Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Reduction of NOx on metal-free hydrogenated hexagonal boron nitride

Anthony J. R. Payne *, Neubi F. Xavier Jr and Marco Sacchi *
School of Chemistry and Chemical Engineering, University of Surrey, Guildford, GU2 7XH, UK. E-mail: apayne@surrey.ac.uk; m.sacchi@surrey.ac.uk

Received 15th February 2024 , Accepted 12th June 2024

First published on 19th June 2024


Abstract

Sustainable catalysts are essential for critical industrial and environmental processes. 2D materials have exceptional surface area and unique thermal and electronic properties, making them excellent candidates for catalytic applications. Moreover, 2D materials can be functionalised to create metal-free active sites, which provide sustainable alternatives to transition and precious metals. Among the pollutants emitted by combustion engines, NOx stands out as one of the most detrimental gases, contributing to environmental pollution and posing risks to human health. We demonstrate that functionalised defects in hexagonal boron nitride (hBN) provide a thermodynamically viable route to removing NOx by reaction with a hydrogenated boron vacancy (3HVB). The decomposition of NO2 proceeds by initially overcoming an activation energy barrier of 1.12 eV to transfer a hydrogen atom from the surface, forming a NO2H species, followed by the elimination of a water molecule. A thermodynamically favourable product consisting of a surface-bound hydroxyl adjacent to a nitrogen antisite defect (where a nitrogen atom occupies a site typically occupied by a boron atom) forms after overcoming an energy barrier of 1.28 eV. NO can further decompose by overcoming an activation energy barrier of 2.23 eV to form a surface HNO species. A rearrangement of the HNO species takes place with an activation energy of 1.96 eV, followed by the elimination of water. The overall reactions reduce NOx into defective hBN and H2O.


1 Introduction

The high temperatures and pressures experienced during fuel combustion enable the fixation of N2 and O2, resulting in NO and NO2 emissions (referred to as NOx).1 NOx is known to cause ailments such as impaired lung function and irritation of the eyes and throat.2,3 The environment is adversely affected by NOx dissolving in rainwater, forming nitric acid (HNO3), and creating acid rain.4 Acid rain has many detrimental effects, including hindering plant growth above and below ground and damaging plant leaves, with food-producing plants most severely affected.5,6 NOx pollution can contaminate soils and water bodies, leading to eutrophication and acidification.7 Furthermore, NO2 is the starting point for the photochemical formation of ground-level ozone,8,9 leading to further detrimental effects on ecosystems and thousands of premature deaths.10–12 The World Health Organisation Global Air Quality Guidelines specify a NO2 limit of 10 g m−3 as an annual average; however, this limit is often exceeded in major cities such as New York, London, Paris, Delhi and Beijing.13

The main emitters of NOx are vehicles (with diesel engines being exceptionally high emitters),14 power plants, agricultural and industrial processes.15,16 Despite technological advances in engine manufacturing and the rapid increase of electric vehicles, fossil fuels still make up a large proportion of the energy market to meet demand.17 Hence, much research has focused on eliminating the toxic and polluting gases emitted by fossil fuel combustion.18

Using metal-based heterogeneous catalysts for emission control is widely accepted as an effective way of reducing pollution from NOx and gas exhaust emissions, with many vehicles equipped with selective catalytic reduction (SCR) systems.19,20 SCR reduces NOx into environmentally benign N2 and H2O by utilising a catalyst and reducing agent.21 Reducing agents include NH3 and H2 in NH3-SCR and H2-SCR. NH3-SCR has shown to be an attractive choice for low-temperature reduction; however, challenges include N2O formation and the need for an additional system to supply and store ammonia in applications.21,22 Safe storage and distribution of H2 remain technological challenges which delay the widespread employment of hydrogen use as a fuel.23 However, recently, there has been significant interest in producing H2 “on-demand” via water-gas shift reaction post-combustion (CO + H2O → CO2 + H2),24 thus providing a source of H2 for H2-SCR. Presently, automotive catalysts dominate the global demand for platinum group metals,25 with platinum and palladium commonly employed for H2-SCR.21 However, platinum and palladium catalysts often exhibit peak activity at temperatures around 200 °C,21 whereas catalytic converters in vehicles frequently surpass this threshold with passenger vehicles operating over a wide range of temperatures (280–650 °C).26 Therefore, there is potential to find alternative catalysts that are active at higher temperatures. Perovskites have recently been of interest as a greener alternative for NOx removal but still rely (although quantitatively less) on metal atoms in their structure.27,28 Metal-free NH3-SCR has been investigated using ketone-terminated graphene nanoribbon edges,29,30 yet metal-free H2-SCR remains underexplored within a critically significant field.

Hexagonal boron nitride (hBN) consists of sp2 hybridised boron and nitrogen atoms forming single-atom thick nanosheets.31 These so-called ‘2D’ nanosheets are held together by van der Waals forces and stacked in an ABAB pattern.32 Due to the B–N bond's ionic character, charge localisation33 hBN is an insulator with a large band gap of 6 eV. Potential applications of hBN include electrical,34 optoelectronic35 and spintronic36 devices. Due to its outstanding antiwear and antifriction properties, hBN is also employed in the automotive industry as a lubricant.37 hBN can be grown by various methods. Bottom-up approaches include chemical vapour deposition (CVD), which uses boron and nitrogen-containing precursors, such as borazine or ammonia borane, to grow epitaxial, single layers of hBN on a metal substrate.38 Alternatively, hBN can be formed by reacting molecules containing boron and nitrogen, such as boric oxide and urea.39 In each case, a single layer of hBN can be produced by subsequent physical or chemical exfoliation.40

hBN has high thermal stability,31,41 a low coefficient of thermal expansion,42 and a high specific surface area;43 these properties make hBN ideally suited for industrial catalysis. Through high-resolution transmission electron microscopy, hBN has been found to contain various vacancy defects, including boron (VB) and nitrogen (VN) single-atom vacancies.44,45 Ball milling has been established as an effective method for creating additional defect sites on hBN.46,47 Using VN as an active site to remove NO2 has already been the subject of previous density functional theory (DFT) studies.48,49 In these earlier studies, the VN defect was repaired, and molecular oxygen was produced. Pd and Ni-doped hBN have also been considered to catalyse NO reduction by CO to form N2 and CO2.50 Other dopants such as carbon,51 silicon,52 and rhodium53 doped hBN have also been investigated for the fabrication of sensors for NO2.

Research on hBN defects has shown the hydrogenated VB (HVB) to be significantly more stable than the hydrogenated VN defect and the non-hydrogenated counterpart VB supported by X-ray absorption and luminescence measurements.54,55 We have previously reported that HVB vacancies can be effectively produced by exposing boron–nitrogen divacancies to ammonia.56 Furthermore, as hydrogen atoms are ordinarily contained in hBN precursors, HVB defects are likely to be generated during hBN processing, such as during the growth of hBN nanowalls.57

Previous work by Nash et al.58 investigated HVB defects as active sites for the hydrogenation of alkenes on metal-free hBN. The authors reported a reactor based on a pebble mill to introduce HVB defects by mechanical force. The hydrogenation reaction was achieved with high yields of 97–100% for several alkenes. The work was supported by DFT calculations, which determined a barrier for the hydrogenation of propane to propane of 1.53 eV. Given the activity of HVB defects in the hydrogenation of alkenes, other similar reactions may be possible.

A deep understanding of the molecular level mechanisms is required to develop more efficient and greener NOx catalysts. This study investigates the reactivity of the HVB with NO and NO2 molecules. Our computational study shows that HVB defects in hBN can provide an excellent metal-free site for removing NOx by providing a stable source of hydrogen for reduction.

2 Computational methodology

First-principles electronic structure calculations were carried out using CASTEP.59 The PBE functional60 was used to treat exchange and correlation in combination with the Tkatchenko–Scheffler dispersion correction method.61 The plane wave basis set was expanded to a cut-off energy of 400 eV. All calculations used Vanderbilt ultrasoft pseudopotentials62 and a Monkhorst–Pack k-point sampling of the first Brillouin zone63 with a uniform grid of (10 × 10 × 1) or (4 × 4 × 1) for modelling of (1 × 1) and (5 × 5) unit cells, respectively. The surface structures were optimised until the maximum force on each atom was less than 0.025 eV Å−1 and a self-consistent field energy tolerance of 1 × 10−6 eV. During relaxation, the lattice parameters remained fixed to those of a pristine hBN unit cell. The transition states along the reaction pathways were identified with the linear and quadratic synchronous transit (LST/QST) algorithm,64 with a force tolerance of 0.1 eV Å−1. A (5 × 5) unit cell was employed for all defect calculations, with a vacuum region of at least 20 Å to separate the periodically repeated images and avoid spurious interactions. Molecular diagrams were produced using VESTA.65

Adsorption energies EAds were calculated using the standard formula:

 
EAds = EABEAEB(1)
where EAB is the energy of the adsorbed species, EA and EB are the energy of the dissociated species.

Formation energies Ef were calculated using the formula:

 
image file: d4cy00206g-t1.tif(2)
where EV is the energy of a vacancy or system, EhBN is the energy of the pristine hBN sheet, nN and nB are the number of nitrogen and boron atoms removed to create a vacancy, respectively. EN2 is the energy of an isolated N2 molecule, EB is the reference energy of a boron atom in its α-rhombohedral phase,66nH is the number of hydrogen atoms added to hydrogenated vacancies and EH2 is the energy of an isolated H2 molecule.

Reaction rate coefficients (KTST) for the surface reactions were calculated according to the transition state theory:

 
image file: d4cy00206g-t2.tif(3)
where kb is the Boltzmann constant, T is the temperature, h is the Plank constant, q and q are the partition functions for the transition state and initial state, respectively, and ΔEa is the activation energy. The adsorption rate coefficients kads were calculated using Hertz–Knudsen kinetics:67
 
image file: d4cy00206g-t3.tif(4)
where P is the partial pressure of the molecule, A is the surface area of the surface site, m is the mass of the particle kb is the Boltzmann constant and T is the temperature.

3 Results and discussion

Each of the vacancy models can be seen in Fig. 1. The structure and properties of VB and VN vacancies have been discussed in detail in previous work and are shown in b and c.56 VN is more stable than VB with Ef of 8.63 eV and 9.75 eV, respectively. VB has C2v symmetry, whereas VN has D3h symmetry due to Jahn–Teller distortion.68–71Fig. 1d shows an HVB with all three nitrogen dangling bonds saturated by hydrogen atoms (3HVB); it is by far the most stable configuration investigated, with a low Ef of 1.98 eV. The geometry of 3HVB can be rationalised through a combination of electronic and steric factors. Comparing the N–H–N angle (θ) variation between vacancies, we can rationalise the increase in in-plane steric hindrance as a function of the increasing number of adsorbed hydrogens. In 3HVB, only one of the three hydrogen atoms points into the hBN plane with a θ of 180° and Hirshfeld charge of +0.07 e. The remaining two hydrogens point above and below the hBN plane where θ is 130° with a Hirshfeld charge of +0.11 e. This difference in angle can be explained by the need to prevent steric clashes of hydrogen atoms and minimise partial positive charges yet maintain sp2 hybridisation of the hBN as much as possible. The H–H distance is 2.49 Å between the outer hydrogens and 1.58 Å between the in-plane hydrogen and each external hydrogen, with an average H–H distance of 1.88 Å. The in-plane nitrogen has a 0.08 e higher N–H bond population than the average N–H bond, reflecting a greater extent of sp2 hybridisation and stronger bonding. The minor change in bond population results in the N–H in-plane bond being shorter by 0.008 Å (0.8%) than the out-of-plane N–H bonds. This minor change in bond length is unlikely to affect the reactivity of 3HVB significantly.
image file: d4cy00206g-f1.tif
Fig. 1 (a) Pristine hBN sheet, (b) VB (Ef = 9.75 eV, spin = 1 μB) (c) VN (Ef = 8.63 eV, spin = 1 μB), (d) 3HVB (Ef = 1.98 eV, spin = 0 μB), (e) 2HVB (Ef = 4.22 eV, spin = 1 μB), (f) 1HVB (Ef = 6.82 eV, spin = 0 μB), (g) 3HVN (Ef = 5.29 eV, spin = 0 μB), (h) 2HVN (Ef = 6.86 eV, spin = 1 μB), (i) 1HVN (Ef = 6.93 eV, spin = 0 μB). Boron atoms are shown in green, and nitrogen atoms are shown in grey.

Removing a hydrogen atom to create a 2HVB increases Ef to 4.22 eV. In 2HVB, the hydrogen atoms are further coplanar with hBN, with an average θ of 155°, and have an H–H distance of 1.83 Å. The loss of a hydrogen atom allows the remaining hydrogens to be orientated more in the plane as there is now less steric hindrance inside the vacancy, allowing for θ values closer to 180°. In 2HVB, the N–H distances are 2.01 Å and 2.16 Å; therefore, one hydrogen atom is slightly closer to the nitrogen atom in this structure. The hydrogens interact with the lone pair on the nitrogen dangling bond asymmetrically, thus reducing the N–H distance for the interacting H. 2HVB has a magnetic moment of 1 μB due to an unpaired electron on the nitrogen dangling bond. These changes in the geometry of 2HVB are not enough to compensate for creating a dangling bond, decreasing the stability of this defect relative to that of 3HVB.

The 1HVB has a higher Ef of 6.82 eV due to the creation of an additional nitrogen-dangling bond. Here, only one hydrogen atom remains on the vacancy; the reduced steric clashes allow for an θ of 157°. It follows that the remaining hydrogen is positioned only slightly above the hBN plane with H–N distances of 1.98 Å and 2.02 Å. Removal of the final hydrogen will form a VB with an Ef of 9.75 eV. It is clear that the stability of HVB increases with an increasing number of hydrogen atoms; hence, 3HVB is the most likely to form and will be the subject of our investigation for the hydrogenation of NO2.

In addition to 1–3HVB, we also investigated the structure and geometry of hydrogenated VN vacancies (1–3HVN) for comparison and are shown in Fig. 1(g–i). The configurations obtained by DFT are in good agreement with the work presented in Surya et al.72 with a detailed comparison provided in the ESI. Of these vacancies, 3HVN has the lowest Ef of 5.29 eV, significantly higher than that of 3HVB (Ef 1.98 eV). The greater decrease in formation energy of boron vacancies compared to nitrogen vacancies upon the addition of hydrogen atoms can be rationalised by considering the electronegativity and geometry of the atoms around the vacancy. As N is more electronegative than B, the N–H bonding is stronger than that of B–H, with average bond lengths of 1.10 Å and 1.19 Å for 3HVB and 3HVN. Furthermore, longer B–H bonds lead to an increased steric clash of hydrogen atoms, requiring greater deformation of the vacancy site. Therefore, 3HVB vacancies maintain stronger N–B bonding than 3HVN with average B–N bond lengths of H-connected atoms of 1.42 Å and 1.44 Å, respectively. Given the stability observed in the 3HVB vacancy and the likelihood of the formation of hydrogenated VB vacancies compared to hydrogenated VN, we have elected to focus our investigation primarily on this site.

3.1 NO2 decomposition

The adsorption of NO2 on pristine hBN was initially considered. A range of sites and orientations were investigated (as described in the ESI), with the most stable adsorption site pictured in Fig. 2. Here, NO2 is adsorbed with the nitrogen atom at a height of 3.32 Å above hBN with the O–N–O plane almost parallel to the hBN plane. In this position, NO2 has an EAds of −0.202 eV, indicative of physisorption. Previous works have also investigated the adsorption of NO2 on hBN and have found the orientation with oxygen atoms pointing upwards, away from the surface, to be the preferred orientation,52 with one study providing an EAds of −0.09 eV.49 However, these studies used generalised gradient approximation exchange–correlation functionals without dispersion corrections, thus underestimating dispersion forces and the adsorption energy of physisorption. In our study, initiating geometry optimisation with the oxygen atoms pointing away from the surface leads to a rotation of the NO2 molecule, resulting in a parallel orientation. The particular stability of NO2 parallel to the plane can be rationalised by considering that the dispersion interaction between the hBN and NO2 is maximised by increasing the contact area between the adsorbate and the substrate. The boron atoms of hBN and NO2 oxygen atoms have average Hirshfeld charges of 0.21 e and −0.07 e, respectively; these charges attract stabilising the adsorbed NO2.
image file: d4cy00206g-f2.tif
Fig. 2 Adsorption site and orientation of NO2 on hBN. The oxygen atoms are angled slightly towards the hBN plane at a height of 3.219 Å. The nitrogen is situated at a height of 3.324 Å. This adsorption position maximises attractive dispersion forces between hBN and NO2 with an EAds of −0.202 eV.

The step-wise mechanism for the reaction of NO2 with 3HVB was then considered by exploring multiple possible reaction pathways to ensure the minimum energy path was found (as described in the ESI). The reaction mechanism, starting from the adsorption of NO2 above 3HVB and ending with the creation of a nitrogen antisite defect (NB) and H2O, can be seen in Fig. 3 with the overall equation NO2 + 3HVB ⇌ NB + H2O + OH. This reaction pathway, TS1–9, consists of nine elementary steps that connect reactants, intermediates, and products through transition states. Fig. S5 of the ESI displays each structure along the reaction path from both perpendicular and parallel orientations. The highest transition state barrier in TS1–9 is 1.28 eV (TS7), and the ΔE of the reaction is −0.99 eV, indicating that this reaction is possible at moderate temperatures and is thermodynamically favourable. We will examine each of these steps in turn.


image file: d4cy00206g-f3.tif
Fig. 3 The reaction pathway of NO2 with a 3HVB to NB and H2O. The product of the reaction, a nitrogen antisite adjacent to a boron site with a hydroxyl group, is used as the reference energy. Transition states are identified by ‡ and labelled TS1–9. The highest energy barrier, TS7, reflects the movement of a hydrogen atom through the hBN plane in preparation for making an NB site. Other steps occur with energy barriers between 0.04 and 1.09 eV. The reaction is thermodynamically favourable with a ΔE of −0.99 eV.

Firstly, a range of NO2 adsorption sites on 3HVB and orientations were examined (Fig. S2 of the ESI). The adsorption energy minimum can be seen in Fig. 4a with an EAds of −0.247 eV, compared to −0.202 eV for physisorption on pristine hBN. Here, NO2 is oriented to bring one of the oxygen atoms close to the surface at a height of 2.93 Å (Fig. 4a). Hydrogen bonding between the oxygen lone pair and the hydrogen pointing above the hBN plane stabilises this orientation with an O–H distance of 2.28 Å, reducing its adsorption energy relative to pristine hBN. Prior to the reaction, the adsorbed NO2 diffuses to a position where an oxygen atom is directly above a hydrogen atom of 3HVB after overcoming a low energy barrier of 0.11 eV (TS1).


image file: d4cy00206g-f4.tif
Fig. 4 (a) Adsorption geometry of NO2 over a 3HVB with an EAds of −0.25 eV. (b) NO2 adsorbed over a 3HVB prior to reaction with an EAds of −0.16 eV. NO2 can migrate from (a) to (b) by overcoming a low energy barrier of 0.11 eV. (c) NO2H adsorbed over 2HVB with an EAds of −0.38 eV. The barrier to transferring a hydrogen atom from (b) to (c) is 1.12 eV.

Next, hydrogen transfer from 3HVB to NO2 was considered to create a NO2H species in a NO2 + 3HVB ⇌ NO2H reaction step. A range of orientations was assessed to find the most stable NO2H conformation. The most stable position of NO2H is close to the vacancy, with an O–H bond directly above the nitrogen dangling bond and with the second oxygen atom above a surface boron atom. Structural relaxations of other geometries either resulted in the formation of 3HVB or produced a higher energy NO2H orientation. As with NO2 adsorption on pristine hBN, the NO2H is stabilised by van der Waals interactions between the oxygen lone pair and surface boron (Fig. 4b). The N–H bond-breaking step requires an activation energy of 1.02 eV (TS2). The products of this step are a 2HVB and a NO2H species. Hydrogen bonding between the NO2H hydrogen moiety and the uncoordinated surface nitrogen gives NO2H an EAds of −0.28 eV.

The NO2H species can easily rotate parallel to the hBN plane. The NO2H rotation proceeds via an initial step with a low activation barrier of only 0.15 eV to bring the oxygen atoms parallel to the plane (TS3), as shown in Fig. 5b. This barrier reflects the breaking of van der Waals forces between the surface and NO2H. A second rotation occurs with a more significant energy barrier of 0.72 eV (TS4). This higher barrier reflects the need to overcome the repulsive forces of the nitrogen lone pairs and form an N–N bond between NO2H and the surface nitrogen dangling bond. NO2H is now chemisorbed onto the surface, as shown in Fig. 5c. Here, the surface nitrogen is puckered 0.5 Å above the plane, and the N–N bonded nitrogen atoms have sp3 hybridised character. Subsequently, after forming a NO2H species, the hydroxyl can accept another hydrogen atom to produce water (TS5, NO2H + H ⇌ NO + H2O) after overcoming an energy barrier of 0.96 eV. The produced water molecule is hydrogen bonded to the surface. Adsorbed water could remain on the surface and contribute to stabilising further structures and transition states through hydrogen bonding; therefore, it continues to feature in our simulations.


image file: d4cy00206g-f5.tif
Fig. 5 (a) Adsorption site of NO2H over 2HVB with an EAds of −0.38 eV. (b) NO2H adsorbed over 2HVB before reaction to form a chemisorbed NO2H species (c). The rotation barrier from (a) to (b) is 0.15 eV. The transition from (b) to (c) has a barrier of 0.72 eV. Charge accumulation is shown in blue, and charge depletion in red with respect to an isolated NO2H (isosurface cut-off in (a) and (b) is 0.01 e Å−3). (a) Shows a much greater change in charge densities due to the stronger van der Waals interactions with the surface. NO2H decomposes to produce H2O after overcoming a reaction barrier of 0.96 eV to change from (c) to (d).

Following the formation of a water molecule (TS5), the surface structure now contains an O[double bond, length as m-dash]N–N moiety, a nitrogen dangling bond in the vacancy, and an N–H group. Bonding of the O[double bond, length as m-dash]N nitrogen to the nitrogen dangling bond can substantially lower the total energy of the surface. This step occurs after overcoming a low barrier of 0.04 eV (TS6) and forms a nitrogen antisite defect (NB), where the nitrogen atom occupies the site usually occupied by boron in hBN. The NB nitrogen atom is bound to both the nitrogen atoms of the vacancy and the oxygen atom above, as seen in Fig. 6a. Further surface rearrangement takes place to bring the sub-layer hydrogen through the vacancy before positioning itself above the plane 1.87 Å away from the surface oxygen. This rearrangement occurs with an energy barrier of 1.28 eV (TS7) due to steric hindrance destabilising the transition state as the hydrogen needs to pass through an in-plane position (Fig. 6b), and the subsequent geometry has moved the N–O bond into a sp3-like orientation. These steric factors cause TS7 to proceed with the highest barrier in the decomposition of NO2.


image file: d4cy00206g-f6.tif
Fig. 6 (a) NB with additional oxygen and hydrogen atoms positioned above and below the hBN plane, respectively (b) transition state to bring the hydrogen atom from below to an above plane position with a barrier of 1.28 eV (TS7). (c) NB with additional oxygen and hydrogen atoms both positioned above the plane.

Surface hydrogen and oxygen atoms can now form a surface hydroxyl group bound to the antisite nitrogen (NB) after overcoming an energy barrier of 0.82 eV, as shown in TS8, reflecting the breaking of an N–H bond and the forming of an H–O bond (RN–H + NBO ⇌ RN + NBO–H). Given the electron density accumulation provided by the four nitrogen atoms and the oxygen around the NB, it is thermodynamically favourable for the hydroxyl species to migrate to an adjacent boron atom as it has a lower electron density, as shown in Fig. 7. The migration through TS9 (with an energy barrier of 0.33 eV) minimises the electron repulsion and satisfies the remaining nitrogen dangling bond, reducing the overall energy of the product.


image file: d4cy00206g-f7.tif
Fig. 7 Electron density slice contour map from 0 to 3 e Å−3 of (a) NB with attached hydroxyl functional group and (b) NB adjacent to a surface boron site with a hydroxyl group attached. The barrier to transition from (a) to (b) is 0.33 eV (TS9) in an exothermic process with a reaction enthalpy of −2.73 eV. Note: the water molecule is in the plane of the slice in (b) but not in (a). In (b), the electron density surrounding the NB site repels that of the adsorbed oxygen. Therefore, the hydroxyl migrates to an adjacent boron site (a), lowering electron–electron repulsion and producing a more thermodynamically stable product.

The overall reaction forms a water molecule and a surface-bound hydroxyl adjacent to a nitrogen antisite defect (NO2 + 3HVB ⇌ NB + H2O + OH). This reaction is thermodynamically favourable to form with a ΔE of −0.98 eV. The highest energy barrier for a single step in the overall NO2 reduction mechanism is 1.28 eV. The barrier for the reaction of NO2 with HVB is comparable to that of NO2 with VN as explored in work by Feng et al.49 where NO2 decomposes to form O2 and the ‘healing’ of VN to form a pristine hBN sheet with a barrier of 1.17 eV. We note that this paper did not include dispersion correction in the DFT calculations, therefore likely underestimating the energy of adsorption. Given the low formation energy of the initial 3HVB defect, the relatively low energy barrier of the rate-determining step and the production of a thermodynamically favourable product, the catalytic reduction of NO2 over hydrogenated hBN defects appears to be a promising way of controlling NO2 emissions from combustion engines.

If an additional hydrogen atom is added, further steps could be taken to produce a second water molecule, as shown in Fig. 8. The hydroxyl group can then be hydrogenated to form a second water molecule (OH + ½H2 ⇌ H2O) after overcoming an energy barrier of 0.36 eV (TS10). This step is exothermic with an overall ΔE of −1.30 eV. The thermodynamic stability is due to the creation of a second water molecule, the return to sp2 binding of the surface boron atom and the original water molecule being able to adsorb strongly on top of the NB site. An isolated NB has an Ef of 5.45 eV, higher than that of a 2HVB vacancy with an Ef of 4.22 eV. The overall product of the reaction NO2 + 3HVB + H ⇌ NB + 2H2O between NO2 and a 3HVB, in the presence of hydrogen atoms, is an NB and two water molecules. Once exhausted, new vacancy defects need to be introduced to regenerate active sites. The difference in Ef between NB and 3HVB is 3.47 eV, thus requiring an input of energy to regenerate the active site. The introduction of such sites can be achieved by mechanical methods such as ball milling,46,47 and has been achieved in work by Nash et al.58 to complete a catalytic cycle.


image file: d4cy00206g-f8.tif
Fig. 8 Reaction step of the hydroxyl group on a surface-born atom adjacent to a NB site with an adsorbed hydrogen atom. Transition states are identified by ‡ labelled TS10. An energy barrier of 0.36 eV facilitates the hydrogenation of the hydroxyl group to create a water molecule as the product. This step is thermodynamically favourable with a ΔE of −1.30 eV.

The overall reaction of NO2 to NB and two water molecules is thermodynamically feasible, yet many activated steps are involved. Therefore, a microkinetic model consisting of 10 elementary steps was constructed to evaluate the reaction's kinetic viability. The consumption of NO2 and production of H2O at different temperatures, and a pressure of 10 bar was chosen to align with the optimal conditions for the water-gas shift reaction24 (further details of the microkinetic model are reported in the ESI). The rate of NO2 consumption compared to the rate of H2O formation was calculated with results shown in Fig. 9. Below 800 K, the rate of NO2 consumption remains very low (<1.10 × 10−6 s−1); increasing the temperature, the rate of H2O production rises to a maximum of 7.48 × 10−5 s−1 at 990 K, accompanied by a NO2 consumption rate of 3.68 × 10−5 s−1. As temperatures rise, the energy of the molecules increases, allowing multiple reaction barriers to be overcome. Beyond the optimal temperature, the reactants cease to adsorb onto the surface and persist in the gas phase. Our results show that NO2 decomposition can occur at 3HVB defect sites in hBN via a Langmuir–Hinshelwood mechanism.


image file: d4cy00206g-f9.tif
Fig. 9 Microkinetic simulations for the reaction of NO2 and 3HVB displaying the consumption of NO2 compared to the production of H2O as a function of temperature.

3.2 NO decomposition

NO pollution shares many of the same detrimental effects as NO2 pollution leading to acid rain; hence, these two species and their effects are often combined in NOx. Moreover, NO oxidised to NO2 by reacting with sources of O.73 We included the reaction of NO in the scope of our work to explore any possible reactions with 3HVB. Adsorption of NO was first considered above pristine hBN. The most favourable adsorption site is pictured in Fig. 10 and adsorbs with an EAds of −0.17 eV, similar to NO2.
image file: d4cy00206g-f10.tif
Fig. 10 Adsorption site and orientation of NO on hBN. The nitrogen atom is angled slightly towards the hBN plane at a height of 3.07 Å. The oxygen atom is situated in the centre of a hBN ring. This adsorption position maximises attractive dispersion forces between hBN and NO2 with an EAds of −0.17 eV.

NO adsorption on 3HVB was considered by starting with the most stable orientation pictured in Fig. 11 with an EAds of −0.240 eV at a height of 3.041 Å, similar to NO2. Multiple reaction routes were then analysed; here, the reaction begins with the direct reaction of NO with 3HVB to form a chemisorbed HNO species after overcoming a high energy barrier of 2.23 eV (TS11, NO + 3HVB ⇌ HNO + 2HVB). The size of this barrier reflects the breaking of the N–H bond and stretching of the N–O bond to accommodate the additional coordination of the nitrogen atom, partially stabilised by the nitrogen lone pair orientated towards the surface hydrogen. This step has the highest barrier for the reaction of NO with 3HVB. Hydrogen transfer to produce a hydroxyl group occurs with another high energy barrier of 1.96 eV (TS12). Following the creation of a hydroxyl group, the NOH intermediate can accept an N–H hydrogen atom in TS13 (ΔE of 1.63 eV) to form a chemisorbed HNOH species and a dangling bond over the nitrogen atom. The reaction's final step (TS14) involves the migration of the nitrogen atom to bind with the nitrogen dangling bond of the vacancy to create an NB site with a remaining hydrogen attached (HNB). Simultaneously, the hydroxyl group accepts a hydrogen atom to form a water molecule. The overall reaction NO + 3HVB ⇌ H2O + HNB is slightly exothermic with a ΔE of −0.09 eV. The surface product of the reaction (HNB) has an Ef of 5.42 eV.


image file: d4cy00206g-f11.tif
Fig. 11 The reaction pathway of NO with a 3HVB to NB and H2O. The product of the reaction, a molecule of water and an NB site with the final surface hydrogen remaining, is used as a reference energy. Transition states are identified by ‡ and labelled TS11–14. The highest energy barrier at TS1 (2.23 eV) reflects the energy to break an N–H bond and lengthen the N–O bond simultaneously. The reaction is thermodynamically favourable with a ΔE of −0.09 eV.

The reaction pathway of NO with 3HVB significantly differs from that of NO2. For NO reduction, the hydrogen transfer to create a chemisorbed HNOx species occurs with a much higher energy barrier than NO2, with activation energies of 2.23 eV and 1.28 eV, respectively, for this step. The weaker N–O bonds in NO2 with O[double bond, length as m-dash]N–O and O–N[double bond, length as m-dash]O resonance structures result in greater electron density on the oxygen atoms, making an O[double bond, length as m-dash]N–O–H species possible. Thus, NO2 can accept a hydrogen atom from 3HVB to form a physisorbed NO2H species before chemisorption. In contrast, NO has a stronger N–O bond with less electron density available to accept a hydrogen atom on the O or N atoms. Therefore, hydrogen transfers and chemisorption occur in a single step with a higher energy barrier, with O–H and N–N bonds forming simultaneously. The barriers relating to NO decomposition on 3HVB are significantly higher than those revealed in work by Xiao et al.48 where the reaction of NO with a VN results in the formation of a B–O–N epoxy structure. The reaction of 3HVB with NO is less exothermic with a ΔE of −0.09 compared to NO2 with a ΔE of −0.98 eV. Here, the product of the reaction is an NB site with an additional N–H bond pointing hydrogen below the hBN monolayer.

Our results show that NO2 decomposition is energetically facile at 3HVB defect sites in hBN with a barrier of 1.28 eV and achieves optimum NO2 consumption at 990 K. The decomposition product is H2O, an NB site adjacent to a surface hydroxyl species. This hydroxyl species can then be hydrogenated to yield a second water molecule. Similarly, NO can be decomposed by NO with a limiting step of 2.23 eV to create an NB and a vacancy hydrogen.

4 Conclusions

In conclusion, we have investigated the reaction of NOx with a 3HVB site in a “free-standing” hBN mono-layer using DFT calculations. NO2 decomposition proceeds via Langmuir–Hinshelwood hydrogenation of NO2 to form a NO2H species before the chemisorption of this intermediate on the hBN vacancy. The reaction follows several further steps that end with eliminating a molecule of water, forming an NB site adjacent to a surface-bound hydroxyl. An additional hydrogen atom diffusing from other sites can react with the hydroxyl and yield a second water molecule. The atomistic details of the NO2 reduction mechanism, encompassing ten elementary steps, have been described in section 3, focusing on linking reaction energy and barriers to steric and electronic effects. Our calculations prove that the decomposition of NO2 over hydrogenated hBN vacancies is a feasible approach to removing NO2. The rate-determining step has an energy barrier of 1.28 eV. Using micro-kinetics to evaluate the reaction path as a whole, we have calculated an optimum temperature of 990 K under a pressure of 10 bar with a very low rate of reaction below 800 K. The NO decomposition proceeds via the formation of a surface HNO species before rearrangement and elimination of a water molecule before forming an NB. The NO reaction mechanism progresses through four steps (discussed in section 3.2). The decomposition of NO to NB + H2 has an energy barrier of 2.32 eV and is slightly exothermic (0.09 eV). Therefore, higher temperatures are required to achieve NO reduction. These reactions transform harmful and polluting NO2 into benign H2O using a selective and metal-free catalyst. This paper has thoroughly investigated the complete removal of NOx 3HVB defects in hBN, demonstrating its potential to be developed as an effective metal-free catalyst for deNOx applications.

Data availability

The datasets generated and/or analysed during the current study are available in the University of Surry research repository: https://doi.org/10.15126/surreydata.901080.

Author contributions

A. P. conceived the presented idea, performed the DFT simulations, and wrote the manuscript under M. S.' guidance. N. X. provided expertise in simulations and kinetic calculations.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

A. P. acknowledges the University of Surrey for access to the Eureka High Performance Computing facility and funding in the form of a PhD studentship. MS acknowledge the Royal Society for personal funding through a Royal Society University Research Fellowship URF\R\191029. This work used ARCHER2, the UK's national high-performance computing service, via the UK's HPC Materials Chemistry Consortium, which is funded by EPSRC (EP/R029431 and EP/X035859).

References

  1. A. A. Abdel-Rahman, Int. J. Energy Res., 1998, 22, 483–513 CrossRef CAS.
  2. M. Kampa and E. Castanas, Environ. Pollut., 2008, 151, 362–367 CrossRef CAS PubMed.
  3. M. Guarnieri and J. R. Balmes, Lancet, 2014, 383, 1581–1592 CrossRef CAS PubMed.
  4. A. Singh and M. Agrawal, J. Environ. Biol., 2008, 29, 15–24 CAS.
  5. Z. Shi, J. Zhang, Z. Xiao, T. Lu, X. Ren and H. Wei, J. Environ. Manage., 2021, 297, 113213 CrossRef CAS PubMed.
  6. E. Du, D. Dong, X. Zeng, Z. Sun, X. Jiang and W. de Vries, Sci. Total Environ., 2017, 605–606, 764–769 CrossRef CAS PubMed.
  7. W. de Vries, Curr. Opin. Environ. Sci. Health, 2021, 21, 100249 CrossRef.
  8. L. F. Gamon and U. Wille, Acc. Chem. Res., 2016, 49, 2136–2145 CrossRef CAS PubMed.
  9. P. Sicard, E. Paoletti, E. Agathokleous, V. Araminienė, C. Proietti, F. Coulibaly and A. De Marco, Environ. Res., 2020, 191, 110193 CrossRef CAS PubMed.
  10. E. Agathokleous, Z. Feng, E. Oksanen, P. Sicard, Q. Wang, C. J. Saitanis, V. Araminiene, J. D. Blande, F. Hayes, V. Calatayud, M. Domingos, S. D. Veresoglou, J. Peñuelas, D. A. Wardle, A. de Marco, Z. Li, H. Harmens, X. Yuan, M. Vitale and E. Paoletti, Sci. Adv., 2020, 6, 1–18 Search PubMed.
  11. D. Nuvolone, D. Petri and F. Voller, Environ. Sci. Pollut. Res., 2018, 25, 8074–8088 CrossRef CAS PubMed.
  12. P. Sicard, Y. O. Khaniabadi, S. Perez, M. Gualtieri and A. De Marco, Environ. Sci. Pollut. Res., 2019, 26, 32645–32665 CrossRef CAS PubMed.
  13. A. Goshua, C. A. Akdis and K. C. Nadeau, Allergy, 2022, 77, 1955–1960 CrossRef PubMed.
  14. R. O'Driscoll, M. E. Stettler, N. Molden, T. Oxley and H. M. ApSimon, Sci. Total Environ., 2018, 621, 282–290 CrossRef PubMed.
  15. A. T. Hoang, T. T. Huynh, X. P. Nguyen, T. K. T. Nguyen and T. H. Le, Energy Sources, Part A, 2021, 00, 1–21 Search PubMed.
  16. D. H. Nguyen, C. Lin, C. T. Vu, N. K. Cheruiyot, M. K. Nguyen, T. H. Le, W. Lukkhasorn, T. D. H. Vo and X. T. Bui, Environ. Technol. Innovation, 2022, 28, 102809 CrossRef CAS.
  17. T. Ahmad and D. Zhang, Energy Rep., 2020, 6, 1973–1991 CrossRef.
  18. A. G. Olabi, D. Maizak and T. Wilberforce, Sci. Total Environ., 2020, 748, 141249 CrossRef CAS PubMed.
  19. S. Dey and N. S. Mehta, Resour. Environ. Sustain., 2020, 2, 100006 Search PubMed.
  20. T. Selleri, A. D. Melas, A. Joshi, D. Manara, A. Perujo and R. Suarez-Bertoa, Catalysts, 2021, 11, 404 CrossRef CAS.
  21. S. Muhammad Farhan, W. Pan, C. Zhijian and Y. JianJun, Fuel, 2024, 355, 129364 CrossRef CAS.
  22. W. Shan and H. Song, Catal. Sci. Technol., 2015, 5, 4280–4288 RSC.
  23. H. Nazir, N. Muthuswamy, C. Louis, S. Jose, J. Prakash, M. E. Buan, C. Flox, S. Chavan, X. Shi, P. Kauranen, T. Kallio, G. Maia, K. Tammeveski, N. Lymperopoulos, E. Carcadea, E. Veziroglu, A. Iranzo and A. M. Kannan, Int. J. Hydrogen Energy, 2020, 45, 20693–20708 CrossRef CAS.
  24. W. H. Chen and C. Y. Chen, Appl. Energy, 2020, 258, 114078 CrossRef CAS.
  25. A. E. Hughes, N. Haque, S. A. Northey and S. Giddey, Resources, 2021, 10, 93 CrossRef.
  26. A. H. Laskar, M. Y. Soesanto and M. C. Liang, Environ. Sci. Technol., 2021, 55, 4378–4388 CrossRef CAS PubMed.
  27. S. Royer, D. Duprez, F. Can, X. Courtois, C. Batiot-Dupeyrat, S. Laassiri and H. Alamdari, Chem. Rev., 2014, 114, 10292–10368 CrossRef CAS PubMed.
  28. I. V. Yentekakis, A. G. Georgiadis, C. Drosou, N. D. Charisiou and M. A. Goula, Nanomaterials, 2022, 12, 1042 CrossRef CAS PubMed.
  29. J. Yuan, J. X. Mi, R. Yin, T. Yan, H. Liu, X. Chen, J. Liu, W. Si, Y. Peng, J. Chen and J. Li, ACS Catal., 2022, 12, 1024–1030 CrossRef CAS.
  30. J. Yuan, Z. Wang, J. Liu, J. Li and J. Chen, Environ. Sci. Technol., 2023, 57, 606–614 CrossRef CAS PubMed.
  31. S. N. Perevislov, Refract. Ind. Ceram., 2019, 60, 291–295 CrossRef CAS.
  32. C. R. Hsing, C. Cheng, J. P. Chou, C. M. Chang and C. M. Wei, New J. Phys., 2014, 16, 113015 CrossRef.
  33. G. Cassabois, P. Valvin and B. Gil, Phys. Rev. B, 2016, 93, 1–6 CrossRef.
  34. S. A. Tawfik, S. Ali, M. Fronzi, M. Kianinia, T. T. Tran, C. Stampfl, I. Aharonovich, M. Toth and M. J. Ford, Nanoscale, 2017, 9, 13575–13582 RSC.
  35. X. Lü, S. He, H. Lian, S. Lv, Y. Shen, W. Dong and Q. Duan, Appl. Surf. Sci., 2020, 531, 147262 CrossRef.
  36. F. Zheng, G. Zhou, Z. Liu, J. Wu, W. Duan, B. L. Gu and S. B. Zhang, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 1–5 Search PubMed.
  37. P. M. Revabhai, R. K. Singhal, H. Basu and S. K. Kailasa, J. Nanostruct. Chem., 2023, 13, 1–41 CrossRef CAS.
  38. M. Jana and R. N. Singh, Int. Mater. Rev., 2018, 63, 162–203 CrossRef CAS.
  39. R. T. Paine and C. K. Narulat, Chem. Rev., 1990, 90, 73–91 CrossRef CAS.
  40. T. H. Le, Y. Oh, H. Kim and H. Yoon, Chem. – Eur. J., 2020, 26, 6360–6401 CrossRef CAS PubMed.
  41. D. R. Lide, G. Baysinger, L. I. Berger, R. N. Goldberg, H. V. Kehiaian, K. Kuchitsu, G. Rosenblatt, D. L. Roth and D. Zwillinger, CRC Handbook of Chemistry and Physics, Internet Version 2005, CRC Press, Boca Raton, FL, 2005 Search PubMed.
  42. Q. Cai, D. Scullion, W. Gan, A. Falin, S. Zhang, K. Watanabe, T. Taniguchi, Y. Chen, E. J. Santos and L. H. Li, Sci. Adv., 2019, 5, eaav0129 CrossRef PubMed.
  43. L. Xue, B. Lu, Z. S. Wu, C. Ge, P. Wang, R. Zhang and X. D. Zhang, Chem. Eng. J., 2014, 243, 494–499 CrossRef CAS.
  44. C. Jin, F. Lin, K. Suenaga and S. Iijima, Phys. Rev. Lett., 2009, 102, 1–4 Search PubMed.
  45. J. C. Meyer, A. Chuvilin, G. Algara-Siller, J. Biskupek and U. Kaiser, Nano Lett., 2009, 9, 2683–2689 CrossRef CAS PubMed.
  46. Y. Ding, F. Torres-Davila, A. Khater, D. Nash, R. Blair and L. Tetard, MRS Commun., 2018, 8, 1236–1243 CrossRef CAS.
  47. L. H. Lu, Y. Chen, B. Gavin, Z. Hongzhou, P. Mladen and M. G. Alexey, J. Mater. Chem. C, 2011, 21, 11862–11866 RSC.
  48. B. Xiao, X. F. Yu and Y. H. Ding, RSC Adv., 2014, 4, 22688–22696 RSC.
  49. J. W. Feng, Y. J. Liu and J. X. Zhao, J. Mol. Model., 2014, 20, 2307 CrossRef PubMed.
  50. M. D. Esrafili, S. Heydari and L. Dinparast, Struct. Chem., 2019, 30, 1647–1657 CrossRef CAS.
  51. M. D. Esrafili and F. Arjomandi Rad, Vacuum, 2019, 166, 127–134 CrossRef CAS.
  52. Y. He, D. Li, W. Gao, H. Yin, F. Chen and Y. Sun, Nanoscale, 2019, 11, 21909–21916 RSC.
  53. J. Zhang, J. Tian, Q. Zhang, Y. Lu, L. Li and Y. Xu, ChemistrySelect, 2021, 6, 13609–13615 CrossRef CAS.
  54. N. L. McDougall, J. G. Partridge, R. J. Nicholls, S. P. Russo and D. G. McCulloch, Phys. Rev. B, 2017, 96, 144106 CrossRef.
  55. L. Weston, D. Wickramaratne, M. Mackoit, A. Alkauskas and C. G. Van De Walle, Phys. Rev. B, 2018, 97, 214104 CrossRef CAS.
  56. A. J. R. Payne, N. F. Xavier, G. Bauerfeldt and M. Sacchi, Phys. Chem. Chem. Phys., 2022, 24, 20426–20436 RSC.
  57. D. Q. Hoang, P. Pobedinskas, S. S. Nicley, S. Turner, S. D. Janssens, M. K. Van Bael, J. D'Haen and K. Haenen, Cryst. Growth Des., 2016, 16, 3699–3708 CrossRef CAS.
  58. D. J. Nash, D. T. Restrepo, N. S. Parra, K. E. Giesler, R. A. Penabade, M. Aminpour, D. Le, Z. Li, O. K. Farha, J. K. Harper, T. S. Rahman and R. G. Blair, ACS Omega, 2016, 1, 1343–1354 CrossRef CAS PubMed.
  59. M. K. Agusta and I. Prasetiyo, J. Phys.: Condens. Matter, 2002, 14, 2717–2744 CrossRef.
  60. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
  61. A. Tkatchenko and M. Scheffler, Phys. Rev. Lett., 2009, 102, 6–9 CrossRef PubMed.
  62. D. Vanderbilt, Phys. Rev. B: Condens. Matter Mater. Phys., 1990, 41, 7892 CrossRef PubMed.
  63. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Solid State, 1976, 13, 5188 CrossRef.
  64. T. A. Halgren and W. N. Lipscomb, Chem. Phys. Lett., 1977, 49, 225–232 CrossRef CAS.
  65. K. Momma and F. Izumi, J. Appl. Crystallogr., 2011, 44, 1272–1276 CrossRef CAS.
  66. K. Shirai, H. Dekura, Y. Mori, Y. Fujii, H. Hyodo and K. Kimura, J. Phys. Soc. Jpn., 2011, 80, 084601 CrossRef.
  67. I. Filot, Introduction to microkinetic modeling, Technische Universiteit Eindhoven, 2018 Search PubMed.
  68. M. S. Si and D. S. Xue, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 1–4 CrossRef.
  69. B. Huang and H. Lee, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 86, 1–8 Search PubMed.
  70. J. Strand, L. Larcher and A. L. Shluger, J. Phys.: Condens. Matter, 2020, 32, 055706 CrossRef CAS PubMed.
  71. K. Watanabe and T. Taniguchi, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 79, 1–4 Search PubMed.
  72. V. Surya, K. Choutipalli, K. Esackraj and E. Varathan, Appl. Surf. Sci., 2022, 602, 15 Search PubMed.
  73. J. H. Seinfeld and S. N. Pandis, Atmospheric Chemistry and Physics: From air pollution to climate change, John Wiley & Sons, 3rd edn, 2016 Search PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4cy00206g

This journal is © The Royal Society of Chemistry 2024
Click here to see how this site uses Cookies. View our privacy policy here.