic reactions [41,50,51]. The M06L functional has been examined with some variants of the coupled clusters [52,53] and other DFT functionals for systems including metals [52,53,54]. Although the energy barrier is sometimes underestimated, it works well in comparison with other global hybrid functionals. A basis set of double zeta plus polarization, i.e., 6-31G(d,p) was adopted for C, O, N, and H atoms, and the relativistic effective core potential of Hay–Wadt (LANL2DZ) was used for the Ti atom. All calculations were carried out with the Gaussian09 suite of programs Rev. B01 [55].
For the deNOx simulation, H2O and N2O gases were used as the reactants. We considered both singlet and triplet spin states to simulate the reactions under thermal and photo-assisted conditions, respectively, the latter of which is assumed to be continuous photo-excitation with a considerable lifetime in the excited state without radiative or nonradiative decay. A vibrational analysis was done at each stationary point; all the intermediates were confirmed as true local minima and the transition states were first-order saddle points with only one imaginary frequency. The reaction energy profiles in terms of electronic with ZPE correction were calculated. The energy profile at each step was presented in the relative energy, which is defined as
Δ E = E c o m p l e x ( E c a t a l y s t + E a d s o r b a t e )
where E c o m p l e x , E c a t a l y s t , and E a d s o r b a t e are the total energies of the catalyst-adsorbate complex, the starting TiO-por complex at each step, and the isolated reactant molecules, respectively. The negative (positive) value indicates the stable (unstable) adsorption complex relative to the isolated systems.

4. Conclusions

The elementary reactions of N2O decomposition over the oxotitanium porphyrin under thermal and photo-assisted conditions in the presence of water have been examined using DFT method with the unrestricted M06L/6-31G(d, p) level of theory. The potential energy profiles show that the entire reaction is an exothermic process. The reaction steps of N2O decomposition over hydroxyl-oxotitanium porphyrin can be summarized as follows. For the first N2O decomposition, the N–O bond dissociation over the Ti(OH)2-por intermediate is the rate-determining step and requires an activation energy of 27.57 kcal/mol. Then, a spontaneous process of releasing the H2O from the TiO(OH)2-por catalyst intermediates is the preferable route, as examined in Path B, and the second N2O decomposition over the TiO2-por requires a lower activation energy of 12.37 kcal/mol than the first one. In the last step of catalyst regeneration, the O2 formation from the TiO3-por intermediate is a spontaneous process with a barrierless activation energy and a desorption energy of oxygen molecule only 4.96 kcal/mol, which is an advantage of this kind of catalyst.
Therefore, from this theoretical investigation, it is worth noting that the presence of water does not inhibit the N2O decomposition on the catalyst surface. As a final note, the exhaust gas contains various compounds such as H2O, O2, NxOy, CO2, and therefore, a theoretical study on the N2O decomposition reaction needs to consider the effect of other gases in the reaction mechanism as well. The comprehensive study of overall reaction mechanism, which involves all effects, will be useful to guide experimental catalyst development for deN2O applications.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/2/157/s1, Absorption spectrum of TiO-Por and N2O calculated by TD-M06L method, relevant MOs (a) HOMO-1, (b) HOMO, (c) LUMO, and (d) LUMO+1, of TiO-Por upon excitations, and the spin density plot of triplet state TiO-Por.

Author Contributions

P.M. (Conceptualization, Methodology, Data curation, Formal analysis, Writing—original draft and review), V.P. (Funding acquisition, Supervision), S.N. (Investigation, Methodology, Resources, Supervision, Review & editing) and L.S. (Funding acquisition, Supervision). All authors have read and agreed to the published version of the manuscript.

Funding

This research was partially funded by the NANOTEC, NSTDA, Ministry of Science and Technology, Thailand, through its program of Research Network NANOTEC and Thailand Research Fund (RSA6180080 and RTA6080005) and the Shanghai Municipal Science and Technology Commission of Professional and Technical Service Platform for Designing and Manufacturing of Advanced Composite Materials (16DZ2292100).

Acknowledgments

We thank Nanoscale Simulation Laboratory at National Nanotechnology Center (NANOTEC). P.M. would like to specially thank to Ehara for kindly discussion.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ravishankara, A.R.; Daniel, J.S.; Portmann, R.W. Nitrous Oxide (N2O): The Dominant Ozone-Depleting Substance Emitted in the 21st Century. Science 2009, 326, 123–125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kapteijn, F.; Rodriguez-Mirasol, J.; Moulijn, J.A. Heterogeneous catalytic decomposition of nitrous oxide. Appl. Catal. B Environ. 1996, 9, 25–64. [Google Scholar] [CrossRef]
  3. Konsolakis, M. Recent Advances on Nitrous Oxide (N2O) Decomposition over Non-Noble-Metal Oxide Catalysts: Catalytic Performance, Mechanistic Considerations, and Surface Chemistry Aspects. ACS Catal. 2015, 5, 6397–6421. [Google Scholar] [CrossRef]
  4. Liu, Z.; He, F.; Ma, L.; Peng, S. Recent Advances in Catalytic Decomposition of N2O on Noble Metal and Metal Oxide Catalysts. Catal. Surv. Asia 2016, 20, 121–132. [Google Scholar] [CrossRef]
  5. Leont’ev, A.V.; Fomicheva, O.A.; Proskurnina, M.V.; Zefirov, N.S. Modern chemistry of nitrous oxide. Russ. Chem. Rev. 2001, 70, 91–104. [Google Scholar] [CrossRef]
  6. Doi, K.; Noda, T.; Arai, T.L.N.; Tagawa, T.; Goto, S. Effects of Isoflurane on Kinetics of N2O Catalytic Decomposition over Rh/Al2O3 in Medical Operating Rooms. J. Chem. Eng. Jpn. 2003, 36, 322–327. [Google Scholar] [CrossRef]
  7. Doi, K.; Wu, Y.Y.; Takeda, R.; Matsunami, A.; Arai, N.; Tagawa, T.; Goto, S. Catalytic decomposition of N2O in medical operating rooms over Rh/Al2O3, Pd/Al2O3, and Pt/Al2O3. Appl. Catal. B Environ. 2001, 35, 43–51. [Google Scholar] [CrossRef]
  8. Wei, X.; Yang, X.-F.; Wang, A.-Q.; Li, L.; Liu, X.-Y.; Zhang, T.; Mou, C.-Y.; Li, J. Bimetallic Au–Pd Alloy Catalysts for N2O Decomposition: Effects of Surface Structures on Catalytic Activity. J. Phys. Chem. C 2012, 116, 6222–6232. [Google Scholar] [CrossRef]
  9. Hermes, A.C.; Hamilton, S.M.; Hopkins, W.S.; Harding, D.J.; Kerpal, C.; Meijer, G.; Fielicke, A.; Mackenzie, S.R. Effects of Coadsorbed Oxygen on the Infrared Driven Decomposition of N2O on Isolated Rh5+ Clusters. J. Phys. Chem. Lett. 2011, 2, 3053–3057. [Google Scholar] [CrossRef]
  10. Rodríguez-Kessler, P.L.; Rodríguez-Domínguez, A.R. N2O dissociation on small Rh clusters: A density functional study. Comput. Mater. Sci. 2015, 97, 32–35. [Google Scholar] [CrossRef]
  11. Rodríguez-Kessler, P.L.; Pan, S.; Florez, E.; Cabellos, J.L.; Merino, G. Structural Evolution of the Rhodium-Doped Silver Clusters AgnRh (n ≤ 15) and Their Reactivity toward NO. J. Phys. Chem. C 2017, 121, 19420–19427. [Google Scholar] [CrossRef]
  12. Russo, N.; Mescia, D.; Fino, D.; Saracco, G.; Specchia, V. N2O Decomposition over Perovskite Catalysts. Ind. Eng. Chem. Res. 2007, 46, 4226–4231. [Google Scholar] [CrossRef]
  13. Jiang, H.; Wang, H.; Liang, F.; Werth, S.; Schiestel, T.; Caro, J. Direct Decomposition of Nitrous Oxide to Nitrogen by In Situ Oxygen Removal with a Perovskite Membrane. Angew. Chem. Int. Ed. 2009, 48, 2983–2986. [Google Scholar] [CrossRef] [PubMed]
  14. Sui, C.; Niu, X.; Wang, Z.; Yuan, F.; Zhu, Y. Activity and deactivation of Ru supported on La1.6Sr0.4NiO4 perovskite-like catalysts prepared by different methods for decomposition of N2O. Catal. Sci. Technol. 2016, 6, 8505–8515. [Google Scholar] [CrossRef]
  15. Liu, N.; Chen, P.; Li, Y.; Zhang, R. N2O Direct Dissociation over MgxCeyCo1−x−yCo2O4 Composite Spinel Metal Oxide. Catalysts 2017, 7, 10. [Google Scholar] [CrossRef]
  16. Kaczmarczyk, J.; Zasada, F.; Janas, J.; Indyka, P.; Piskorz, W.; Kotarba, A.; Sojka, Z. Thermodynamic Stability, Redox Properties, and Reactivity of Mn3O4, Fe3O4, and Co3O4 Model Catalysts for N2O Decomposition: Resolving the Origins of Steady Turnover. ACS Catal. 2016, 6, 1235–1246. [Google Scholar] [CrossRef]
  17. Abu-Zied, B.M.; Bawaked, S.M.; Kosa, S.A.; Ali, T.T.; Schwieger, W.; Aqlan, F.M. Effects of Nd-, Pr-, Tb- and Y-doping on the structural, textural, electrical and N2O decomposition activity of mesoporous NiO nanoparticles. Appl. Surf. Sci. 2017, 419, 399–408. [Google Scholar] [CrossRef]
  18. Carabineiro, S.A.C.; Papista, E.; Marnellos, G.E.; Tavares, P.B.; Maldonado-Hódar, F.J.; Konsolakis, M. Catalytic decomposition of N2O on inorganic oxides: Εffect of doping with Au nanoparticles. Mol. Catal. 2017, 436, 78–89. [Google Scholar] [CrossRef]
  19. Kondratenko, E.V.; Pérez-Ramírez, J. Mechanism and Kinetics of Direct N2O Decomposition over Fe−MFI Zeolites with Different Iron Speciation from Temporal Analysis of Products. J. Phys. Chem. B 2006, 110, 22586–22595. [Google Scholar] [CrossRef]
  20. Liu, N.; Zhang, R.; Chen, B.; Li, Y.; Li, Y. Comparative study on the direct decomposition of nitrous oxide over M (Fe, Co, Cu)–BEA zeolites. J. Catal. 2012, 294, 99–112. [Google Scholar] [CrossRef]
  21. Uzunova, E.L.; Mikosch, H. A theoretical study of nitric oxide adsorption and dissociation on copper-exchanged zeolites SSZ-13 and SAPO-34: The impact of framework acid-base properties. Phys. Chem. Chem. Phys. 2016, 18, 11233–11242. [Google Scholar] [CrossRef] [PubMed]
  22. Guesmi, H.; Berthomieu, D.; Bromley, B.; Coq, B.; Kiwi-Minsker, L. Theoretical evidence of the observed kinetic order dependence on temperature during the N2O decomposition over Fe-ZSM-5. Phys. Chem. Chem. Phys. 2010, 12, 2873–2878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Marnellos, G.E.; Efthimiadis, E.A.; Vasalos, I.A. Effect of SO2 and H2O on the N2O decomposition in the presence of O2 over Ru/Al2O3. Appl. Catal. B Environ. 2003, 46, 523–539. [Google Scholar] [CrossRef]
  24. Marnellos, G.E.; Efthimiadis, E.A.; Vasalos, I.A. Simultaneous Catalytic Reduction of NOX and N2O in an In/Al2O3−Ru/Al2O3 Dual-Bed Reactor in the Presence of SO2 and H2O. Ind. Eng. Chem. Res. 2004, 43, 2413–2419. [Google Scholar] [CrossRef]
  25. Hansen, N.; Heyden, A.; Bell, A.T.; Keil, F.J. A Reaction Mechanism for the Nitrous Oxide Decomposition on Binuclear Oxygen Bridged Iron Sites in Fe-ZSM-5. J. Phys. Chem. C 2007, 111, 2092–2101. [Google Scholar] [CrossRef]
  26. Heyden, A.; Hansen, N.; Bell, A.T.; Keil, F.J. Nitrous Oxide Decomposition over Fe-ZSM-5 in the Presence of Nitric Oxide: A Comprehensive DFT Study. J. Phys. Chem. B 2006, 110, 17096–17114. [Google Scholar] [CrossRef]
  27. Heyden, A.; Peters, B.; Bell, A.T.; Keil, F.J. Comprehensive DFT Study of Nitrous Oxide Decomposition over Fe-ZSM-5. J. Phys. Chem. B 2005, 109, 1857–1873. [Google Scholar] [CrossRef]
  28. Bulushev, D.A.; Prechtl, P.M.; Renken, A.; Kiwi-Minsker, L. Water Vapor Effects in N2O Decomposition over Fe−ZSM-5 Catalysts with Low Iron Content. Ind. Eng. Chem. Res. 2007, 46, 4178–4185. [Google Scholar] [CrossRef]
  29. Ju, W.-S.; Matsuoka, M.; Iino, K.; Yamashita, H.; Anpo, M. The Local Structures of Silver(I) Ion Catalysts Anchored within Zeolite Cavities and Their Photocatalytic Reactivities for the Elimination of N2O into N2 and O2. J. Phys. Chem. B 2004, 108, 2128–2133. [Google Scholar] [CrossRef]
  30. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef]
  31. Higashimoto, S.; Nishimoto, K.; Ono, T.; Anpo, M. Characterization of Fe-Oxide Species Prepared onto ZSM-5 Zeolites and Their Role in the Photocatalytic Decomposition of N2O into N2 and O2. Chem. Lett. 2000, 29, 1160–1161. [Google Scholar] [CrossRef]
  32. Rusu, C.N.; Yates, J.T. N2O Adsorption and Photochemistry on High Area TiO2 Powder. J. Phys. Chem. B 2001, 105, 2596–2603. [Google Scholar] [CrossRef]
  33. Sano, T.; Negishi, N.; Mas, D.; Takeuchi, K. Photocatalytic Decomposition of N2O on Highly Dispersed Ag+ Ions on TiO2 Prepared by Photodeposition. J. Catal. 2000, 194, 71–79. [Google Scholar] [CrossRef]
  34. Kim, B.; Li, Z.; Kay, B.D.; Dohnálek, Z.; Kim, Y.K. Low-Temperature Desorption of N2O from NO on Rutile TiO2(110)-1 × 1. J. Phys. Chem. C 2014, 118, 9544–9550. [Google Scholar] [CrossRef]
  35. Obalová, L.; Reli, M.; Lang, J.; Matějka, V.; Kukutschová, J.; Lacný, Z.; Kočí, K. Photocatalytic decomposition of nitrous oxide using TiO2 and Ag-TiO2 nanocomposite thin films. Catal. Today 2013, 209, 170–175. [Google Scholar] [CrossRef] [Green Version]
  36. Kočí, K.; Krejčíková, S.; Šolcová, O.; Obalová, L. Photocatalytic decomposition of N2O on Ag-TiO2. Catal. Today 2012, 191, 134–137. [Google Scholar] [CrossRef]
  37. Chatterjee, T.; Shetti, V.S.; Sharma, R.; Ravikanth, M. Heteroatom-Containing Porphyrin Analogues. Chem. Rev. 2017, 117, 3254–3328. [Google Scholar] [CrossRef]
  38. Zhang, W.; Lai, W.; Cao, R. Energy-Related Small Molecule Activation Reactions: Oxygen Reduction and Hydrogen and Oxygen Evolution Reactions Catalyzed by Porphyrin- and Corrole-Based Systems. Chem. Rev. 2017, 117, 3717–3797. [Google Scholar] [CrossRef]
  39. Gao, W.-Y.; Chrzanowski, M.; Ma, S. Metal-metalloporphyrin frameworks: A resurging class of functional materials. Chem. Soc. Rev. 2014, 43, 5841–5866. [Google Scholar] [CrossRef]
  40. Antonangelo, A.R.; Grazia Bezzu, C.; Mughal, S.S.; Malewschik, T.; McKeown, N.B.; Nakagaki, S. A porphyrin-based microporous network polymer that acts as an efficient catalyst for cyclooctene and cyclohexane oxidation under mild conditions. Catal. Commun. 2017, 99, 100–104. [Google Scholar] [CrossRef]
  41. Maitarad, P.; Namuangruk, S.; Zhang, D.; Shi, L.; Li, H.; Huang, L.; Boekfa, B.; Ehara, M. Metal–Porphyrin: A Potential Catalyst for Direct Decomposition of N2O by Theoretical Reaction Mechanism Investigation. Environ. Sci. Technol. 2014, 48, 7101–7110. [Google Scholar] [CrossRef] [PubMed]
  42. Guilard, R.; Latour, J.M.; Lecomte, C.; Marchon, J.C.; Protas, J.; Ripoll, D. Peroxotitanium (IV) porphyrins. Synthesis, stereochemistry, and properties. Inorg. Chem. 1978, 17, 1228–1237. [Google Scholar] [CrossRef]
  43. Woo, L.K.; Hays, J.A.; Goll, J.G. Oxo-transfer reactions of chromium and titanium porphyrins. Inorg. Chem. 1990, 29, 3916–3917. [Google Scholar] [CrossRef] [Green Version]
  44. Latour, J.-M.; Galland, B.; Marchon, J.-C. Retention of configuration at the titanium atom upon oxo-peroxo ligand substitution of titanium (IV) porphyrins. J. Chem. Soc. Chem. Commun. 1979, 13, 570–571. [Google Scholar] [CrossRef]
  45. Pirngruber, G.D. The surface chemistry of N2O decomposition on iron containing zeolites (I). J. Catal. 2003, 219, 456–463. [Google Scholar] [CrossRef]
  46. Winter, E.R.S. The decomposition of nitrous oxide on the rare-earth sesquioxides and related oxides. J. Catal. 1969, 15, 144–152. [Google Scholar] [CrossRef]
  47. Liu, X.; Yang, Z.; Zhang, R.; Li, Q.; Li, Y. Density Functional Theory Study of Mechanism of N2O Decomposition over Cu-ZSM-5 Zeolites. J. Phys. Chem. C 2012, 116, 20262–20268. [Google Scholar] [CrossRef]
  48. Zhao, Y.; Truhlar, D.G. The M06 suite of density functionals for main group thermochemistry, thermochemical kinetics, noncovalent interactions, excited states, and transition elements: Two new functionals and systematic testing of four M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar]
  49. Zhao, Y.; Truhlar, D.G. Density Functional Theory for Reaction Energies: Test of Meta and Hybrid Meta Functionals, Range-Separated Functionals, and Other High-Performance Functionals. J. Chem. Theory Comput. 2011, 7, 669–676. [Google Scholar] [CrossRef]
  50. Mom, R.V.; Cheng, J.; Koper, M.T.M.; Sprik, M. Modeling the Oxygen Evolution Reaction on Metal Oxides: The Infuence of Unrestricted DFT Calculations. J. Phys. Chem. C 2014, 118, 4095–4102. [Google Scholar] [CrossRef]
  51. Enriquez, J.I.G.; Moreno, J.L.V.; David, M.Y.; Arboleda, N.B.; Lin, O.H.; Villagracia, A.R.C. DFT Investigation on the Electronic and Water Adsorption Properties of Pristine and N-Doped TiO2 Nanotubes for Photocatalytic Water Splitting Applications. J. Electron. Mater. 2017, 46, 3592–3602. [Google Scholar] [CrossRef]
  52. Hansen, J.A.; Ehara, M.; Piecuch, P. Aerobic Oxidation of Methanol to Formic Acid on Au8: Benchmark Analysis Based on Completely Renormalized Coupled-Cluster and Density Functional Theory Calculations. J. Phys. Chem. A 2013, 117, 10416–10427. [Google Scholar] [CrossRef] [PubMed]
  53. Boekfa, B.; Pahl, E.; Gaston, N.; Sakurai, H.; Limtrakul, J.; Ehara, M. C–Cl Bond Activation on Au/Pd Bimetallic Nanocatalysts Studied by Density Functional Theory and Genetic Algorithm Calculations. J. Phys. Chem. C 2014, 118, 22188–22196. [Google Scholar] [CrossRef]
  54. Meeprasert, J.; Namuangruk, S.; Boekfa, B.; Dhital, R.N.; Sakurai, H.; Ehara, M. Mechanism of Ullmann Coupling Reaction of Chloroarene on Au/Pd Alloy Nanocluster: A DFT Study. Organometallics 2016, 35, 1192–1201. [Google Scholar] [CrossRef]
  55. 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 Rev. B.01; Gaussian Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
Figure 1. Structure of oxotitanium(IV) porphyrin (TiO-por).
Figure 1. Structure of oxotitanium(IV) porphyrin (TiO-por).
Catalysts 10 00157 g001
Figure 2. Water dissociation over the TiO-por active site. Light grey, red, blue, dark grey, and white balls represent Ti, O, N, C, and H atoms, respectively. Bond lengths are in Å.
Figure 2. Water dissociation over the TiO-por active site. Light grey, red, blue, dark grey, and white balls represent Ti, O, N, C, and H atoms, respectively. Bond lengths are in Å.
Catalysts 10 00157 g002
Figure 3. The first N2O decomposition over the hydroxyl-oxotitanium porphyrin (Ti(OH)2-por).
Figure 3. The first N2O decomposition over the hydroxyl-oxotitanium porphyrin (Ti(OH)2-por).
Catalysts 10 00157 g003
Figure 4. Path A: The second N2O decomposition followed by water desorption and oxygen formation.
Figure 4. Path A: The second N2O decomposition followed by water desorption and oxygen formation.
Catalysts 10 00157 g004
Figure 5. Path B: Water desorption followed by the second N2O decomposition and oxygen formation.
Figure 5. Path B: Water desorption followed by the second N2O decomposition and oxygen formation.
Catalysts 10 00157 g005
Figure 6. Full energy profile of N2O decomposition over the oxotitanium porphyrin.
Figure 6. Full energy profile of N2O decomposition over the oxotitanium porphyrin.
Catalysts 10 00157 g006
Scheme 1. Overall reaction pathways of the photo-assisted N2O decomposition with/without a water molecule (the energy is in kcal/mol).
Scheme 1. Overall reaction pathways of the photo-assisted N2O decomposition with/without a water molecule (the energy is in kcal/mol).
Catalysts 10 00157 sch001
Table 1. Activation energy (kcal/mol) for N2O direct decomposition including the water molecule on singlet and triplet states of TiOH-por catalysts.
Table 1. Activation energy (kcal/mol) for N2O direct decomposition including the water molecule on singlet and triplet states of TiOH-por catalysts.
Reaction StepActivation Energy (kcal/mol)
1TiOH-por 3TiOH-por
1st N2O decomposition
Ti(OH)2-por + N2O → TiO(OH)2-por + N2
63.5727.57
Path A 2nd N2O decomposition and water desorption
TiO(OH)2-por + N2O → Ti(OOH)2-por + N2
67.8729.01
Path B water desorption and 2nd N2O decomposition
TiO2-por + N2O → TiO3-por + N2
68.4112.36
Oxygen formation
TiO3-por → TiO-por + O2
5.12barrierless
Table 2. Activation energy (kcal/mol) of the first and second N2O decompositions and oxygen formation over the Fe-ZSM5 and oxotitanium(IV) porphyrin.
Table 2. Activation energy (kcal/mol) of the first and second N2O decompositions and oxygen formation over the Fe-ZSM5 and oxotitanium(IV) porphyrin.
ProcessActivation Energy (kcal/mol)
Z[FeO]+ aZ[Fe(OH)2] + a3TiO-por b3Ti(OH)2-por
1st N2O30.442.829.927.6
2nd N2O20.161.711.512.4
O2 formation8.016.00.6barrier less
a Ref [27], b Ref [41].

Share and Cite

MDPI and ACS Style

Maitarad, P.; Promarak, V.; Shi, L.; Namuangruk, S. Effect of Water Molecule on Photo-Assisted Nitrous Oxide Decomposition over Oxotitanium Porphyrin: A Theoretical Study. Catalysts 2020, 10, 157. https://doi.org/10.3390/catal10020157

AMA Style

Maitarad P, Promarak V, Shi L, Namuangruk S. Effect of Water Molecule on Photo-Assisted Nitrous Oxide Decomposition over Oxotitanium Porphyrin: A Theoretical Study. Catalysts. 2020; 10(2):157. https://doi.org/10.3390/catal10020157

Chicago/Turabian Style

Maitarad, Phornphimon, Vinich Promarak, Liyi Shi, and Supawadee Namuangruk. 2020. "Effect of Water Molecule on Photo-Assisted Nitrous Oxide Decomposition over Oxotitanium Porphyrin: A Theoretical Study" Catalysts 10, no. 2: 157. https://doi.org/10.3390/catal10020157

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