Next Article in Journal
The Photocatalytic Activity of CaTiO3 Derived from the Microwave-Melting Heating Process of Blast Furnace Slag
Next Article in Special Issue
On the Origin of Raman Activity in Anatase TiO2 (Nano)Materials: An Ab Initio Investigation of Surface and Size Effects
Previous Article in Journal
Exact Solutions for Non-Isothermal Flows of Second Grade Fluid between Parallel Plates
Previous Article in Special Issue
High-Accuracy Neural Network Interatomic Potential for Silicon Nitride
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DFT Study of Adsorption Behavior of Nitro Species on Carbon-Doped Boron Nitride Nanoribbons for Toxic Gas Sensing

by
Francisco Villanueva-Mejia
1,*,
Santiago José Guevara-Martínez
2,
Manuel Arroyo-Albiter
3,
José Juan Alvarado-Flores
4 and
Adalberto Zamudio-Ojeda
5
1
Instituto Tecnológico de Pabellón de Arteaga, Carretera a la estación de Rincón de Romos, km 1, Pabellón de Arteaga 20670, Aguascalientes, Mexico
2
Department of Pharmacology, School of Exact Sciences and Engineering, University of Guadalajara, Boulevard Gral. Marcelino García Barragán 1421, Olímpica, Guadalajara 44840, Jalisco, Mexico
3
Instituto de Investigaciones Químico Biológicas, Universidad Michoacana de San Nicolás de Hidalgo, Morelia 58030, Michoacán, Mexico
4
Facultad de Ingeniería en Tecnología de la Madera, Universidad Michoacana de San Nicolás de Hidalgo, Morelia 58040, Michoacán, Mexico
5
Department of Physics, School of Exact Sciences and Engineering, University of Guadalajara, Boulevard Gral. Marcelino García Barragán 1421, Olímpica, Guadalajara 44840, Jalisco, Mexico
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(8), 1410; https://doi.org/10.3390/nano13081410
Submission received: 28 March 2023 / Revised: 14 April 2023 / Accepted: 17 April 2023 / Published: 19 April 2023
(This article belongs to the Special Issue Theoretical Calculation and Molecular Modeling of Nanomaterials)

Abstract

:
The modifications of the electronic properties on carbon-doped boron nitride nanoribbons (BNNRs) as a response to the adsorption of different nitro species were investigated in the framework of the density functional theory within the generalized gradient approximation. Calculations were performed using the SIESTA code. We found that the main response involved tuning the original magnetic behavior to a non-magnetic system when the molecule was chemisorbed on the carbon-doped BNNR. It was also revealed that some species could be dissociated through the adsorption process. Furthermore, the nitro species preferred to interact over nanosurfaces where dopants substituted the B sublattice of the carbon-doped BNNRs. Most importantly, the switch on the magnetic behavior offers the opportunity to apply these systems to fit novel technological applications.

1. Introduction

Toxic gases are major contributors to ambient air pollution and pose a serious threat to both human health and the environment. Rapid detection of these gases is crucial, and in some cases, they need to be neutralized to prevent harm, becoming harmless [1]. The use of low-dimensional gas-sensor systems for indoor and outdoor environmental monitoring is becoming increasingly prevalent [2]. Nitrogen-containing gases play a critical role in protecting the environment, promoting human health, and supporting biomedicine and agriculture [3]. Toxic gases, such as CO and NO, are particularly harmful and can cause severe health problems, including reduced oxygen delivery to vital organs, skin issues, and respiratory complications [1]. Exposure to even low concentrations of nitrogen dioxide in enclosed spaces can lead to respiratory complications and even death [4]. Additionally, ammonia can harm human health through tropospheric ozone formation, acidification of rain, and eutrophication [5]. Recent advancements in gas-sensor technology have allowed for the detection of NO2 and NH3 at the level of a single gas molecule on a thin-film material [6]. Theoretical analyses have demonstrated the effectiveness of graphene-based systems in detecting nitrogen-containing gases, and these promising results have prompted further experimentation and application of low-dimensional gas sensors [7,8]. Theoretical calculations have also shown that boron nitride systems, which are analog materials with similarities to graphene, can be utilized for gas sensing. Therefore, conducting theoretical calculations to determine the gas sensing properties of boron nitride systems could be a significant milestone in discovering a more effective sensing material.
The desire to downsize gas sensors while maintaining high sensitivity, selectivity, rapid response, and low cost has led scientists to investigate gas sensing systems based on novel nanomaterials [9,10,11]. Graphene, a pure two-dimensional crystal made of a sp2-hybridized carbon, has opened up a vast area of research exploring related low-dimensional geometries due to its exceptional properties, including high chemical stability, thermal stability, surface–volume ratio, carrier mobility, and fast response time [12,13,14,15,16]. Both experimental and theoretical investigations have been conducted on graphene’s applications in gas sensing and biosensing [17,18,19,20,21]. Such studies have sparked an increased interest in exploring related inorganic two-dimensional geometries for further research.
Hexagonal boron nitride sheets are another geometry that has recently gained significant research interest due to its analogical similarity to graphene [22,23,24]. Similar to graphene nanoribbons, boron nitride nanoribbons (BNNRs) can also be classified into zigzag and armchair configurations [25]. Previous studies have described the band gap behavior for each configuration of BNNR, and the introduction of functionalization and substitutional carbon doping has modulated their band gap, tuning their electronic properties from insulators to semiconducting materials [26,27]. The introduction of doped systems not only modifies the electronic properties of the boron nitride nanoribbon but also affects its intrinsic reactivity, describing the more interactive areas on the material and explaining the preferred interaction mechanism [28]. This makes BNNRs a promising material for novel technological applications, such as gas sensing. Theoretical DFT calculations have demonstrated efficient sensing capabilities of BNNRs for biosensing [29], alkaloid sensing [30], PH3 gas sensing [31], and NO reduction [32]. Therefore, addressing theoretical calculations may provide interesting results before developing an efficient gas sensor.
Some of the toxic gas molecules (NH3, NH4, CO, and NO2) show low sensitivity towards pristine GNR, which limits its potential use in the detection of individual gas molecules [33,34,35]. However, various reports discovered strong interactions between doped graphene and the same mentioned four gases, enhancing gas sensors [7,36,37,38]. Metal-doped graphene has also been shown to exhibit high sensitivity towards O2 adsorption [39], while defected graphene has been reported to have higher interactions with H2S than pristine graphene [40]. Furthermore, metal-decorated graphene has been found to have a high sensing performance towards NO, H2S, and HCN molecules, as the gas molecules bind firmly to the carbon structures [41,42].
In this study, we utilized first-principles methods based on density functional theory to investigate the electronic response of carbon-doped boron nitride nanoribbons upon exposure to toxic gas molecules, including NO2, NH3, and NH4. By analyzing the changes in their electronic properties, we evaluated the sensitivity of the carbon-doped BNNR sensor. Our findings suggest that these systems have great potential in the development of ultrahigh-sensitivity gas sensors.

2. Computation Details

The SIESTA code was used to perform all calculations for this work, with the projector augmented-wave method (PAW) describing the interaction between ionic cores and valence electrons [43]. The Perdew, Burke, and Ernzerhof parameterization (PBE) of the generalized gradient approximation (GGA) was used for the exchange-correlation functional [44]. The plane-wave basis-set cutoff energy was set to 450 eV, and the conjugate gradient method was used until the forces on each atom were lower than 0.01 eV/Å. The orthorhombic unit cell dimensions on the non-periodic axis were large enough to avoid interactions between images and edged hydrogen atoms, passivating the N and B atoms on the edges. The unit cell dimensions were optimized through the periodic dimension (along the y-axis (Figure S1)) until the cell total stress was close to a minimum, while a space of ~10 Å along the non-periodic directions was fixed. A Monkhorst-Pack grid of 1 × 20 × 1 k-points was used to sample the Brillouin zone in the reciprocal space, with a plane-wave cutoff energy of 450 eV. Spin-polarized calculations were performed for the carbon-doped BNNR nanostructures, and the total energy per unit cell was more negative in the spin-polarized approximation than in the non-polarized approximation (more than 2.27 eV of negative difference); these significant differences are in agreement with the literature. The results will be discussed in terms of the spin-polarized solution, with the total energy per unit cell being more negative than the non-polarized approximation. Therefore, nanoribbons with magnetic behavior will be analyzed (M-BNNR). In spin-restricted or -unrestricted calculations, closed shell systems are characterized by having doubly occupied molecular orbitals (restricted case), whereas in open shell systems, the spin degree of freedom makes it possible to obtain a complete set of orbitals for α-spin electrons and another set for β-spin electrons. The calculations were also used to describe systems that can be periodic in one dimension, with good agreement with experimental data.

3. Results

3.1. System Description

Firstly, the M-BNNRs considered in this work were built by passivating the edge N and B atoms with one hydrogen atom attached per edge atom. The selected size of these nanoribbons was 12 × 2, following a convention of size M × N, with M dimer lines along the width of the M-BNNR and N being the number of six-member rings along the periodic direction. The optimized geometries resulted in average bond lengths of 1.43, 1.45, and 1.08 Å for the pairs C-N, B-N, and C-H, respectively. Figure 1 provides the schematic representation of 12 × 2 M-BNNRs passivated by hydrogen atoms. In Figure 1a, the optimized structure of the nanoribbon shows a C atom substituting a N atom on the edge, whereas in Figure 1b, a C atom substitutes a N atom at the center of the nanoribbon. With a constant concentration of dopants equal to 1.78% and only one carbon atom on the BNNR, the relative position of the substitute atom shown in Figure 1 determines the identification for each possible arrangement, which is named P-M-BNNR, where P represents the relative position of the carbon atom on the nanoribbon (P = E or C), E indicating the edge and C the center of the M-BNNR.
Spin-polarized and non-spin-polarized solutions of the Kohn–Sham equations were analyzed, finding significant differences in the proposed systems. The doping effect in the system causes an unpaired number of electrons that, in contrast to a pristine BNNR, may be the reason for these significant differences. A previous study by Du et al. reported the spontaneous magnetization of BNNRs with a single C substitution [45].

3.2. System Electronic Properties

In this section, we discuss the band structure and total density of states of M-BNNRs, as presented in Figure 2. To obtain these figures, a Monkhorst-Pack grid with twice the density employed for energy calculations was used, and an integration of the Brillouin zone was performed. To improve the convergence of the integrals, a Gaussian smearing with a width of 0.01 eV for the levels was employed.
Figure 2b,c clearly show that the substitution of a N atom with a C atom results in an unoccupied band with a spin up contribution that shifts downwards from its original position near the Fermi level. In contrast, a pristine BNNR system exhibits insulator properties, with a band gap of approximately 4.5 eV, as shown in Figure 2a.
The magnetic behavior observed in the M-BNNRs can be attributed to the contribution of the C atom, which results in an unpaired number of electrons compared to the pristine system. Specifically, the substitution of a N atom with a C atom results in one less electron, which gives rise to a system with magnetic behavior.

3.3. Energetic Stability

In order to assess the energetic stability of the proposed systems, the cohesive energy (CE) was computed, and the results are reported in Table 1. The cohesive energy is defined as the amount of energy required to separate isolated atomic species of a solid, and a positive CE indicates a bound and stable state. Higher values of CE indicate more stable states. The computed CE values for a cubic BN and h-BN sheet are reported to be 7.12 and 7.09 eV atom−1, respectively, which are in good agreement with the size of the analyzed nanoribbon. It is theoretically expected that the cohesive energy decreases as the system size is reduced. The molar Gibbs free energy (δG) is also displayed in parenthesis in Table 1 and is given by
δ G = E x + i = 1 n X i μ i
where E(x) corresponds to the binding energy per atom and Xi is the molar fraction of the proper atomic species (H, B, C, N), which satisfies ∑Xi = 1. The chemical potential is approximated by the binding energy per atom of the ground state of H2 and N2 in the singlet state, the ground state of B2 in its triplet state, and the cohesive energy per atom of a graphene sheet. The energetic stability trend observed in the computed systems shows that doping at the edge is more energetically favorable than at the center of the ribbon, with the trend being pristine BNNR > E_M-BNNR > C_M-BNNR, indicating that the systems become increasingly energetically stable with dopant atoms on the edge rather that at the center of the nanoribbon.

3.4. Interaction of Toxic Gases on M-BNNRs

In this section, to further explore the potential of M-BNNRs in gas-sensing applications, the response to different molecules of interest was analyzed. While pristine graphene nanoribbons can detect these toxic molecules, doped graphene nanoribbons have shown improved gas-sensing capabilities due to the presence of dopant atoms. In the case of M-BNNRs, these toxic molecules can interact with the possible sites of the nanoribbons, leading to a corresponding change in the magnetic electronic behavior. This sensitivity to the adsorption of a single toxic molecule suggests that M-BNNRs could serve as highly responsive material for the detection of these harmful gases.
The molecular electrostatic potential analysis has provided additional insights into the interaction between the molecules of interest and the M-BNNRs. Specifically, the substitution of the C atom on the proposed site activates the surrounding N atoms, resulting in a more positive and larger site for interaction with the molecules. To study the interaction process, various initial configurations were proposed, taking into account the orientation and location of the molecules with respect to the activated area.
The interaction energies, optimized complexes, and corresponding modifications in the electronic properties were analyzed to understand the interaction process between the molecules and M-BNNRs. This information can be used to synthesize M-BNNRs that exhibit a high sensitivity toward the toxic molecules of interest and potentially develop them for use in gas-sensing applications.
Table 2 summarizes the interaction energies between the M-BNNRs and the analyzed toxic molecules. The interaction energy (IE) was calculated by using the following expression:
I E = E c o m p l e x E P _ M B N N R + E M
where Ecomplex is the computed energy of the interaction system, EP_M-BNNR represents the energy of the optimized magnetic nanoribbon, and EM designates the energy of the toxic molecule of interest, all of them in gas phase.
Table 2 suggests that the edge of the ribbon is more reactive and provides a more attractive site for the interaction of the studied molecules. The negative interaction energies indicate that the molecules are adsorbed on the surface of the M-BNNRs, and this adsorption process can modify the electronic properties of the system. The larger negative interaction energies observed for the E_M-BNNRs suggest that the substitution for the C atom at the center of the ribbon results in a less favorable site for interaction compared to the edge one. The modification of the electronic properties due to the adsorption of the studied molecules will be discussed in the next section.

4. Discussion

The final configurations of the interaction process between the systems are presented in Figure 3. Figure 3a,b demonstrate that the NO2 molecule can directly interact with the C dopant atom in the M-BNNR. These interaction complexes bind the N atom of the NO2 molecule to the C atom on the M-BNNR. The binding distances between the C-N atoms are reported to be 1.49 and 1.52 Å for the edge and center M-BNNR, respectively. In addition, a curve deformation on the M-BNNR for the C atom is observed, which may be due to the small size of the M-BNNR, where the cohesive energy is not sufficient to maintain the original six-member ring form.
Figure 3c,d illustrate the interaction of ammonium with the M-BNNR. In both systems, one H atom from the ammonium molecule binds to the C atom on the M-BNNR, resulting in an average bond distance of 1.12 Å between the C-H pair. Additionally, these calculations reveal the deprotonation for the remaining portion of the ammonium molecule, as evidenced by the reported distance between the bonded H atom and the nearest N atom. The separation distances between N-H pairs were found to be 2.34 and 2.21 Å for the edge and center M-BNNR, respectively. A curve deformation on the C atom can also be observed, although these deformations are less pronounced than the ones previously mentioned.
Figure 3e showcases the interaction between ammonia and E_M-BNNR, which represents the final and most favorable interaction observed in this study. The bonding between the species results in the most evident deformation. Initial configurations of the ammonia molecule oriented on top of the C atom were unresponsive, prompting a shift to the three nearest B atoms. The superior B atom exhibited the strongest attraction to the ammonia molecule, forming bonds described by B-N pairs with a separation of 1.62 Å. Similar analyses were performed on the C_M-BNNR, exploring initial configurations of the ammonia molecule on top of the C atom, the nearest B atoms, and even on the nearest N atoms. However, the interaction process was found to be unfavorable in all cases. In some instances, repulsion processes were observed at the nearest N atoms. Further studies are required to determine why the interaction process on this site is not favorable, as represented in Figure 3d.
Figure 4 shows the modifications in the magnetic electronic properties of the M-BNNRs following the interaction process. In all cases, with the exception of Figure 4f, the original magnetic behavior of the M-BNNR transitioned to a non-magnetic behavior after bonding with a molecule such as nitrogen dioxide or ammonia, a part of the original molecule, as in the case of ammonium. Following bonding onto the M-BNNR, the corresponding graphs describing the electronic properties now exhibit an electron pair in each band in both the band-structure graph and the local-density-of-states graph. Only Figure 4f retains an unpaired number of electrons in each band-structure and local-density-of-states graph.
The results of the present study demonstrated a stronger interaction between the M-BNNR and NO2 compared to a similar interaction reported by Bagheri and Peyghan, where they observed interaction separation distances of approximately 2 Å between a nanocone and the same molecule [46]. Additionally, a similar deformation behavior was observed in the study conducted by Esrafili in 2019, where the NO2 adsorption on a boron and nitrogen-doped graphene nanoribbon resulted in a slight curvature directly where the molecule interacted. The separation distance for their system was larger than 1.66 Å [37]. Moreover, the ammonium interaction observed in the present study involved a partial deprotonation of the molecule resulting in the separation of the molecule into ammonia and a hydrogen atom bonded to the M-BNNR. To our knowledge, this type of interaction has not yet been reported in the literature. Finally, the NH3 interaction on top of a zigzag graphene nanoribbon converged with a separation of 3.03 Å reported by E. Salih et al., more than double the separation reported in this study [47].
The effect of the carbon-dopant relative position in the same sublattice was also examined, while maintaining a dopant concentration of 1.78% of C atoms. Our findings are consistent with previous reports, which suggest that the edges of the M-BNNRs exhibit a higher interaction with nitrogenated molecules than the center part.

5. Conclusions

In this study, we conducted a first-principles investigation of the electronic properties of M-BNNRs using DFT. This study presented an investigation of the interaction between M-BNNRs and three nitrogenated molecules, namely nitrogen dioxide, ammonium, and ammonia.
Our study indicates that M-BNNRs can be utilized for the detection of nitrogenated molecules. By analyzing the shift from magnetic to non-magnetic behavior due to the interaction of these molecules with the M-BNNRs, we took the first step towards designing a low-dimensional gas sensor. The order of interaction in terms of the energy involved in the interaction of these nitro molecules is as follows: ammonium > ammonia > nitrogen dioxide.
To gain a better understanding of these processes, future research could investigate mixtures of these molecules, including their interactions with both the dopant atom and the delocalized π-systems.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/nano13081410/s1, Figure S1: Schematic representation of a M-BNNR with 4 member rings on the periodic direction; Optimized coordinates of individual systems.

Author Contributions

Conceptualization, F.V.-M. and S.J.G.-M.; methodology, F.V.-M.; validation, F.V.-M., M.A.-A., and A.Z.-O.; formal analysis, F.V.-M.; investigation, F.V.-M. and S.J.G.-M.; resources, J.J.A.-F., M.A.-A., and A.Z.-O.; data curation, A.Z.-O.; writing—original draft preparation, F.V.-M. and S.J.G.-M.; writing—review and editing, J.J.A.-F., M.A.-A., and A.Z.-O.; visualization, F.V.-M. and S.J.G.-M.; supervision, F.V.-M. and A.Z.-O.; project administration, F.V.-M.; funding acquisition, J.J.A.-F., M.A.-A., and A.Z.-O. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on a reasonable request from the corresponding author.

Acknowledgments

We are grateful for the support provided by the Centro Universitario de Ciencias Exactas e Ingenierías of the Universidad de Guadalajara for this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ramanathan, A.A. Defect Functionalization of MoS2 nanostructures as toxic gas sensors: A review. IOP Conf. Ser. Mater. Sci. Eng. 2018, 305, 012001. [Google Scholar] [CrossRef]
  2. Tsujita, W.; Yoshino, A.; Ishida, H.; Moriizumi, T. Gas sensor network for air-pollution monitoring. Sens. Actuators B Chem. 2005, 110, 304–311. [Google Scholar] [CrossRef]
  3. Kumar, V.; Azhikodan, D.; Roy, D.R. 2D Sb2C3 monolayer: A promising material for the recyclable gas sensor for environmentally toxic nitrogen-containing gases (NCGs). J. Hazard. Mater. 2021, 405, 124168. [Google Scholar] [CrossRef] [PubMed]
  4. Chen, T.-M.; Kuschner, W.G.; Gokhale, J.; Shofer, S. Outdoor Air Pollution: Nitrogen Dioxide, Sulfur Dioxide, and Carbon Monoxide Health Effects. Am. J. Med. Sci. 2007, 333, 249–256. [Google Scholar] [CrossRef]
  5. Winiwarter, W.; Klimont, Z. The role of N-gases (N2O, NOx, NH3) in cost-effective strategies to reduce greenhouse gas emissions and air pollution in Europe. Curr. Opin. Environ. Sustain. 2011, 3, 438–445. [Google Scholar] [CrossRef]
  6. He, Q.; Zeng, Z.; Yin, Z.; Li, H.; Wu, S.; Huang, X.; Zhang, H. Fabrication of Flexible MoS2 Thin-Film Transistor Arrays for Practical Gas-Sensing Applications. Small 2012, 8, 2994–2999. [Google Scholar] [CrossRef]
  7. Jappor, H.R.; Khudair, S.A.M. Electronic Properties of Adsorption of CO, CO2, NH3, NO, NO2 and SO2 on Nitrogen Doped Graphene for Gas Sensor Applications. Sens. Lett. 2017, 15, 432–439. [Google Scholar] [CrossRef]
  8. Harada, N.; Sato, S. Electronic properties of NH4-adsorbed graphene nanoribbon as a promising candidate for a gas sensor. AIP Adv. 2016, 6, 055023. [Google Scholar] [CrossRef]
  9. Schedin, F.; Geim, A.K.; Morozov, S.V.; Hill, E.W.; Blake, P.; Katsnelson, M.I.; Novoselov, K.S. Detection of individual gas molecules adsorbed on graphene. Nat Mater. 2007, 6, 652–655. [Google Scholar] [CrossRef]
  10. Lu, G.; Ocola, L.E.; Chen, J. Reduced graphene oxide for room-temperature gas sensors. Nanotechnology 2009, 20, 445502. [Google Scholar] [CrossRef]
  11. Dan, Y.; Lu, Y.; Kybert, N.J.; Luo, Z.; Johnson, A.T.C. Intrinsic Response of Graphene Vapor Sensors. Nano Lett. 2009, 9, 1472–1475. [Google Scholar] [CrossRef]
  12. Geim, A.K.; Novoselov, K.S. The rise of graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  13. Zhu, J.; Yang, D.; Yin, Z.; Yan, Q.; Zhang, H. Graphene and Graphene-Based Materials for Energy Storage Applications. Small 2014, 10, 3480–3498. [Google Scholar] [CrossRef] [PubMed]
  14. Galashev, A.E.; Rakhmanova, O.R. Mechanical and thermal stability of graphene and graphene-based materials. Physics-Uspekhi 2014, 57, 970–989. [Google Scholar] [CrossRef]
  15. Gan, T.; Hu, S. Electrochemical sensors based on graphene materials. Microchim. Acta 2011, 175, 1–19. [Google Scholar] [CrossRef]
  16. Fu, X.-W.; Liao, Z.-M.; Zhou, Y.-B.; Wu, H.-C.; Bie, Y.-Q.; Xu, J.; Yu, D.-P. Graphene/ZnO nanowire/graphene vertical structure based fast-response ultraviolet photodetector. Appl. Phys. Lett. 2012, 100, 223114. [Google Scholar] [CrossRef]
  17. Tian, W.; Liu, X.; Yu, W. Research Progress of Gas Sensor Based on Graphene and Its Derivatives: A Review. Appl. Sci. 2018, 8, 1118. [Google Scholar] [CrossRef]
  18. Varghese, S.S.; Lonkar, S.; Singh, K.K.; Swaminathan, S.; Abdala, A. Recent advances in graphene based gas sensors. Sens. Actuators B Chem. 2015, 218, 160–183. [Google Scholar] [CrossRef]
  19. Yuan, W.; Shi, G. Graphene-based gas sensors. J. Mater. Chem. A Mater. 2013, 1, 10078. [Google Scholar] [CrossRef]
  20. Cruz-Martínez, H.; Rojas-Chávez, H.; Montejo-Alvaro, F.; Peña-Castañeda, Y.A.; Matadamas-Ortiz, P.T.; Medina, D.I. Recent Developments in Graphene-Based Toxic Gas Sensors: A Theoretical Overview. Sensors 2021, 21, 1992. [Google Scholar] [CrossRef] [PubMed]
  21. Abergel, D.S.L.; Apalkov, V.; Berashevich, J.; Ziegler, K.; Chakraborty, T. Properties of graphene: A theoretical perspective. Adv. Phys. 2010, 59, 261–482. [Google Scholar] [CrossRef]
  22. Paine, R.T.; Narula, C.K. Synthetic routes to boron nitride. Chem. Rev. 1990, 90, 73–91. [Google Scholar] [CrossRef]
  23. Liu, L.; Feng, Y.P.; Shen, Z.X. Structural and electronic properties of h -BN. Phys. Rev. B 2003, 68, 104102. [Google Scholar] [CrossRef]
  24. Ooi, N.; Rajan, V.; Gottlieb, J.; Catherine, Y.; Adams, J.B. Structural properties of hexagonal boron nitride. Model. Simul. Mater. Sci. Eng. 2006, 14, 515–535. [Google Scholar] [CrossRef]
  25. Barone, V.; Peralta, J.E. Magnetic Boron Nitride Nanoribbons with Tunable Electronic Properties. Nano Lett. 2008, 8, 2210–2214. [Google Scholar] [CrossRef]
  26. Liu, N.; Liu, J.B.; Gao, G.Y.; Yao, K.L. Carbon doping induced giant low bias negative differential resistance in boron nitride nanoribbon. Phys. Lett. A 2014, 378, 2217–2221. [Google Scholar] [CrossRef]
  27. Park, C.H.; Louie, S.G. Energy Gaps and Stark Effect in Boron Nitride Nanoribbons. Nano Lett. 2008, 8, 2200–2203. [Google Scholar] [CrossRef]
  28. Villanueva-Mejia, F.; Navarro-Santos, P.; Rodríguez-Kessler, P.L.; Herrera-Bucio, R.; Rivera, J.L. Reactivity of atomically functionalized C-doped boron nitride nanoribbons and their interaction with organosulfur compounds. Nanomaterials 2019, 9, 452. [Google Scholar] [CrossRef]
  29. Dabhi, S.D.; Roondhe, B.; Jha, P.K. Nucleobases-decorated boron nitride nanoribbons for electrochemical biosensing: A dispersion-corrected DFT study. Phys. Chem. Chem. Phys. 2018, 20, 8943–8950. [Google Scholar] [CrossRef]
  30. Roondhe, B.; Dabhi, S.D.; Jha, P.K. Sensing properties of pristine boron nitride nanostructures towards alkaloids: A first principles dispersion corrected study. Appl. Surf. Sci. 2018, 441, 588–598. [Google Scholar] [CrossRef]
  31. Srivastava, P.; Jaiswal, N.K.; Sharma, V. First-principles investigation of armchair boron nitride nanoribbons for sensing PH3 gas molecules. Superlattices Microstruct 2014, 73, 350–358. [Google Scholar] [CrossRef]
  32. Li, Y.; Nie, J.; Gao, D.; Zhao, S.; Zhang, Y.; Bian, B.; Guo, Z.; Huang, Y.; Fang, Y.; Tang, C. Hexagonal boron nitride nanoribbon as a novel metal-free catalyst for high-efficiency NO reduction to NH3. Fuel 2023, 339, 126943. [Google Scholar] [CrossRef]
  33. Yoon, H.J.; Jun, D.H.; Yang, J.H.; Zhou, Z.; Yang, S.S.; Cheng, M.M.-C. Carbon dioxide gas sensor using a graphene sheet. Sens. Actuators B Chem. 2011, 157, 310–313. [Google Scholar] [CrossRef]
  34. Tang, S.; Cao, Z. Adsorption of nitrogen oxides on graphene and graphene oxides: Insights from density functional calculations. J. Chem. Phys. 2011, 134, 044710. [Google Scholar] [CrossRef]
  35. Lin, X.; Ni, J.; Fang, C. Adsorption capacity of H2O, NH3, CO, and NO2 on the pristine graphene. J. Appl. Phys. 2013, 113, 034306. [Google Scholar] [CrossRef]
  36. Hosseinian, A.; Asadi, Z.; Edjlali, L.; Bekhradnia, A.; Vessally, E. NO2 sensing properties of a borazine doped nanographene: A DFT study. Comput Theor Chem. 2017, 1106, 36–42. [Google Scholar] [CrossRef]
  37. Esrafili, M.D. Boron and nitrogen co-doped graphene nanosheets for NO and NO2 gas sensing. Phys. Lett. A 2019, 383, 1607–1614. [Google Scholar] [CrossRef]
  38. Zheng, Z.; Wang, H. Different elements doped graphene sensor for CO2 greenhouse gases detection: The DFT study. Chem. Phys. Lett. 2019, 721, 33–37. [Google Scholar] [CrossRef]
  39. Nasehnia, F.; Seifi, M. Adsorption of molecular oxygen on VIIIB transition metal-doped graphene: A DFT study. Mod. Phys. Lett. B 2014, 28, 1450237. [Google Scholar] [CrossRef]
  40. Borisova, D.; Antonov, V.; Proykova, A. Hydrogen sulfide adsorption on a defective graphene. Int. J. Quantum. Chem. 2013, 113, 786–791. [Google Scholar] [CrossRef]
  41. Tang, Y.; Liu, Z.; Shen, Z.; Chen, W.; Ma, D.; Dai, X. Adsorption sensitivity of metal atom decorated bilayer graphene toward toxic gas molecules (CO, NO, SO2 and HCN). Sens. Actuators B Chem. 2017, 238, 182–195. [Google Scholar] [CrossRef]
  42. Bo, Z.; Guo, X.; Wei, X.; Yang, H.; Yan, J.; Cen, K. Density functional theory calculations of NO2 and H2S adsorption on the group 10 transition metal (Ni, Pd and Pt) decorated graphene. Phys. E Low Dimens. Syst. Nanostruct. 2019, 109, 156–163. [Google Scholar] [CrossRef]
  43. Soler, J.M.; Artacho, E.; Gale, J.D.; García, A.; Junquera, J.; Ordejón, P.; Sánchez-Portal, D. The SIESTA method for ab initio order- N materials simulation. J. Phys. Condens. Matter. 2002, 14, 2745–2779. [Google Scholar] [CrossRef]
  44. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  45. Du, A.J.; Smith, S.C.; Lu, G.Q. First-principle studies of electronic structure and C-doping effect in boron nitride nanoribbon. Chem. Phys. Lett. 2007, 447, 181–186. [Google Scholar] [CrossRef]
  46. Bagheri, Z.; Peyghan, A.A. DFT study of NO2 adsorption on the AlN nanocones. Comput. Theor. Chem. 2013, 1008, 20–26. [Google Scholar] [CrossRef]
  47. Salih, E.; Ayesh, A.I. Enhancing the Sensing Performance of Zigzag Graphene Nanoribbon to Detect NO, NO2, and NH3 Gases. Sensors 2020, 20, 3932. [Google Scholar] [CrossRef]
Figure 1. Schematic representations of (a) E_M-BNNR and (b) C_M-BNNR. The green sphere represents the B atom, grey sphere N atom, brown sphere C atom, and pink sphere H atom.
Figure 1. Schematic representations of (a) E_M-BNNR and (b) C_M-BNNR. The green sphere represents the B atom, grey sphere N atom, brown sphere C atom, and pink sphere H atom.
Nanomaterials 13 01410 g001
Figure 2. Electronic properties of the analyzed systems: (a) pristine BNNR, (b) E_M-BNNR, and (c) C_M-BNNR. Graphs represent the band structure with the corresponding total density of states.
Figure 2. Electronic properties of the analyzed systems: (a) pristine BNNR, (b) E_M-BNNR, and (c) C_M-BNNR. Graphs represent the band structure with the corresponding total density of states.
Nanomaterials 13 01410 g002
Figure 3. Schematic representation of the most stable interaction complex between M-BNNR and toxic molecules: (a) NO2 on the edge, (b) NO2 at the center, (c) NH4 on the edge, (d) NH4 at the center, (e) NH3 on the edge, and (f) NH3 at the center.
Figure 3. Schematic representation of the most stable interaction complex between M-BNNR and toxic molecules: (a) NO2 on the edge, (b) NO2 at the center, (c) NH4 on the edge, (d) NH4 at the center, (e) NH3 on the edge, and (f) NH3 at the center.
Nanomaterials 13 01410 g003
Figure 4. Electronic properties after the adsorption process of NO2 for (a) E_M-BNNR and (b) C_M-BNNR, NH3 for (c) E_M-BNNR and (d) C_M-BNNR, and NH4 for (e) E_M-BNNR and (f) C_M-BNNR. Graphs represent band structure with the corresponding local density of states.
Figure 4. Electronic properties after the adsorption process of NO2 for (a) E_M-BNNR and (b) C_M-BNNR, NH3 for (c) E_M-BNNR and (d) C_M-BNNR, and NH4 for (e) E_M-BNNR and (f) C_M-BNNR. Graphs represent band structure with the corresponding local density of states.
Nanomaterials 13 01410 g004aNanomaterials 13 01410 g004b
Table 1. Cohesive energy per atom (and Gibbs free energy) in eV of the pristine BNNR and M-BNNR.
Table 1. Cohesive energy per atom (and Gibbs free energy) in eV of the pristine BNNR and M-BNNR.
M × NPristineE_M-BBNRC_M-BNNR
12 × 26.32 (−5.5)6.30 (−5.39)6.29 (−5.38)
Table 2. Interaction energy of the most stable complex between the analyzed toxic molecules on the M-BNNR, measured in kcal/mol.
Table 2. Interaction energy of the most stable complex between the analyzed toxic molecules on the M-BNNR, measured in kcal/mol.
M-BNNRNO2NH4NH3
E_M-BNNR−50.04−87.09−80.47
C_M-BNNR−44.23−85.02NR
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

Villanueva-Mejia, F.; Guevara-Martínez, S.J.; Arroyo-Albiter, M.; Alvarado-Flores, J.J.; Zamudio-Ojeda, A. DFT Study of Adsorption Behavior of Nitro Species on Carbon-Doped Boron Nitride Nanoribbons for Toxic Gas Sensing. Nanomaterials 2023, 13, 1410. https://doi.org/10.3390/nano13081410

AMA Style

Villanueva-Mejia F, Guevara-Martínez SJ, Arroyo-Albiter M, Alvarado-Flores JJ, Zamudio-Ojeda A. DFT Study of Adsorption Behavior of Nitro Species on Carbon-Doped Boron Nitride Nanoribbons for Toxic Gas Sensing. Nanomaterials. 2023; 13(8):1410. https://doi.org/10.3390/nano13081410

Chicago/Turabian Style

Villanueva-Mejia, Francisco, Santiago José Guevara-Martínez, Manuel Arroyo-Albiter, José Juan Alvarado-Flores, and Adalberto Zamudio-Ojeda. 2023. "DFT Study of Adsorption Behavior of Nitro Species on Carbon-Doped Boron Nitride Nanoribbons for Toxic Gas Sensing" Nanomaterials 13, no. 8: 1410. https://doi.org/10.3390/nano13081410

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