Next Article in Journal
Stable Fluorescence of Eu3+ Complex Nanostructures Beneath a Protein Skin for Potential Biometric Recognition
Next Article in Special Issue
Extensive Broadband Near-Infrared Emissions from GexSi1−x Alloys on Micro-Hole Patterned Si(001) Substrates
Previous Article in Journal
Carbon and Neon Ion Bombardment Induced Smoothing and Surface Relaxation of Titania Nanotubes
Previous Article in Special Issue
Photocatalytic Degradation of Tobacco Tar Using CsPbBr3 Quantum Dots Modified Bi2WO6 Composite Photocatalyst
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nitrogen-Containing Gas Sensing Properties of 2-D Ti2N and Its Derivative Nanosheets: Electronic Structures Insight

1
Key Laboratory of Liquid-Solid Structural Evolution and Processing of Materials (Ministry of Education), Shandong University, Jinan 250061, China
2
College of Industry and Commerce, Shandong Management University, Jinan 250357, China
3
Condensed Matter Theory, Department of Physics and Astronomy, Ångström Laboratory, Uppsala University, 75120 Uppsala, Sweden
4
Department of Materials Science and Engineering, KTH Royal Institute of Technology, 10044 Stockholm, Sweden
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(9), 2459; https://doi.org/10.3390/nano11092459
Submission received: 12 August 2021 / Revised: 8 September 2021 / Accepted: 14 September 2021 / Published: 21 September 2021
(This article belongs to the Special Issue Nanotechnologies and Nanomaterials: Selected Papers from CCMR)

Abstract

:
In this work, the potentials of two-dimensional Ti2N and its derivative nanosheets Ti2NT2(T=O, F, OH) for some harmful nitrogen-containing gas (NCG) adsorption and sensing applications have been unveiled based on the quantum-mechanical Density Functional Theory calculations. It is found that the interactions between pure Ti2N and NCGs (including NO, NO2, and NH3 in this study) are very strong, in which NO and NO2 can even be dissociated, and this would poison the substrate of Ti2N monolayer and affect the stability of the sensing material. For the monolayer of Ti2NT2(T=O, F, OH) that is terminated by functional groups on surface, the adsorption energies of NCGs are greatly reduced, and a large amount of charges are transferred to the functional group, which is beneficial to the reversibility of the sensing material. The significant changes in work function imply the good sensitivity of the above mentioned materials. In addition, the fast response time further consolidates the prospect of two-dimensional Ti2NT2 as efficient NCGs’ sensing materials. This theoretical study would supply physical insight into the NCGs’ sensing mechanism of Ti2N based nanosheets and help experimentalists to design better 2-D materials for gas adsorption or sensing applications.

1. Introduction

Industrial pollutant emissions account for a large share of environmental pollution, and one of the most harmful is nitrogen-containing gas (NCGs), such as NO, NO2, and NH3. At the same time, the nitrogen-containing gases themselves are generally toxic and can damage the respiratory cells of humans and animals [1,2,3,4]. Therefore, there is an urgent need to explore efficient and alternative materials for the sensing or capturing of these gases. Although some current solid-state sensors can also detect toxic gases, there is still much room for improvement due to the increasingly high demand for sensitivity, selectivity, stability, and other factors [5,6]. Two-dimensional (2-D) materials have large contact areas with gas and special electronic properties; thus, their interactions with gas molecules have been widely studied, especially in gas sensing and capturing [7]. For example, Zhang et al. studied the adsorption behavior between four modified graphene substrates and small gas molecules (CO, NO, NO2, and NH3) and significantly improved the performance of the gas sensor by introducing doping or defects [8]. Liu et al. found that the GeSe monolayer had potential for NH3 gas capture and storage, as well as for the detection and catalysis of SO2 and NO2. The optical properties of this GeSe monolayer can also be tuned as a candidate material for optical sensors by adsorbing different small molecules [9]. Li et al. compared the structures and electronic properties of pure and two vacancy-defected MoS2 nanosheets with NO adsorption. Defects can effectively improve the NO-sensing performance [10].
Transition metal carbide or nitride MXenes are also one family of promising gas-sensing materials [11,12], where M is a transition metal and X is carbon or nitrogen. The general formula of MXenes is written as MnXn−1. During the preparation process in an aqueous solution containing fluoride ions, there may be some O, F, and OH functional groups formed [13,14,15]. This type of MXene with functional groups is written as MnXn−1Tn(T=O, F, OH). MXenes are also widely reported in energy storage, electrode materials, ion batteries, and the spintronic and optical devices, due to their unique structures and properties [16,17,18,19,20,21,22,23,24].
After our review of previous literature, we found that some carbide-based MXenes have the potential to be used as sensors for molecules, such as amino acids and nucleobase [25,26]. With less exfoliation energy compared with other types, the synthesizability of the nitride MXenes is good [27], while there are fewer studies on their use in sensing NCGs [28,29,30,31,32]. In order to obtain a theoretical understanding of the nitride MXenes in terms of their sensing or capturing performance towards nitrogen-containing gas (NCGs), in this work, we conducted a detailed computational study on the adsorption behaviors of NCGs (including NO, NO2, and NH3) of Ti2N monolayer with bare surface and terminated with O, F, and OH functional groups. Density Functional Theory–based first-principles calculations were performed to unveil the adsorption structures, charge transfer, band structures, density of states, and work functions to stimulate and inspire the deep mechanism investigations.

2. Methods

DFT calculations were employed by using the projector augmented wave (PAW) [33] pseudo-potentials in a Vienna Ab initio Simulation Package (VASP) [34]. For the exchange-correlation functional, we used the Perdew-Burke-Ernzerhof (PBE) [35] form of the generalized gradient approximation (GGA). Grimme’s DFT-D2 method was used to describe the van der Waals interactions [36]. This method adds semi-empirical dispersion on the basis of conventional Kohn–Sham DFT energy. In calculations, Ti2N and Ti2NT2(T=O, F, OH) were constructed as 4 × 4 × 1 supercells, and a 20 Å vacuum layer was added in Z-axis direction to eliminate the interlayer interaction. The energy cutoff of 520 eV was used for wave functions’ expansion. The Brillouin zone integration was sampled with 5 × 5 k-grid mesh for geometry optimizations and 15 × 15 k-grid mesh for electronic properties calculations to achieve high accuracy. Geometry optimizations were performed by using the conjugated gradient method, and the convergence criterion was set to be 0.02 eV/Å on force and 1 × 10−5 eV/atom on energy. The adsorption energy (Eads) of nitrogen-containing gas (NCGs) on Ti2N or Ti2NT2(T=O, F, OH) is defined as follows:
Eads = EMXene + NCGs − E NCGs − EMXene
where Eads, EMXene + NCGs, EMXene, and ENCGs stand for the adsorption energy, total energies of optimized Ti2N or Ti2NT2(T=O, F, OH) MXene with adsorbed NCGs molecules, energies of pure MXene, and isolated NCGs (NO, NO2, and NH3) molecules respectively.

3. Results and Discussion

First of all, the Ti2N MXene has a hexagonal structure with three layers of atoms: the middle is N-atoms layer; and the upper (a) and lower (b) layers are Ti atoms, as is shown in Figure 1a. After structural optimizations, the Ti-N bond length is 2.07 Å, the bond angle of the same layer Ti and N (Ti–N–Ti) and N–Ti–N are both 92.4°, and the bond angle of different-layer Ti and N (Ti–N–Ti) is 87.6°. The lattice parameter of the 4 × 4 supercell is a = b = 11.98 Å. These parameters are very consistent with the results of Ti2N studied in other works in the literature [37,38], indicating the reliability of the calculation results. The band structure and PDOS of the material are shown in Figure 1b; it can be seen that there are many electronic states available for occupation at the Fermi level, and Ti2N exhibits metallic properties. The states near the Fermi level are occupied by Ti electrons, where N’s contribution is relatively small, meaning that low-energy carriers mainly originate from Ti.
In order to study the most stable adsorption configurations of NCGs on Ti2N MXene, the NO, NO2, and NH3 molecules were respectively placed at three different adsorption sites, namely on top of the Ti(a) atom shown in Figure 1a (site “a”), on top of the hollow Ti(b) atom (site “b”), or on top of the N atom (site “c”). The screened most stable configurations after adsorbing NCGs are shown in Figure 2. Both NO and NO2 show dissociative adsorption. The adsorption energy is far greater than those of other two-dimensional materials [39,40,41,42,43,44]. When NO is adsorbed on Ti2N, N and O atoms are adsorbed at the adjacent site “b” respectively. The distance between N and O is 3.07 Å, which is much larger than 1.15 Å of free NO gas molecule, indicating that the N–O bond is broken and NO dissociates. The dissociated N and O atoms form new bonds with Ti atoms in the upper layer of Ti2N, with bond lengths of 1.93 and 1.97 Å, respectively. Ti atoms protrude slightly upwards from its binding, and the Ti–N bond length on the surrounding matrix increases by about 0.14 Å. It can be seen from Figure 3a that, after dissociation, the N 2p and O 2p in NO and the nearby Ti 3d-orbitals have similar peak positions and obvious p-d hybridization in PDOS diagram. There is a strong bonding between NO and Ti2N, and the adsorption energy is as high as −9.809 eV. At the same time, the band structure shows that the dissociative adsorption of NO on Ti2N does not change the original metallic properties.
When NO2 interacts with Ti2N, it is interesting that NO2will also dissociate similar to NO adsorption, and all the N and O atoms are adsorbed at the site “b”. One O-atom keeps the same shape adsorbed at adjacent position, while the another O-atom is adsorbed away from it. The N-O bond length in NO2is increased to 3.07 Å. The dissociative adsorption energy is up to −14.336 eV. As can be seen from the electronic structure in Figure 3b, the p-d hybridization of N 2p-, O 2p-, and Ti 3d-orbitals corresponds with the strong bonding of Ti-N and Ti-O, respectively. The metallic behavior of Ti2N remains after NO2 dissociative adsorption. From the differential charge density in Figure 3d–f, it can be seen that the dissociation of NO and NO2 is caused by the excessive charge transfer from Ti2N MXene to the gas molecule. The Bader charge analysis shows that 2.57 e and 3.77 e charges were transferred from Ti2N to NO and NO2, respectively. The bulk of these transferred charges comes from the surrounding Ti atoms, and this is also the fundamental reason why the N-O bond is broken.
The adsorption of NH3 is completely different from those of NO and NO2. It is bound by associative adsorption with the binding distance of 2.2 Å and the Eads value of−1.286 eV. As shown in Figure 3c, NH3 tends to be adsorbed at site “a” with N-atom below and three H-atoms evenly distributed above. The distance between the Ti atom at site “a” and the surrounding N atom is elongated to 2.14 Å, and the N–Ti–N bond angle is reduced to 88.3°. To figure out the binding and charge transfer mechanism between NH3 molecule and Ti2N monolayer, the band structure and PDOS were analyzed. It is difficult to find a pronounced overlap between N p and Tid close to the Fermi level in Figure 3c. The amount of charge transfer is only 0.01 e, which is far less than those of NO and NO2. The bonding between NH3 and Ti2N is very weak, and the band structure that is almost the same as the bare Ti2N monolayer also confirms this.
For ideal gas sensing application, excessive adsorption energy is obviously not feasible. The dissociative adsorption of NO and NO2 on Ti2N monolayer will obviously poison the substrate and make it fail to work cyclically. The dissociation of gas molecules results from the strong interaction between the bare Ti2N surface and O atoms. Thus, the pure Ti2N monolayer can be used as a kind of gas-trapping material instead of highly efficient sensor for NO and NO2. Compared with those two gases, the adsorption effect of NH3 is satisfactory. The associative adsorption with lower energy will not cause chemical poisoning, which is beneficial to Ti2N as a promising candidate material for reversible NH3 sensors. In actual preparation process, some functional groups are usually introduced onto the Ti2N monolayer, and these functional groups may have significant impacts on adsorption properties of the material. Considering this reality, in the following part, we recount how we studied the adsorption behaviors of the Ti2N monolayer terminated by O, F, and OH towards NCGs.
We introduce O, F, and OH functional groups at three adsorption positions (site “a”, “b”, and “c”) of Ti2N, of which OH only considers the vertical direction (O atom is close to the matrix). All the functional groups tend to adsorb at site “b” with the lowest energy, as shown in Figure 4a–c. After adsorbing different functional groups, the Ti–N bonds show different degrees of extension: 2.164, 2.052, and 2.065 Å for Ti2NO2,Ti2NF2, and Ti2N(OH)2, respectively, which is due to the bonding between Ti and functional groups. Gouveia et al. reported possible atomic-layer-stacking-phase transitions of some MXenes [41]; thus, we also carefully checked this point in our research. It is found that the material system maintains phase stability and stacking way after introduction of those functional groups. The Bader charge analysis shows that 1.1, 0.8, and 0.8 e are respectively transferred to O, F, and OH from Ti2N. It can also be observed in Figure 4d–f that there is a strong overlap between the Tid-orbital and the T (O, F, or OH)d-orbitals.
After obtaining the stable configurations of Ti2NT2(T=O, F, OH), we placed various NCGs molecules at different adsorption sites (site “a”, “b”, and “c”), such as those in pure Ti2N. It can be seen from the most stable adsorption structures after optimizations in Figure 5 that all NCGs have no dissociative adsorption on Ti2NT2(T=O, F, OH). The binding distance is defined as the shortest direct distance between the NCGs molecule and the matrix; hereby, the binding distances of NO, NO2, and NH3 on Ti2NO2 are 2.082, 1.824, and 2.172 Å, respectively. The bond lengths of NO and NH3 after adsorption are the same as those of free gas molecules, while one of the N–O bonds of NO2 is stretched to 1.530 Å. The 2-D Ti2NO2 does not change significantly under NO environment, while the Ti atoms exposed to the adsorption positions of NO2 and NH3 are obviously dragged out. For NCGs adsorption on the surface of Ti2NF2, the molecular bond lengths and the matrix are almost unchanged with the adsorption distances of 2.951, 2.800, and 2.947 Å, respectively. It is surprising when NO is adsorbed on Ti2N(OH)2. The H atoms on the substrate bind with NO, where three H atoms combine with the N atom and one H atom combines with the O atom to form a hydroxyl lammonium-like product. The adsorption distances of NO2 and NH3are 1.625 and 1.702 Å, which are measured from the nearest H atom; thus, they are shorter than those of other Ti2NT2(T=O, F). While, the adsorption is not strong, especially the NH3 molecule has basically no change compared with the free molecule.
We can clearly see that the absolute values of adsorption energy of NCGs on Ti2NT2(T=O, F, OH) decrease sharply compared with the pure substrate in Table 1, especially on the surface of Ti2NF2. Since the H atoms of the OH functional groups fall off the Ti2N(OH)2 substrate and bind with NO, their decrease is not that large. On the pure Ti2N monolayer, a large amount of charge is transferred from Ti2N to NO and NO2. After the introduction of the functional group O, F, or OH, most of the charge transfer from Ti2N is captured by these functional groups, which leaves a very small charge amount for exchange with NCGs. In addition, one can see that the difference in adsorption energy of each NCG on substrate is obvious, indicating that the selectivity is good. As mentioned above, NO and NO2 are dissociatively adsorbed on pure Ti2N, which results in the poisoning of the Ti2N surface. Similarly, NO binds with H atoms of Ti2N(OH)2 and destroys the surface integrity of the substrate, which is detrimental to the sensing material. Thus, Ti2N(OH)2 is not suitable for NO sensing. From the perspective of adsorption energy, those of NO on Ti2NO2, NO2on Ti2NF2, and NH3on Ti2N(OH)2 are between physisorption and chemisorption, which are important to make sensors work reversibly and increase their efficiency [45].
To better understand the effects of NCGs molecules on Ti2NT2(T=O, F, OH), the band structure and PDOS of all the adsorption systems are calculated and shown in Figure 6. Regardless of the type of NCGs adsorption, the system retains metallic properties. Only when Ti2NO2 is exposed to NO2, the Tid-, Op-,and N p-orbitals show a certain degree of d-p hybridization, while the p- and d-orbitals in other PDOS do not overlap. When Ti2NO2 adsorbs NO, a flat band appears near the Fermi energy and indicates that the electrons are delocalized, which further proves that the interaction between NO and Ti2NO2 is not strong. This phenomenon is more obvious when NCGs are adsorbed on Ti2NF2 shown in Figure 6d–f. These results are consistent with the above discussion on energetics and charge-transfer profiles.
An ideal sensing material not only needs to possess excellent gas-binding capability but also has a requirement for good sensitivity. The conductivity of materials, especially two-dimensional materials, is closely related with work function. During adsorption, if the electron affinity of gas molecule surpasses the work function of substrate materials, the molecule will grab electrons from the surface of materials and the molecule would be charged negatively; in contrast, the adsorbed molecule would be charged positively. Both two cases would change the electronic structure, as well as conductivity of materials. The work function [46] is defined as the least energy required to free an electron from the surface of a system, which can be calculated by the following equation:
φ= V − Ef
where φ, V, and Ef are the work function, electrostatic potential at the vacuum level, and the Fermi energy, respectively. Figure 7 shows the change of work function of Ti2N and Ti2NT2(T=O, F, OH) before and after adsorption of various NCGs. “Bare” represents the bare system (before adsorption), and the names of NCGs in abscissa represent various systems after adsorption of different gases. The work functions of Ti2NO2 and Ti2N(OH)2 decrease obviously after adsorbing NO, while that of Ti2NF2 increases significantly after adsorbing NO2. After adsorbing NH3, the work functions of Ti2N, Ti2NF2, and Ti2NO2 all decrease, and the change of Ti2NO2 is the most obvious. Because all the systems are metallic, even small changes in work function can cause large variation in conductivity [43]. Comprehensively considering the above results on adsorption structure and energy, Ti2NO2 can be used as a superior sensing candidate material for NO and NH3, and Ti2NF2 can be utilized as that for NO2. In addition, the potential of sensing properties of Ti2N, Ti2NF2, and Ti2N(OH)2 towards NH3 is also good.
Finally, we discuss the recovery time (τ). An ideal gas-sensing material needs appropriate adsorption energy, and a shorter recovery time is better. To make a gas sensor work efficiently, a short recovery time is required. It can be calculated through transition state theory, namely by the following relationship [47,48]:
τ = υ−1exp(−Eads/kT)
where υ, Eads, k, and T are the attempt frequency, adsorption energy, Boltzmann’s constant (8.62 × 10−5 eV K−1), and operational temperature, respectively. Under the same conditions, the increase in adsorption energy will lead to an exponential increase in recovery time; then the higher operating temperature compensation is required, which is detrimental to the sensor performance. In order to get a relatively reasonable value, we set the attempt frequency here to 1012s−1 based on traditional transition state theory [49,50,51]. The recovery time of NO on Ti2NO2 is calculated to be 2.16 × 105 us, and that of NO2/Ti2NF2 and NH3/Ti2NF2 is 1.29 and 1.71 ×10−4 us, respectively. Here it can be seen in Figure 8 that the adsorption energy difference of less than 1 eV can be enlarged to several orders of magnitude in recovery time. Although NO on Ti2NO2 has the longest recovery time, the value is only 0.216 s. For the system of NH3/Ti2NF2, it has the shortest recovery time, which shows that the MXene system in this research can be promising. Therefore, when other conditions are met, the candidate materials with appropriate gas adsorption energy and short response time should be selected.

4. Summary and Outlook

In this work, we used the first-principles DFT calculations to investigate the 2-D Ti2N nanosheet and its Ti2NT2(T=O, F, OH) derivative materials for harmful nitrogen-containing gas (NGCs) adsorption and sensing applications. When NCGs are exposed on the surface of Ti2N, the dissociative adsorption of NO and NO2 occurs and destroys the stability of the nanosheet, which would make it hard to work reversibly, while NH3 does not have this effect on Ti2N. When studying the interaction mechanisms between NCGs and Ti2NT2(T=O, F, OH), it can be found that a large amount of charges accumulates at functional groups, thus greatly reducing the adsorption energy of NCGs in absolute value. The obvious change in work function, coupled with the metallic nature of the systems, would improve gas sensitivity. Appropriate gas adsorption energy can directly determine fast recovery time to ensure the efficiency and reversibility of sensors. This research gives the electronic structures origin and application prospect of Ti2NT2(T=O, F, OH) as efficient NCGs sensing materials and would inspire experimentalists to explore better 2-D candidates in the field.

Author Contributions

Formal analysis, H.Z., W.D., J.Z., R.A., and Z.Q.; project administration, Z.Q.; supervision, Z.Q.; writing—original draft, H.Z., W.D., and Z.Q.; writing—review and editing, H.Z. and Z.Q.; funding acquisition, Z.Q. All authors have read and agreed to the published version of the manuscript.

Funding

We are grateful for the support from the Natural Science Foundation of China (51801113 and 52171182), the Sino-German Center for Research Promotion and NSFC (GZ1673), the Natural Science Foundation of Shandong Province (ZR2018MEM001), the Young Scholars Program of Shandong University (YSPSDU), the Education Program of Shandong University (2020Y299), and the Fundamental Research Funds for the Central Universities.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The National Supercomputer Centre (NSC) and the HPC Cloud Platform of Shandong University are acknowledged.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ramanathan, V.; Feng, Y. Air pollution, greenhouse gases and climate change: Global and regional perspectives. Atmos. Environ. 2009, 43, 37–50. [Google Scholar] [CrossRef]
  2. Koolen, C.D.; Rothenberg, G. Air pollution in Europe. ChemSusChem 2018, 12, 164–172. [Google Scholar] [CrossRef]
  3. Hasselblad, V.; Eddy, D.M.; Kotchmar, D.J. Synthesis of environmental evidence: Nitrogen dioxide epidemiology studies. J. Air Waste Manag. Assoc. 1992, 42, 662–671. [Google Scholar] [CrossRef]
  4. Garrett, M.H.; Hooper, M.A.; Hooper, B.M.; Abramson, M.J. Respiratory symptoms in children and indoor exposure to nitrogen dioxide and gas stoves. Am. J. Respir. Crit. Care Med. 1998, 158, 891–895. [Google Scholar] [CrossRef] [PubMed]
  5. Ouyang, T.; Qian, Z.; Ahuja, R.; Liu, X. First-principles investigation of CO adsorption on pristine, C-doped and N-vacancy defected hexagonal AlN nanosheets. Appl. Surf. Sci. 2018, 439, 196–201. [Google Scholar] [CrossRef]
  6. Liu, Y.; Parisi, J.; Sun, X.; Lei, Y. Solid-state gas sensors for high temperature applications—A review. J. Mater. Chem. A 2014, 2, 9919–9943. [Google Scholar] [CrossRef]
  7. Tang, X.; Du, A.; Kou, L. Gas sensing and capturing based on two-dimensional layered materials: Overview from theoretical perspective. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2018, 8, e1361. [Google Scholar] [CrossRef]
  8. Zhang, Y.-H.; Chen, Y.-B.; Zhou, K.-G.; Liu, C.; Zeng, J.; Zhang, H.-L.; Peng, Y. Improving gas sensing properties of graphene by introducing dopants and defects: A first-principles study. Nanotechnology 2009, 20, 185504. [Google Scholar] [CrossRef] [Green Version]
  9. Liu, L.; Yang, Q.; Wang, Z.; Ye, H.; Chen, X.; Fan, X.; Zhang, G. High Selective Gas Detection for small molecules based on Germanium selenide monolayer. Appl. Surf. Sci. 2018, 433, 575–581. [Google Scholar] [CrossRef]
  10. Li, F.; Shi, C. NO-sensing performance of vacancy defective monolayer MoS2 predicted by density function theory. Appl. Surf. Sci. 2018, 434, 294–306. [Google Scholar] [CrossRef]
  11. Yu, X.-F.; Li, Y.-C.; Cheng, J.-B.; Liu, Z.-B.; Li, Q.-Z.; Li, W.-Z.; Yang, X.; Xiao, B. Monolayer Ti2CO2: A Promising Candidate for NH3 Sensor or Capturer with High Sensitivity and Selectivity. ACS Appl. Mater. Interfaces 2015, 7, 13707–13713. [Google Scholar] [CrossRef]
  12. Xiao, B.; Li, Y.-C.; Yu, X.-F.; Cheng, J.-B. MXenes: Reusable materials for NH3 sensor or capturer by controlling the charge injection. Sens. Actuators B Chem. 2016, 235, 103–109. [Google Scholar] [CrossRef]
  13. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Naguib, M.; Mashtalir, O.; Carle, J.; Presser, V.; Lu, J.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional transition metal carbides. ACS Nano 2012, 6, 1322–1331. [Google Scholar] [CrossRef]
  15. Feng, A.; Yu, Y.; Wang, Y.; Jiang, F.; Yu, Y.; Mi, L.; Song, L. Two-dimensional MXene Ti3C2 produced by exfoliation of Ti3AlC2. Mater. Des. 2017, 114, 161–166. [Google Scholar] [CrossRef]
  16. Lashgari, H.; Abolhassani, M.; Boochani, A.; Elahi, S.; Khodadadi, J. Electronic and optical properties of 2D graphene-like compounds titanium carbides and nitrides: DFT calculations. Solid State Commun. 2014, 195, 61–69. [Google Scholar] [CrossRef]
  17. Weng, H.; Ranjbar, A.; Liang, Y.; Song, Z.; Khazaei, M.; Yunoki, S.; Arai, M.; Kawazoe, Y.; Fang, Z.; Dai, X. Large-gap two-dimensional topological insulator in oxygen functionalized MXene. Phys. Rev. B 2015, 92, 075436–0754367. [Google Scholar] [CrossRef] [Green Version]
  18. Pan, H. Electronic properties and lithium storage capacities of two-dimensional transition-metal nitride monolayers. J. Mater. Chem. A 2015, 3, 21486–21493. [Google Scholar] [CrossRef]
  19. Wang, D.; Liu, Y.; Meng, X.; Wei, Y.; Zhao, Y.; Pang, Q.; Chen, G. Two-dimensional VS2 monolayers as potential anode materials for lithium-ion batteries and beyond: First-principles calculations. J. Mater. Chem. A 2017, 5, 21370–21377. [Google Scholar] [CrossRef]
  20. Kumar, H.; Frey, N.C.; Dong, L.; Anasori, B.; Gogotsi, Y.; Shenoy, V.B. Tunable magnetism and transport properties in nitride MXenes. ACS Nano 2017, 11, 7648–7655. [Google Scholar] [CrossRef]
  21. Yang, Z.; Zheng, Y.; Li, W.; Zhang, J. Investigation of two-dimensional hf-based MXenes as the anode materials for li/na-ion batteries: A DFT study. J. Comput. Chem. 2019, 40, 1352–1359. [Google Scholar] [CrossRef] [PubMed]
  22. Gao, G.; Ding, G.; Li, J.; Yao, K.; Wu, M.; Guangqian, D. Monolayer MXenes: Promising half-metals and spin gapless semiconductors. Nanoscale 2016, 8, 8986–8994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Pang, J.; Mendes, R.G.; Bachmatiuk, A.; Zhao, L.; Ta, H.Q.; Gemming, T.; Liu, H.; Liu, Z.; Rummeli, M.H. Applications of 2D MXenes in energy conversion and storage systems. Chem. Soc. Rev. 2018, 48, 72–133. [Google Scholar] [CrossRef] [PubMed]
  24. Shukla, V.; Jena, N.K.; Naqvi, S.R.; Luo, W.; Ahuja, R. Modelling high-performing batteries with Mxenes: The case of S-functionalized two-dimensional nitride Mxene electrode. Nano Energy 2019, 58, 877–885. [Google Scholar] [CrossRef]
  25. Gouveia, J.; Novell-Leruth, G.; Vines, F.; Illas, F.; Gomes, J. The Ti2CO2MXene as a nucleobase 2D sensor: A first-principles study. Appl. Surf. Sci. 2021, 544, 148946. [Google Scholar] [CrossRef]
  26. Gouveia, J.; Novell-Leruth, G.; Reis, P.M.L.S.; Viñes, F.; Illas, F.; Gomes, J. First-principles calculations on the adsorption behavior of amino acids on a titanium carbide MXene. ACS Appl. Biol. Mater. 2020, 3, 5913–5921. [Google Scholar] [CrossRef]
  27. Dolz, D.; Morales-García, A.; Viñes, F.; Illas, F. Exfoliation energy as a descriptor of MXenes synthesizability and surface chemical activity. Nanomaterials 2021, 11, 127. [Google Scholar] [CrossRef] [PubMed]
  28. Soundiraraju, B.; George, B.K. Two-dimensional Titanium Nitride (Ti2N) MXene: Synthesis, characterization, and potential application as surface-enhanced ramanscattering substrate. ACS Nano 2017, 11, 8892–8900. [Google Scholar] [CrossRef]
  29. Li, Y.; Guo, Y.; Chen, W.; Jiao, Z.; Ma, S. Reversible hydrogen storage behaviors of Ti2NMXenes predicted by first-principles calculations. J. Mater. Sci. 2018, 54, 493–505. [Google Scholar] [CrossRef]
  30. Lin, H.; Yang, D.-D.; Lou, N.; Zhu, S.-G.; Li, H.-Z. Functionalized titanium nitride-based MXenes as promising host materials for lithium-sulfur batteries: A first principles study. Ceram. Int. 2018, 45, 1588–1594. [Google Scholar] [CrossRef]
  31. Urbankowski, P.; Anasori, B.; Makaryan, T.; Er, D.; Kota, S.; Walsh, P.L.; Zhao, M.; Shenoy, V.B.; Barsoum, M.W.; Gogotsi, Y. Synthesis of two-dimensional titanium nitride Ti4N3(MXene). Nanoscale 2016, 8, 11385–11391. [Google Scholar] [CrossRef]
  32. Xiao, X.; Yu, H.; Jin, H.; Wu, M.; Fang, Y.; Sun, J.; Hu, Z.; Li, T.; Wu, J.; Huang, L.; et al. Salt-templated synthesis of 2D metallic MoN and other nitrides. ACS Nano 2017, 11, 2180–2186. [Google Scholar] [CrossRef]
  33. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  34. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  35. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  36. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  37. Wang, D.; Gao, Y.; Liu, Y.; Jin, D.; Gogotsi, Y.; Meng, X.; Du, F.; Chen, G.; Wei, Y. First-principles calculations of Ti2N and Ti2NT2 (T=O, F, OH) monolayers as potential anode materials for lithium-ion batteries and beyond. J. Phys. Chem. C 2017, 121, 13025–13034. [Google Scholar] [CrossRef]
  38. Xie, Y.; Kent, P.R.C. Hybrid density functional study of structural and electronic properties of functionalized Tin+1Xn (X=C, N) monolayers. Phys. Rev. B 2013, 87, 235441. [Google Scholar] [CrossRef] [Green Version]
  39. Prasongkit, J.; Amorim, R.; Chakraborty, S.; Ahuja, R.; Scheicher, R.; Amornkitbamrung, V. Highly sensitive and selective gas detection based on silicene. J. Phys. Chem. C 2015, 119, 16934–16940. [Google Scholar] [CrossRef]
  40. Leenaerts, O.; Partoens, B.; Peeters, F.M. Adsorption of H2O, NH3, CO, NO2, and NO on graphene: A first-principles study. Phys. Rev. B 2008, 77, 125416. [Google Scholar] [CrossRef] [Green Version]
  41. Gouveia, J.D.; Viñes, F.; Illas, F.; Gomes, J.R.B. MXenes atomic layer stacking phase transitions and their chemical activity consequences. Phys. Rev. Mater. 2020, 4, 054003. [Google Scholar] [CrossRef]
  42. Ouyang, T.; Qian, Z.; Hao, X.; Ahuja, R.; Liu, X. Effect of defects on adsorption characteristics of AlN monolayer towards SO2 and NO2: Ab initio exposure. Appl. Surf. Sci. 2018, 462, 615–622. [Google Scholar] [CrossRef]
  43. Kou, L.; Frauenheim, T.; Chen, C. Phosphorene as a superior gas sensor: Selective adsorption and distinct I–V response. J. Phys. Chem. Lett. 2014, 5, 2675–2681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Yue, Q.; Shao, Z.; Chang, S.; Li, J. Adsorption of gas molecules on monolayer MoS2 and effect of applied electric field. Nanoscale Res. Lett. 2013, 8, 425. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Hussain, T.; Singh, D.; Gupta, S.K.; Karton, A.; Sonvane, Y.; Ahuja, R. Efficient and selective sensing of nitrogen-containing gases by Si2BN nanosheets under pristine and pre-oxidized conditions. Appl. Surf. Sci. 2018, 469, 775–780. [Google Scholar] [CrossRef]
  46. Gholizadeh, R.; Yu, Y.-X. Work functions of pristine and heteroatom-doped graphenes under different external electric fields: An ab initio DFT study. J. Phys. Chem. C 2014, 118, 28274–28282. [Google Scholar] [CrossRef]
  47. Barsan, N.; Koziej, D.; Weimar, U. Metal oxide-based gas sensor research: How to? Sens. Actuators B Chem. 2007, 121, 18–35. [Google Scholar] [CrossRef]
  48. Nagarajan, V.; Chandiramouli, R. Investigation of NH3 adsorption behavior on graphdiyne nanosheet and nanotubes: A first-principles study. J. Mol. Liq. 2018, 249, 24–32. [Google Scholar] [CrossRef]
  49. Eyring, H. The activated complex in chemical reactions. J. Chem. Phys. 1935, 3, 107–115. [Google Scholar] [CrossRef]
  50. Hadipour, N.L.; Peyghan, A.A.; Soleymanabadi, H. Theoretical study on the Al-doped ZnO nanoclusters for CO chemical sensors. J. Phys. Chem. C 2015, 119, 6398–6404. [Google Scholar] [CrossRef]
  51. Ahmadi, A.; Hadipour, N.L.; Kamfiroozi, M.; Bagheri, Z. Theoretical study of aluminum nitride nanotubes for chemical sensing of formaldehyde. Sens. Actuators B Chem. 2012, 161, 1025–1029. [Google Scholar] [CrossRef]
Figure 1. (a) Top and side view of the optimized structures for Ti2N monolayer (blue and silver balls represent Ti and N atoms, respectively). (b) Band structure and projected density of states of Ti2N monolayer (Tid-orbital uses top abscissa and other orbitals use bottom coordinate).
Figure 1. (a) Top and side view of the optimized structures for Ti2N monolayer (blue and silver balls represent Ti and N atoms, respectively). (b) Band structure and projected density of states of Ti2N monolayer (Tid-orbital uses top abscissa and other orbitals use bottom coordinate).
Nanomaterials 11 02459 g001
Figure 2. Top and side view of the optimized atomic structures for Ti2N monolayer after adsorbing (a) NO, (b) NO2, and (c) NH3 (blue, silver, red, and pink balls represent Ti, N, O, and H atoms, respectively).
Figure 2. Top and side view of the optimized atomic structures for Ti2N monolayer after adsorbing (a) NO, (b) NO2, and (c) NH3 (blue, silver, red, and pink balls represent Ti, N, O, and H atoms, respectively).
Nanomaterials 11 02459 g002
Figure 3. Band structures, PDOS and differential charge densities of Ti2N after adsorbing (a,d) NO, (b,e) NO2, and (c,f) NH3 (Tid-orbital uses top abscissa and other orbitals use bottom coordinate; isovalue of NO and NO2 is set to 0.01 e/Å3, NH3 is set to 0.005 e/Å3).
Figure 3. Band structures, PDOS and differential charge densities of Ti2N after adsorbing (a,d) NO, (b,e) NO2, and (c,f) NH3 (Tid-orbital uses top abscissa and other orbitals use bottom coordinate; isovalue of NO and NO2 is set to 0.01 e/Å3, NH3 is set to 0.005 e/Å3).
Nanomaterials 11 02459 g003
Figure 4. Top and side view of the optimized structures, band structures, and PDOS of (a,d) Ti2NO2, (b,e) Ti2NF2, and (c,f) Ti2N(OH)2 monolayers (blue, silver, red, pink, and yellow balls represent Ti, N, O, H, and F atoms, respectively. Tid-orbital uses top abscissa and other orbitals use bottom coordinate).
Figure 4. Top and side view of the optimized structures, band structures, and PDOS of (a,d) Ti2NO2, (b,e) Ti2NF2, and (c,f) Ti2N(OH)2 monolayers (blue, silver, red, pink, and yellow balls represent Ti, N, O, H, and F atoms, respectively. Tid-orbital uses top abscissa and other orbitals use bottom coordinate).
Nanomaterials 11 02459 g004
Figure 5. Top and side view of the optimized structures for Ti2NO2 after adsorbing (a) NO, (b) NO2, and (c) NH3; Ti2NF2 after adsorbing (d) NO, (e) NO2, and (f) NH3;and Ti2N(OH)2 after adsorbing (g) NO, (h) NO2, and (i) NH3 (blue, silver, red, pink, and yellow balls represent Ti, N, O, H, and F atoms, respectively).
Figure 5. Top and side view of the optimized structures for Ti2NO2 after adsorbing (a) NO, (b) NO2, and (c) NH3; Ti2NF2 after adsorbing (d) NO, (e) NO2, and (f) NH3;and Ti2N(OH)2 after adsorbing (g) NO, (h) NO2, and (i) NH3 (blue, silver, red, pink, and yellow balls represent Ti, N, O, H, and F atoms, respectively).
Nanomaterials 11 02459 g005
Figure 6. Band structures and PDOS of Ti2NO2 after adsorbing (a) NO, (b) NO2, and (c) NH3; Ti2NF2 after adsorbing (d) NO, (e) NO2, and (f) NH3; Ti2N(OH)2 after adsorbing (g) NO, (h) NO2, and (i) NH3 (Tid-orbital uses top abscissa and other orbitals use bottom coordinate).
Figure 6. Band structures and PDOS of Ti2NO2 after adsorbing (a) NO, (b) NO2, and (c) NH3; Ti2NF2 after adsorbing (d) NO, (e) NO2, and (f) NH3; Ti2N(OH)2 after adsorbing (g) NO, (h) NO2, and (i) NH3 (Tid-orbital uses top abscissa and other orbitals use bottom coordinate).
Nanomaterials 11 02459 g006
Figure 7. Work function change of Ti2N and Ti2NT2(T=O, F, OH) before and after adsorption of various NCGs.
Figure 7. Work function change of Ti2N and Ti2NT2(T=O, F, OH) before and after adsorption of various NCGs.
Nanomaterials 11 02459 g007
Figure 8. Recovery time of different sensing systems.
Figure 8. Recovery time of different sensing systems.
Nanomaterials 11 02459 g008
Table 1. Adsorption energy (Eads) and charge-transfer profiles of NCGs adsorption on 2-D Ti2N and Ti2NT2(T=O, F, OH).
Table 1. Adsorption energy (Eads) and charge-transfer profiles of NCGs adsorption on 2-D Ti2N and Ti2NT2(T=O, F, OH).
NCGsAdsorption Energy (eV)Charge Transfer (e)
Ti2NTi2NO2Ti2NF2Ti2N(OH)2Ti2NTi2NO2Ti2NF2Ti2N(OH)2
NO−9.81−0.68−0.15−5.552.57−0.24−0.070.50
NO2−14.34−2.38−0.36−3.783.770.490.291.78
NH3−1.29−1.16−0.13−0.570.01−0.18−0.030.15
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, H.; Du, W.; Zhang, J.; Ahuja, R.; Qian, Z. Nitrogen-Containing Gas Sensing Properties of 2-D Ti2N and Its Derivative Nanosheets: Electronic Structures Insight. Nanomaterials 2021, 11, 2459. https://doi.org/10.3390/nano11092459

AMA Style

Zhang H, Du W, Zhang J, Ahuja R, Qian Z. Nitrogen-Containing Gas Sensing Properties of 2-D Ti2N and Its Derivative Nanosheets: Electronic Structures Insight. Nanomaterials. 2021; 11(9):2459. https://doi.org/10.3390/nano11092459

Chicago/Turabian Style

Zhang, Hongni, Wenzheng Du, Jianjun Zhang, Rajeev Ahuja, and Zhao Qian. 2021. "Nitrogen-Containing Gas Sensing Properties of 2-D Ti2N and Its Derivative Nanosheets: Electronic Structures Insight" Nanomaterials 11, no. 9: 2459. https://doi.org/10.3390/nano11092459

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