Next Article in Journal
Synthesis and Electronic Properties of Novel Donor–π–Acceptor-Type Functional Dyes with a Carbonyl-Bridged Bithiophene π-Spacer
Previous Article in Journal
Formation and Characterization of Xylitol-Modified Glycidyl Methacrylate-co-Ethyl Methacrylate Matrices for Controlled Release of Antimicrobial Compounds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Phosphorene-Supported Au(I) Fragments for Highly Sensitive Detection of NO

School of Chemistry, State Key Laboratory of Fine Chemicals, Frontier Science Center for Smart Materials, Dalian Key Laboratory of Intelligent Chemistry, Dalian University of Technology, Dalian 116024, China
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(15), 3085; https://doi.org/10.3390/molecules30153085
Submission received: 12 June 2025 / Revised: 17 July 2025 / Accepted: 19 July 2025 / Published: 23 July 2025

Abstract

The fabrication and application of single-site heterogeneous reaction centers are new frontiers in chemistry. Single-site heterogeneous reaction centers are analogous to metal centers in enzymes and transition-metal complexes: they are charged and decorated with ligands and would exhibit superior reactivity and selectivity in chemical conversion. Such high reactivity would also result in significant response, such as a band gap or resistance change, to approaching molecules, which can be used for sensing applications. As a proof of concept, the electronic structure and reaction pathways with NO and NO2 of Au(I) fragments dispersed on phosphorene (Pene) were investigated with first-principle-based calculations. Atomic-deposited Au atoms on Pene (Au1-Pene) have hybridized Au states in the bulk band gap of Pene and a decreased band gap of 0.14 eV and would aggregate into clusters. Passivation of the Au hybrid states with -OH and -CH3 forms thermodynamically plausible HO-Au1-Pene and H3C-Au1-Pene and restores the band gap to that of bulk Pene. Inspired by this, HO-Au1-Pene and H3C-Au1-Pene were examined for detection of NO and NO2 that would react with -OH and -CH3, and the resulting decrease of band gap back to that of Au1-Pene would be measurable. HO-Au1-Pene and H3C-Au1-Pene are highly sensitive to NO and NO2, and their calculated theoretical sensitivities are all 99.99%. The reaction of NO2 with HO-Au1-Pene is endothermic, making the dissociation of product HNO3 more plausible, while the barriers for the reaction of CH3-Au1-Pene with NO and NO2 are too high for spontaneous detection. Therefore, HO-Au1-Pene is not eligible for NO2 sensing and CH3-Au1-Pene is not eligible for NO and NO2 sensing. The calculated energy barrier for the reaction of HO-Au-Pene with NO is 0.36 eV, and the reaction is about thermal neutral, suggesting HO-Au-Pene is highly sensitive for NO sensing and the reaction for NO detection is spontaneous. This work highlights the potential superior sensing performance of transition-metal fragments and their potential for next-generation sensing applications.

1. Introduction

Nitrogen oxides (NOx), represented by NO and NO2, are harmful to both the living environment and human health. Considerable research attention has been devoted continuously to develop novel procedures and devices for efficient and sensitive detection of NOx [1,2,3,4,5]. Two-dimensional (2D)-material-based NOx sensors are emerging for the large surface area, easiness of functionalization, superior electronic and optical properties, etc., of the materials [6,7,8,9,10,11] and are complementary to conventional metal-oxide-based and metal-nitride-based sensors, etc., in NOx detection [12,13].
Phosphorene (Pene), also known as single- or few-layer black phosphorus, is one of the most outstanding 2D materials since the rise of graphene [14,15,16,17]. In Pene, P atoms are covalently interconnected into 2D layers, and these layers are stacked together through van der Waals interactions. Compared with other 2D materials, Pene possesses a tunable direct band gap (from 0.3 to 2.0 eV, depending on its thickness) and a high carrier mobility (approximately 1000 cm2/V·s at r. t.), making it an excellent candidate for gas-sensing applications [18]. The gas-sensing application of Pene has drawn considerable research attention in recent years. Cho et al. investigated the sensing performances of Pene, MoS2, and graphene and found that the performance of Pene is superior in dynamic sensing response, sensitivity, selectivity, and response time. They showed that the sensitivity of Pene to target gases are ~20 times higher than those of graphene and MoS2. The response time of Pene was ~40 times shorter than that of other 2D materials capable of ppb-level gas detection [19]. Cui et al. fabricated a phosphorene-based sensor and demonstrated that this sensor could detect NO2 at ppb levels in dry air at room temperature [20]. Zhang et al. investigated the adsorption of NO, NO2, CO, CO2, and NH3 on Pene under different strains with first-principle-based calculations and showed that the adsorption energies depend strongly on the charge transfer with Pene [21]. Kumawat et al. explored the potential of armchair Pene nanoribbons for explosive detection through first-principle-based calculations and found that these nanoribbons exhibit excellent sensitivity and selectivity toward certain explosive molecules [22]. Apart from being used directly, Pene can be modified further for sensing applications. Surface modification of Pene is primarily achieved through the generating of surface defects or deposition of transition-metal species [23,24,25]. Doping with transition metals [26,27,28,29] or compounds [30,31,32] would not only benefit the air and humidity stabilities and oxidation resistance of Pene [33,34] but also significantly enhance the gas-sensing performance of Pene [35,36,37,38,39,40,41,42]. Cheng et al. proposed that Ag- and Au-doped Pene would be highly sensitive for NO sensing according to the change of band gap [29]. Ghadiri et al. investigated the H2S-sensing performance of pristine Pene and Mn-modified Pene by first-principle-based calculations and showed that Mn-modified Pene is highly selective to H2S and is reusable [43]. Ghambarian et al. investigated the H2-sensing performances of Ni-, Pt-, and Pd-doped Pene and showed that Ni-doped Pene can be used for H2 purification, while Pt-doped Pene can be used as a H2 sensor [44]. Wang et al. reported the synthesis and NO2-sensing application of Ag-nanoparticle-modified Pene [45]. To this end, surface modification with various metal fragments would result in transition-metal-Pene-based sensors with improved gas-sensing performance compared with pristine Pene [46]. Furthermore, the vast combinations of transition-metal fragments on Pene also enable the simultaneous detection and identification of different target molecules [47]. However, it should be noted that most of the aforementioned detection is based on the adsorption of target molecules onto the reactive sites of the senor. The undercoordination of transition-metal atoms on Pene also enables them to adsorb and activate approaching molecules efficiently at the expense of selectivity.
The fabrication and application of single-site heterogeneous reaction centers are new frontiers in chemistry. Single-site heterogeneous reaction centers are analogous to metal centers in enzymes and transition-metal complexes; they would exhibit superior reactivity and selectivity in chemical conversion. Their performance would be tunable by controlling the charge and ligands on them, and they are different from conventional singlet atom sensors. The high reactivity would also result in significant response, such as band gap or resistance change, to approaching molecules, which can be used for sensing applications. Inspired by the superior selectivity of single-site heterogenous reaction centers and the reported sensing performance of Pene-based sensors, we investigated the electronic structure and reaction pathways with NO and NO2 of Au(I) fragments on Pene with extensive first-principle-based calculations [46,48,49]. We expect these Au fragments to exhibit superior sensing performance to NO and NO2 in terms of high reactivity, selectivity, and fast response. We also expect the findings to pave the way for design and fabrication of Pene-based high-performance sensors for NOx detection.

2. Results and Discussions

The adsorption structures of Au atoms and small clusters containing two and three Au atoms on Pene were investigated firstly (Figure 1). The Au adsorption on top of a surface P atom (T), bridging two nearest neighboring P atoms (B), and on top of the center of three surface P atoms (H) were investigated, and all potential spin states were considered. The adsorption of a Au atom at B site is found plausible, and the calculated adsorption energy (Eb) is −1.48 eV, corresponding to a doublet spin symmetry (Au1-Pene, Figure 1a). This is in reasonable agreement with the reported value of −1.61 eV for Au atomic adsorption on Pene obtained with VASP [50]. The slight difference can be attributed to the different implementation of theory within the code used. The adsorption of Au2 and Au3 clusters was also investigated in the same way (Figure 1b,c). The calculated Eb values averaged over the number of Au atoms are −1.61 and −2.06 eV for Au2-Pene and Au3-Pene, respectively, and the calculated free–energy values of these Au species follow the same trend in the T range from 250 to 600 K (Figure 1d). In this sense, Au1-Pene on Pene is not thermodynamically stable, and Au atoms would aggregate to form clusters [51]. This is different from the previously reported case of Pd atomic deposition on Pene, where Pd1-Pene is preferred [52].
The calculated band gap of pristine Pene is 0.87 eV [48,52,53]. The highest occupied state of Au1-Pene is the half-occupied localized state on the Au atom and is within the band gap of Pene at the Fermi level. The calculated band gap of Au1-Pene is only 0.14 eV and is much lower compared with pristine Pene. According to molecular orbital theory, the chemical bonding of Au with a species with a half-occupied state would pair the electrons and downshift this Au state, leading to a drastically enlarged band gap that is similar to that of bulk Pene. The passivation of the state on Au with -OH(HO-Au1-Pene) and -CH3(H3C-Au1-Pene) were investigated as model systems (Figure 2c,d). With the passivation of the unpaired electron state on Au1-Pene, the band gaps of HO-Au1-Pene (Figure 2c) and H3C-Au1-Pene (Figure 2d) increase to 0.88 and 0.80 eV, respectively, as expected. The ground states of these Au(I) fragments on Pene have singlet symmetry. Such change in electronic structure can be considered the response of Au1-Pene to the stimuli induced by the adsorption or bonding of the approaching -OH and -CH3. The band gap change is ~0.74 eV, which would be significant enough to be detected. However, as Au1-Pene is a highly reactive radical with the unpaired electron localized on the Au atom and we have shown that Au1-Pene would aggregate to form plausible Au clusters on Pene, it cannot be used directly to detect NOx. HO-Au1-Pene (Figure 2c) and H3C-Au1-Pene (Figure 2d) were examined further, considering their potential reactions with species with unpaired electrons, such as NOx, etc., may generate more plausible compounds, leaving the band gap decreasing to ~0.14 eV in formed Au1-Pene.
The possibilities for the clustering of HO-Au1-Pene and H3C-Au1-Pene were examined firstly (Figure 3 and Figure 4). After passivation with -OH, the Au atom is at its +1 valence state and coordinates linearly with -OH and one surface P atom (Figure 3a,a’). If the Au(I) fragments are far apart, their contribution to the electronic structure would be the same as a single Au(I) fragment. The possible dimers of HO-Au1-Pene were investigated, and hydrogen bonds between adjacent Au(I) fragments are apparent (Figure 3b–d). The calculated formation free energy (ΔG) in Figure 3f shows that these hydrogen bonds are not strong enough to contribute to additional stability of the dimers, and the monodispersed HO-Au1-Pene is the most plausible among all the structures considered (Figure 3f). The same was found for H3C-Au1-Pene (Figure 4), where the Au atom is also +1 |e| charged and coordinates linearly with both -CH3 and a surface P atom in the most plausible monodispersed form (Figure 4a,a’). As there are only van der Waals interactions between H3C-Au1-Pene and the dimers (Figure 4b–e), the difference in formation free energy between the monomer and the dimer is ~0.3 eV (Figure 4f) and is even larger than that for dimers of HO-Au1-Pene. Based on these, it can be safely concluded that monodispersed Au(I) fragments are plausible when they are far apart and there is no risk for them to aggregate into clusters.
The mechanistic pathways for the reactions between NO and NO2 and HO-Au1-Pene and H3C-Au1-Pene were investigated to highlight the potential sensing performance of these Pene-supported Au(I) fragments (Figure 5). NO would adsorb on and react with HO-Au1-Pene spontaneously, forming HNO2 adsorbed on Au1-Pene by crossing TSNO+OH, and the calculated energy barrier and reaction heat are 0.36 and 0.01 eV, respectively. During this process, the N-O distance increases from 1.16 Å in ISNO+OH (Figure 5a) gradually to 1.18 Å in TSNO+OH (Figure 5b) and finally to 1.20 Å in the formed HNO2 in FSNO+OH (Figure 5c). Correspondingly, the Au-N distance is 2.73 Å in ISNO+OH (Figure 5a), which decreases to 2.29 Å in TSNO+OH (Figure 5b) and further to 2.16 Å to form adsorbed HNO2 in FSNO+OH (Figure 5c). The changes in N-O and Au-N distances can be attributed to the formation of Au-N and N-OH bonds. During the reaction, the Hirshfeld charge on Au decreases from 0.16 |e| in ISNO+OH to 0.12 |e| in FSNO+OH, and the Hirshfeld charge on N increases from 0.02 |e| in ISNO+OH to 0.07 |e| in FSNO+OH, indicating Au(I) is reduced and N is oxidized, and this is in accordance with the proposed mechanism that the Au(I) fragment is reduced to Au1-Pene and NO is oxidized into HNO2 adsorbed with N on Au1-Pene. These findings are further supported by the DOS analysis (Figure 6 and Figure 7). In ISNO+OH, the adsorption of NO is not strong, the molecular states of NO are apparent, and the spin is mainly localized on NO (Figure 6a). At the corresponding transition state, TSNO+OH, NO is approaching the HO-Au1-Pene, so the spin is no longer localized only on NO but also on the sp states of OH and dsp states of Au (Figure 6b). The newly appeared resonance of occupied dsp hybridized states of Au and NO states at the Fermi level indicates the formation of a Au-N bond and the reduction of Au (Figure 6b,c). This, together with the shortened Au-N distance, explains the adsorption of HNO2 (Eads: −0.55 eV). The downshift of -OH sp states and NO states also suggests the plausible formation of N-OH (Figure 6b) and the oxidation of N. Due to the bonding and charge transfer between HNO2 and Au, the spin is delocalized on Au and HNO2 in the product (Figure 6c). The calculated band gap of HNO2 adsorption structure is 0.06 eV, confirming the adsorption and reaction of NO would lead to a detectable band gap change of 0.81 eV (Figure 8a).
During the reaction of adsorbed NO2 with HO-Au1-Pene (ISNO2+OH, Figure 5d), NO2 would move to attack the O of the OH attached to Au, with the insertion of NO2 into the Au-OH by crossing TSNO2+OH (Figure 5e) to form a HNO3 adsorbed on Au1-Pene (FSNO2+OH, Figure 5f), and the calculated energy barrier and reaction heat are 0.77 and 0.61 eV, respectively. In the NO2 adsorption structure (ISNO2+OH), the Au-O(H) and N(O2)-O(H) distances are 2.02 and 2.82 Å, respectively. The sharp spikes of NO2 states are apparent, and only the DOS of NO2 and OH are spin polarized in Figure 7a, suggesting limited interaction between NO2 and HO-Au1-Pene. The approaching of NO2 leads to the increase of Au-O(H) distance to 2.40 Å at TSNO2+OH and the decrease of N(O2)-O(H) distance to 1.63 Å, showing the tendency for the breaking of the Au-O bond and the formation of the N-OH bond. This is further supported by the downshift of DOS peaks of both OH and NO2 states and the resonance among Au, OH, and NO2 states in the range from −8 to −3 eV (Figure 7). The hybridized states of Au resonate with that of NO2 at the Fermi level, confirming the tendency for the insertion of NO2 into Au-OH (Figure 7b). In the product of this step, the Au-O(H) distance increases to 2.83 Å and the N(O2)-O(H) distance decreases further to 1.47 Å, showing the Au-O bond is broken and the OH is attached to NO2, forming a HNO3 adsorbed on Au. This process is also accompanied by charge transfer. The Hirshfeld charges on Au and N change from 0.14 and 0.16 |e|, respectively, in ISNO2+OH, to 0.10 and 0.27 |e|, respectively, in FSNO2+OH, indicating that the Au species is reduced by the charge transferred from the N of NO2 that becomes oxidized. The DOS spikes of the NO2 and OH states are sharp and well separated and resonate in the range from −10 eV to the Fermi level, confirming the plausible formation of HNO3. A new spin-polarized hybridized Au state emerges at the Fermi level, confirming Au gains charge and is reduced during this process and the spin is localized on Au in FSNO2+OH (Figure 7c). To this end, the large barrier at TSNO2+OH can be attributed to the direction of charge transfer from NO2 to Au. The calculated band gap of the HNO3 adsorption structure is 0.09 eV, confirming the adsorption and reaction of NO2 would lead to a detectable band gap change of 0.80 eV (Figure 8). However, the endothermicity and the low reverse-reaction barriers make HO-Au1-Pene less eligible for NO2-sensing applications.
The reactions between NO and NO2 and H3C-Au1-Pene were also investigated (Figure 5g–l). The calculated energy barriers for the formation of H3CNO-Au1-Pene and H3CNO2-Au1-Pene are 1.53 and 1.95 eV, respectively; the calculated reaction heat are −0.48 and −0.46 eV, respectively; and the corresponding band gap changes are 0.75 and 0.79 eV, respectively (Figure 8c,d). Though exothermic, the energy barriers are too high for the reactions between NO and NO2 with H3C-Au1-Pene to take place spontaneously. Therefore, H3C-Au1-Pene is not eligible for NOx-sensing applications.
For comparison, the adsorption and reaction of O2 and H2O on HO-Au1-Pene were also investigated. H2O would form a hydrogen bond with the -OH moiety of HO-Au1-Pene, and the calculated adsorption energy is −0.30 eV, which is slightly more plausible than that on pristine Pene (−0.19 eV). O2 adsorbs physically on P atoms around Au, and the calculated adsorption energy is −0.16 eV, and its dissociation adsorption is slightly less plausible than on pristine Pene considering Au already passivated the lone pair on the P atom beneath it. The further dissociative adsorption of H2O and O2 onto P atoms of HO-Au1-Pene would experience energy barriers of 1.18 and 0.75 eV, respectively, which are also slightly higher than those on pristine Pene [54,55,56,57]. The -OH moiety cannot mediate the dissociation of O2, as this requires oxidation of Au(I) that is more demanding than O2 dissociation on Pene. Considering these high barriers and the limited adsorption energy of H2O and O2, degradation of HO-Au1-Pene would not take place in the dry air condition that is the common operating condition for NO detection. The spontaneous desorption of H2O and O2 and reaction of NO make HO-Au1-Pene highly selective for NO detection. Considering gold clusters are often poisoned by mercaptanes, the reaction of HO-Au1-Pene with H2S forming HS-Au1-Pene and H2O was also investigated. The reaction was found to be 1.03 eV exergonic, suggesting mercaptanes may potentially impact the sensing performance of HO-Au1-Pene.
Finally, the sensitivity of HO-Au1-Pene for detecting NO was also evaluated. The sensitivity was calculated as S = e x p E g E g 2 k T 1 , where k is the Boltzmann constant and E g E g corresponds to the change of band gap (0.81 eV) in existence of NO. The calculated sensitivity of HO-Au1-Pene at 300, 400, and 500 K are all 99.99%. This indicates that HO-Au1-Pene is highly sensitive, making it a promising candidate material for detection of NO. The current work demonstrates the working principle of HO-Au1-Pene as a sensor for NO detection. It is highly sensitive for NO detection with respect to recent proposed NOx sensors based on phosphorene and other 2D materials (Table 1). The change of band gap in existence of NO (0.81 eV) can in principle be detected as the change of resistance of the fabricated composites. Previously, Au and Pt clusters loaded on carbon nanotubes and black phosphorene, and synthesized black phosphorene, were reported as acceptable for NO2 sensing with a similar working principle, and the change of band gap/change of Fermi level due to adsorption/reaction of approaching molecules can be correlated with the change of resistance of the sensor [19,20,46,58].

3. Theoretical Methods

All calculations in this study were performed using the PBE functional within the generalized gradient approximation (GGA) using the DSPP potential and DNP basis set as implemented in DMol3 [63,64]. Such descriptions of electronic states are of at least double-zeta quality [65], so the calculated energies are expected to exhibit small basis-set superposition errors in principle, together with a reasonable description of weak bonds, within the limits of the theory [66]. The contribution of dispersive interactions to the energetic properties of the reaction species was examined in the benchmark calculations with empirical correction developed by Grimme et al. [67] but was found rather limited and was not considered further. Spin-polarized first-principle-based calculations were performed to investigate the deposition of a Au atom and clusters and the adsorption and reaction concerning NO and NO2. Frequency analysis was conducted to ensure that all obtained structures are stable and to derive the partition functions to calculate free energy, in which each transition state structure has only one imaginary frequency in the direction of the reaction. A 4 × 3 supercell of Pene containing 48 atoms was used to mimic the surface of Pene. A 4 × 4 × 1 k-point grid was used to sample the Brillouin zone [68], and the global cutoff radius was 4.50 Å. The convergence criteria for energy and force were set to 1 × 10−5 Ha and 2 × 10−3 Ha/Å, respectively. With these, the lattice parameters of Pene were calculated to be 4.65 and 3.32 Å, respectively [52,69], and the bulk band gap of Pene was calculated to be 0.87 eV, comparing well with recent theoretical investigations [48,52,53]. The adsorption energy of NO, NO2, H2O, and O2 on pristine Pene were calculated to be −0.21, −0.22, −0.19, and −0.16 eV, respectively, and the calculated energy barriers for H2O and O2 dissociation were 1.08 and 0.69 eV, respectively. These data, though without empirical correction developed by Grimme et al. [67], vary only within 0.05 eV from those with dispersive interaction included and agree well with reported results, showing the current theoretical approach is already adequate for the investigation [40,54,55,56,57].

4. Conclusions

Inspired by the potential significant response of single-site heterogeneous reaction centers, in terms of large band-gap change, etc., to approaching molecules that can be used for sensing applications, we investigated the electronic structure and reaction pathways of Au(I) fragments dispersed on Pene with NO and NO2 with extensive first-principle-based calculations. Au1-Pene has hybridized Au states in the bulk band gap of Pene and a decreased band gap of 0.14 eV and would aggregate into clusters. Passivation with -OH and -CH3 forms thermodynamically plausible HO-Au1-Pene and H3C-Au1-Pene and restores the band gap to that of bulk Pene. HO-Au1-Pene and H3C-Au1-Pene were also examined for detection of NO and NO2 that would react with -OH and -CH3, and the resulting decrease of band gap back to that of Au1-Pene would be measurable. HO-Au1-Pene and H3C-Au1-Pene are highly sensitive to NO and NO2, and the calculated theoretical sensitivity are all 99.99%. The calculated energy barrier for the reaction of HO-Au1-Pene with NO is 0.36 eV, and the reaction is about thermal neutral, suggesting HO-Au1-Pene can be used as NO-sensing material, and the reaction for NO detection would be spontaneous. The reaction of NO2 with HO-Au1-Pene is endothermic, making the dissociation of product HNO3 more plausible, while the barriers for the reaction of CH3-Au1-Pene with NO and NO2 are too high for spontaneous detection. Therefore, HO-Au1-Pene is not eligible for NO2 sensing and CH3-Au1-Pene is not eligible for NO and NO2 sensing. The reaction of H2S with HO-Au1-Pene is 1.03 eV exothermic, suggesting HO-Au1-Pene should not be used in the existence of mercaptanes. The current work highlights the working principle of HO-Au1-Pene as a sensor for NO detection. Such change of band gap/Fermi level in existence of NO (0.81 eV) can in principle be detected as the change of resistance of the fabricated composites, especially when Au fragments are loaded on Pene nanoribbons. Au clusters loaded on carbon nanotubes and Pene were reported as acceptable for gas sensing, and the change of electronic structure upon gas adsorption can be correlated with the change of resistance of the sensor [58]. We expect the findings would pave the way for the design and application of single-site heterogeneous reaction centers for sensing applications.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30153085/s1, cartesian coordinates of Au fragments on Pene in xyz format.

Author Contributions

H.G. and X.L. conceived the research and drafted the manuscript. Y.L. performed the investigation and is responsible for the results disclosed. H.G. and X.L. obtained funding, provided research instructions and tools, and commented on the results and the manuscript. The manuscript was revised through contributions of all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Nos. 21771029, 11811530631, 21573034, 21373036, and 21103015) and the Fundamental Research Funds for the Central Universities (DUT22LAB602).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding author.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (Nos. 21771029, 11811530631, 21573034, 21373036, and 21103015) and the Fundamental Research Funds for the Central Universities (DUT22LAB602).

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Drewniak, S.; Drewniak, L.; Pustelny, T. Mechanisms of NO2 Detection in Hybrid Structures Containing Reduced Graphene Oxide: A Review. Sensors 2022, 22, 5316. [Google Scholar] [CrossRef]
  2. Lee, H.Y.; Heish, Y.C.; Lee, C.T. High Sensitivity Detection of Nitrogen Oxide gas at Room Temperature using Zinc Oxide-Reduced Graphene Oxide Sensing Membrane. J. Alloy Compd. 2019, 773, 950–954. [Google Scholar] [CrossRef]
  3. Li, J.; Yang, M.; Cheng, X.; Zhang, X.; Guo, C.; Xu, Y.; Gao, S.; Major, Z.; Zhao, H.; Huo, L. Fast Detection of NO2 by Porous SnO2 Nanotoast Sensor at Low Temperature. J. Hazard. Mater. 2021, 419, 126414. [Google Scholar] [CrossRef] [PubMed]
  4. Wojcika, B.U.; Vincenta, T.A.; Chowdhuryb, M.F.; Gardner, J.W. Ultrasensitive WO3 Gas Sensors for NO2 Detection in Air and Low Oxygen Environment. Sens. Actuator B Chem. 2017, 239, 1051–1059. [Google Scholar] [CrossRef]
  5. Yong, Y.; Su, X.; Cui, H.; Zhou, Q.; Kuang, Y.; Li, X. Two-Dimensional Tetragonal GaN as Potential Molecule Sensors for NO and NO2 Detection: A First-Principle Study. ACS Omega 2017, 2, 8888–8895. [Google Scholar] [CrossRef] [PubMed]
  6. Buckley, D.J.; Black, N.C.G.; Castanon, E.G.; Melios, C.; Hardman, M.; Kazakova, O. Frontiers of Graphene and 2D Material-based Gas Sensors for Environmental Monitoring. 2D Mater. 2020, 7, 032002. [Google Scholar] [CrossRef]
  7. Chen, T.; Cheng, Z.; Tian, Q.; Wang, J.; Yu, X.; Ho, D. Nitrogen Dioxide Gas Sensor Based on Liquid-Phase-Exfoliated Black Phosphorus Nanosheets. ACS Appl. Nano Mater. 2020, 3, 6440–6447. [Google Scholar] [CrossRef]
  8. Donarelli, M.; Ottaviano, L. 2D Materials for Gas Sensing Applications: A Review on Graphene Oxide, MoS2, WS2 and Phosphorene. Sensors 2018, 18, 3638. [Google Scholar] [CrossRef]
  9. Kumar, S.; Pavelyev, V.; Mishra, P.; Tripathi, N.; Sharma, P.; Calle, F. A Review on 2D Transition Metal Di-chalcogenides and Metal Oxide Nanostructures based NO2 Gas Sensors. Mater. Sci. Semicond. Process. 2020, 107, 104865. [Google Scholar] [CrossRef]
  10. Louis, H.; Egemonye, T.C.; Unimuke, T.O.; Inah, B.E.; Edet, H.O.; Eno, E.A.; Adalikwu, S.A.; Adeyinka, A.S. Detection of Carbon, Sulfur, and Nitrogen Dioxide Pollutants with a 2D Ca12O12 Nanostructured Material. ACS Omega 2022, 7, 34929–34943. [Google Scholar] [CrossRef]
  11. Xu, Y.; Zheng, W.; Liu, X.; Zhang, L.; Zheng, L.; Yang, C.; Pinna, N.; Zhang, J. Platinum Single Atoms on Tin Oxide Ultrathin Films for Extremely Sensitive Gas Detection. Mater. Horizons 2020, 7, 1519–1527. [Google Scholar] [CrossRef]
  12. Sun, X.; Lan, Q.; Geng, J.; Yu, M.; Li, Y.; Li, X.; Chen, L. Polyoxometalate as Electron Acceptor in Dye/TiO2 Films to Accelerate Room-temperature NO2 Gas Sensing. Sens. Actuator B Chem. 2023, 374, 132795. [Google Scholar] [CrossRef]
  13. Aghaei, S.M.; Monshi, M.M.; Torres, I.; Zeidi, S.M.J.; Calizo, I. DFT study of adsorption behavior of NO, CO, NO2, and NH3 molecules on graphene-like BC3: A search for highly sensitive molecular sensor. Appl. Surf. Sci. 2018, 427, 326–333. [Google Scholar] [CrossRef]
  14. Gusmao, R.; Sofer, Z.; Pumera, M. Black Phosphorus Rediscovered: From Bulk Material to Monolayers. Angew. Chem. Int. Edit. 2017, 56, 8052–8072. [Google Scholar] [CrossRef] [PubMed]
  15. Jia, Q.; Kong, X.; Qiao, J.; Ji, W. Strain- and Twist-engineered Optical Absorption of Few-layer Black Phosphorus. Sci. China Phys. Mech. Astron. 2016, 59, 696811. [Google Scholar] [CrossRef]
  16. Kou, L.; Chen, C.; Smith, S.C. Phosphorene: Fabrication, Properties, and Applications. J. Phys. Chem. Lett. 2015, 6, 2794–2805. [Google Scholar] [CrossRef]
  17. Lei, S.; Wang, H.; Huang, L.; Sun, Y.Y.; Zhang, S. Stacking Fault Enriching the Electronic and Transport Properties of Few-Layer Phosphorenes and Black Phosphorus. Nano Lett. 2016, 16, 1317–1322. [Google Scholar] [CrossRef]
  18. Lee, G.; Kim, S.; Jung, S.; Jang, S.; Kim, J. Suspended Black Phosphorus Nanosheet Gas Sensors. Sens. Actuator B Chem. 2017, 250, 569–573. [Google Scholar] [CrossRef]
  19. Cho, S.Y.; Lee, Y.; Koh, H.J.; Jung, H.; Kim, J.S.; Yoo, H.W.; Kim, J.; Jung, H.T. Superior Chemical Sensing Performance of Black Phosphorus: Comparison with MoS2 and Graphene. Adv. Mater. 2016, 28, 7020–7028. [Google Scholar] [CrossRef]
  20. Cui, S.; Pu, H.; Wells, S.A.; Wen, Z.; Mao, S.; Chang, J.; Hersam, M.C.; Chen, J. Ultrahigh Sensitivity and Layer-Dependent Sensing Performance of Phosphorene-based Gas Sensors. Nat. Commun. 2015, 6, 8632. [Google Scholar] [CrossRef]
  21. Zhang, H.P.; Kou, L.; Jiao, Y.; Du, A.; Tang, Y.; Ni, Y. Strain Engineering of Selective Chemical Adsorption on Monolayer Black Phosphorous. Appl. Surf. Sci. 2020, 503, 144033. [Google Scholar] [CrossRef]
  22. Kumawat, R.L.; Pathak, B. Prospects of Black Phosphorus Nanoribbon for Explosive Sensing: A Computational Approach. Appl. Surf. Sci. 2020, 529, 147094. [Google Scholar] [CrossRef]
  23. Ghashghaee, M.; Ghambarian, M. Defect Engineering and Zinc Oxide Doping of Black Phosphorene for Nitrogen Dioxide Capture and Detection: Quantum-chemical Calculations. Appl. Surf. Sci. 2020, 523, 146527. [Google Scholar] [CrossRef]
  24. Kaewmaraya, T.; Ngamwongwan, L.; Moontragoon, P.; Karton, A.; Hussain, T. Drastic Improvement in Gas-Sensing Characteristics of Phosphorene Nanosheets under Vacancy Defects and Elemental Functionalization. J. Phys. Chem. C 2018, 122, 20186–20193. [Google Scholar] [CrossRef]
  25. Meshginqalam, B.; Barvestani, J. Vacancy Defected Blue and Black Phosphorene Nanoribbons as Gas Sensor of NOx and SOx Molecules. Appl. Surf. Sci. 2020, 526, 146692. [Google Scholar] [CrossRef]
  26. Moghaddaszadeh, Z.; Toosi, M.R.; Zardoost, M.R.; Arabieh, M. First-principle Study of the Adsorption of Volatile Sulfur Compounds on Black Phosphorene Nanosheets Doped with Some Transition Metals. Mon. Chem. 2020, 151, 1501–1510. [Google Scholar] [CrossRef]
  27. Ou, P.; Song, P.; Liu, X.; Song, J. Superior Sensing Properties of Black Phosphorus as Gas Sensors: A Case Study on the Volatile Organic Compounds. Adv. Theory Simul. 2019, 2, 1800103. [Google Scholar] [CrossRef]
  28. Valt, M.; Caporali, M.; Fabbri, B.; Gaiardo, A.; Krik, S.; Iacob, E.; Vanzetti, L.; Malagu, C.; Banchelli, M.; D’Andrea, C.; et al. Air Stable Nickel-Decorated Black Phosphorus and Its Room-Temperature Chemiresistive Gas Sensor Capabilities. ACS Appl. Mater. Interfaces 2021, 13, 44711–44722. [Google Scholar] [CrossRef]
  29. Yang, D.; Han, N.; Gao, R.; Cheng, Y. Metal doped Black Phosphorene for Gas Sensing and Catalysis: A First-principles Perspective. Appl. Surf. Sci. 2022, 586, 152743. [Google Scholar] [CrossRef]
  30. Ghadiri, M.; Ghashghaee, M.; Ghambarian, M. Influence of NiO Decoration on Adsorption Capabilities of Black Phosphorus Monolayer toward Nitrogen Dioxide: Periodic DFT Calculations. Mol. Simul. 2020, 46, 1062–1072. [Google Scholar] [CrossRef]
  31. Liang, T.; Dai, Z.; Liu, Y.; Zhang, X.; Zeng, H. Suppression of Sn2+ and Lewis acidity in SnS2/black phosphorus heterostructure for ppb-level room temperature NO2 gas sensor. Sci. Bull. 2021, 66, 2471–2478. [Google Scholar] [CrossRef] [PubMed]
  32. Liu, J.; Zhang, C.; Wang, Y.; Chen, X.; Jing, R.; Song, T.; Zhang, Z.; Wang, H.; Fu, C.; Wang, J. Black Phosphorus Nanodot Incorporated Tin Oxide Hollow-spherical Heterojunction for Enhanced Properties of Room-Temperature Gas Sensors. Ceram. Int. 2023, 49, 8248–8258. [Google Scholar] [CrossRef]
  33. Lei, S.Y.; Shen, H.Y.; Sun, Y.Y.; Wan, N.; Yu, H.; Zhang, S. Enhancing the ambient stability of few-layer black phosphorus by surface modification. RSC Adv. 2018, 8, 14676–14683. [Google Scholar] [CrossRef]
  34. Lei, S.Y.; Sun, X.L.; Shen, H.Y. The Effect of Adatom Concentration on the Oxidation Reaction of Phosphorene. Mater. Res. Express 2019, 6, 025905. [Google Scholar] [CrossRef]
  35. Lei, S.Y.; Guo, S.J.; Sun, X.L.; Yu, H.; Xu, F.; Wan, N.; Wang, Z.A. Capture and dissociation of dichloromethane on Fe, Ni, Pd and Pt decorated phosphorene. Appl. Surf. Sci. 2019, 495, 143533. [Google Scholar] [CrossRef]
  36. Lei, S.Y.; Yu, Z.Y.; Shen, H.Y.; Sun, X.L.; Wan, N.; Yu, H. CO Adsorption on Metal-Decorated Phosphorene. Acs Omega 2018, 3, 3957–3965. [Google Scholar] [CrossRef]
  37. Wang, M.Y.; Jiang, Y.C.; Zhang, Q.; Tao, J.N.; Hussain, S.; Liu, G.W.; Wan, N.; Lei, S.Y. Study on the Properties of the Ferroelectric Photocatalytic Hydrogen Evolution Reaction for Chemically Functionalized Phosphorene Nanosheets. ACS Appl. Energ. Mater. 2023, 6, 7985–7995. [Google Scholar] [CrossRef]
  38. Wang, M.Y.; Shi, H.; Tian, M.; Chen, R.W.; Shu, J.P.; Zhang, Q.; Wang, Y.H.; Li, C.Y.; Wan, N.; Lei, S.Y. Single Nickel Atom-Modified Phosphorene Nanosheets for Electrocatalytic CO2 Reduction. ACS Appl. Nano Mater. 2021, 4, 11017–11030. [Google Scholar] [CrossRef]
  39. Wang, Y.H.; Lei, S.Y.; Gao, R.; Sun, X.L.; Chen, J. Effect of metal decoration on sulfur-based gas molecules adsorption on phosphorene. Sci Rep. 2021, 11, 18179. [Google Scholar] [CrossRef]
  40. Lei, S.Y.; Gao, R.; Sun, X.L.; Guo, S.J.; Yu, H.; Wan, N.; Xu, F.; Chen, J. Nitrogen-based Gas Molecule Adsorption of Monolayer Phosphorene under Metal Functionalization. Sci Rep. 2019, 9, 12498. [Google Scholar] [CrossRef]
  41. Wang, M.Y.; Li, L.H.; Zhao, G.H.; Xu, Z.W.; Hussain, S.; Wang, M.S.; Qiao, G.J.; Liu, G.W. Influence of the Surface Decoration of Phosphorene with Ag Nanoclusters on Gas Sensing Properties. Appl. Surf. Sci. 2020, 504, 144374. [Google Scholar] [CrossRef]
  42. Hu, J.X.; Zhao, L.; Du, J.G.; Jiang, G. Adsorption of rare gases on pristine and doped phosphorene. Appl. Surf. Sci. 2020, 504, 144326. [Google Scholar] [CrossRef]
  43. Ghadiri, M.; Ghashghaee, M.; Ghambarian, M. Mn-Doped Black Phosphorene for Ultrasensitive Hydrogen Sulfide Detection: Periodic DFT Calculations. Phys. Chem. Chem. Phys. 2020, 22, 15549–15558. [Google Scholar] [CrossRef] [PubMed]
  44. Ghambarian, M.; Azizi, Z.; Ghashghaee, M. Hydrogen Detection on Black Phosphorene doped with Ni, Pd, and Pt: Periodic Density Functional Calculations. Int. J. Hydrog. Energy 2020, 45, 16298–16309. [Google Scholar] [CrossRef]
  45. Wang, Y.; Zhou, Y.; Li, J.; Zhang, R.; Zhao, H.; Wang, Y. Ag Decoration-enabled Sensitization Enhancement of Black Phosphorus Nanosheets for Trace NO2 Detection at Room Temperature. J. Hazard. Mater. 2022, 435, 129086. [Google Scholar] [CrossRef] [PubMed]
  46. Cho, S.Y.; Koh, H.J.; Yoo, H.W.; Jung, H.T. Tunable Chemical Sensing Performance of Black Phosphorus by Controlled Functionalization with Noble Metals. Chem. Mat. 2017, 29, 7197–7205. [Google Scholar] [CrossRef]
  47. Liu, B.; Zhu, Q.; Pan, Y.; Huang, F.; Tang, L.; Liu, C.; Cheng, Z.; Wang, P.; Ma, J.; Ding, M. Single-Atom Tailoring of Two-Dimensional Atomic Crystals Enables Highly Efficient Detection and Pattern Recognition of Chemical Vapors. ACS Sensors 2022, 7, 1533–1543. [Google Scholar] [CrossRef]
  48. Liang, Z.; Wang, M.; Zhang, X.; Li, Z.; Du, K.; Yang, J.; Lei, S.-Y.; Qiao, G.; Ou, J.Z.; Liu, G. A 2D-0D-2D Sandwich Heterostructure toward High-performance Room-temperature Gas Sensing. ACS Nano 2024, 18, 3669–3680. [Google Scholar] [CrossRef]
  49. Xue, Z.; Yan, M.; Yu, X.; Tong, Y.; Zhou, H.; Zhao, Y.; Wang, Z.; Zhang, Y.; Xiong, C.; Yang, J. One-dimensional Segregated Single Au Sites on Step-rich ZnO Ladder for Ultrasensitive NO2 sensors. Chem 2020, 6, 3364–3373. [Google Scholar] [CrossRef]
  50. Kulish, V.V.; Malyi, O.I.; Persson, C.; Wu, P. Adsorption of Metal Adatoms on Single-layer Phosphorene. Phys. Chem. Chem. Phys. 2015, 17, 992–1000. [Google Scholar] [CrossRef]
  51. Moschetto, S.; Ienco, A.; Manca, G.; Serrano-Ruiz, M.; Peruzzini, M.; Mezzi, A.; Brucale, M.; Bolognesi, M.; Toffanin, S. Easy and Fast In Situ Functionalization of Exfoliated 2D Black Phosphorus with Gold Nanoparticles. Dalton Trans. 2021, 50, 11610–11618. [Google Scholar] [CrossRef]
  52. Liu, X.; Chen, L.K.; Wu, Y.; Zhang, X.; Chambaud, G.; Han, Y.; Meng, C.G. Pd Speciation on Black Phosphorene in a CO and C2H4 Atmosphere: A First-principles Investigation. Phys. Chem. Chem. Phys. 2022, 24, 14284–14293. [Google Scholar] [CrossRef]
  53. Carvalho, A.; Wang, M.; Zhu, X.; Rodin, A.S.; Su, H.B.; Neto, A.H.C. Phosphorene: From Theory to Applications. Nat. Rev. Mater. 2016, 1, 16061. [Google Scholar] [CrossRef]
  54. Benini, F.; Bassoli, N.; Restuccia, P.; Ferrario, M.; Righi, M.C. Interaction of Water and Oxygen Molecules with Phosphorene: An Ab Initio Study. Molecules 2023, 28, 3570. [Google Scholar] [CrossRef] [PubMed]
  55. Huang, Y.; Qiao, J.; He, K.; Bliznakov, S.; Sutter, E.; Chen, X.; Luo, D.; Meng, F.; Su, D.; Decker, J.; et al. Interaction of Black Phosphorus with Oxygen and Water. Chem. Mat. 2016, 28, 8330–8339. [Google Scholar] [CrossRef]
  56. Luo, L.H.; Luo, J.; Li, H.L.; Ren, F.N.; Zhang, Y.F.; Liu, A.D.; Li, W.X.; Zeng, J. Water Enables Mild Oxidation of Methane to Methanol on Gold Single-atom Catalysts. Nat. Commun. 2021, 12, 1218. [Google Scholar] [CrossRef]
  57. Zhang, T.M.; Wan, Y.Y.; Xie, H.Y.; Mu, Y.; Du, P.W.; Wang, D.; Wu, X.J.; Ji, H.X.; Wan, L.J. Degradation Chemistry and Stabilization of Exfoliated Few-Layer Black Phosphorus in Water. J. Am. Chem. Soc. 2018, 140, 7561–7567. [Google Scholar] [CrossRef]
  58. Zanolli, Z.; Leghrib, R.; Felten, A.; Pireaux, J.J.; Llobet, E.; Charlier, J.C. Gas Sensing with Au-Decorated Carbon Nanotubes. ACS Nano 2011, 5, 4592–4599. [Google Scholar] [CrossRef]
  59. Cai, Y.Q.; Ke, Q.Q.; Zhang, G.; Zhang, Y.W. Energetics, Charge Transfer, and Magnetism of Small Molecules Physisorbed on Phosphorene. J. Phys. Chem. C 2015, 119, 3102–3110. [Google Scholar] [CrossRef]
  60. Wang, Z.Y.; Wu, H.; Wu, Q.Y.; Zhao, Y.M.; Shen, L. Magnetic e-Phosphorene for Sensing Greenhouse Gas Molecules. Molecules 2023, 28, 5402. [Google Scholar] [CrossRef]
  61. Abbas, A.N.; Liu, B.L.; Chen, L.; Ma, Y.Q.; Cong, S.; Aroonyadet, N.; Köpf, M.; Nilges, T.; Zhou, C.W. Black Phosphorus Gas Sensors. ACS Nano 2015, 9, 5618–5624. [Google Scholar] [CrossRef]
  62. He, Q.Y.; Zeng, Z.Y.; Yin, Z.Y.; Li, H.; Wu, S.X.; 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]
  63. Delley, B. From Molecules to Solids with the DMol3 Approach. J. Chem. Phys. 2000, 113, 7756–7764. [Google Scholar] [CrossRef]
  64. Delley, B. Hardness Conserving Semilocal Pseudopotentials. Phys. Rev. B 2002, 66, 155125. [Google Scholar] [CrossRef]
  65. Becke, A.D. A Multicenter Numerical-integration Scheme for Polyatomic-molecules. J. Chem. Phys. 1988, 88, 2547–2553. [Google Scholar] [CrossRef]
  66. Inada, Y.; Orita, H. Efficiency of numerical basis sets for predicting the binding energies of hydrogen bonded complexes: Evidence of small basis set superposition error compared to Gaussian basis sets. J. Comput. Chem. 2008, 29, 225–232. [Google Scholar] [CrossRef] [PubMed]
  67. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  68. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-zone Integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  69. Ienco, A.; Manca, G.; Peruzzini, M.; Mealli, C. Modelling strategies for the covalent functionalization of 2D phosphorene. Dalton Trans. 2018, 47, 17243–17256. [Google Scholar] [CrossRef]
Figure 1. Structures (ac) and energetic properties (d) of Au1 (a), Au2 (b), and Au3 (c) clusters deposited on Pene. In (ac), P and Au atoms are in purple and gold, respectively.
Figure 1. Structures (ac) and energetic properties (d) of Au1 (a), Au2 (b), and Au3 (c) clusters deposited on Pene. In (ac), P and Au atoms are in purple and gold, respectively.
Molecules 30 03085 g001
Figure 2. Band structure of pristine Pene (a), Au1-Pene (b), HO-Au1-Pene (c), and H3C-Au1-Pene (d). In (ad), energy “0” corresponds to calculated Fermi level of the system. In (b), the bands of spins are in red and black, respectively.
Figure 2. Band structure of pristine Pene (a), Au1-Pene (b), HO-Au1-Pene (c), and H3C-Au1-Pene (d). In (ad), energy “0” corresponds to calculated Fermi level of the system. In (b), the bands of spins are in red and black, respectively.
Molecules 30 03085 g002
Figure 3. Side (a) and top (a’) views of structure HO-Au1-Pene, top view of structures of potential dimers of HO-Au1-Pene (be), and formation free energy (f) of HO-Au1-Pene and its potential dimers. In (ae), P, Au, O, and H atoms are in purple, gold, red, and white, respectively.
Figure 3. Side (a) and top (a’) views of structure HO-Au1-Pene, top view of structures of potential dimers of HO-Au1-Pene (be), and formation free energy (f) of HO-Au1-Pene and its potential dimers. In (ae), P, Au, O, and H atoms are in purple, gold, red, and white, respectively.
Molecules 30 03085 g003
Figure 4. Side (a) and top (a’) views of structure H3C-Au1-Pene, top view of structures of potential dimers of H3C-Au1-Pene (be), and formation free energy (f) of H3C-Au1-Pene and its potential dimers. In (ae), P, Au, C, and H atoms are in purple, gold, gray, and white, respectively.
Figure 4. Side (a) and top (a’) views of structure H3C-Au1-Pene, top view of structures of potential dimers of H3C-Au1-Pene (be), and formation free energy (f) of H3C-Au1-Pene and its potential dimers. In (ae), P, Au, C, and H atoms are in purple, gold, gray, and white, respectively.
Molecules 30 03085 g004
Figure 5. Side view (top panel) and top view (lower panel) of species involved in reactions between NO and HO-Au1-Pene (ac), NO2 and HO-Au1-Pene (df), NO and H3C-Au1-Pene (gi), and NO2 and H3C-Au1-Pene (jl). In (al), species named IS and FS are the initial and final species involved in the elementary reactions and species named TS are the corresponding transition states. In (al), the N, P, C, O, Au and H are in blue, purple, gray, red, yellow, and white, respectively.
Figure 5. Side view (top panel) and top view (lower panel) of species involved in reactions between NO and HO-Au1-Pene (ac), NO2 and HO-Au1-Pene (df), NO and H3C-Au1-Pene (gi), and NO2 and H3C-Au1-Pene (jl). In (al), species named IS and FS are the initial and final species involved in the elementary reactions and species named TS are the corresponding transition states. In (al), the N, P, C, O, Au and H are in blue, purple, gray, red, yellow, and white, respectively.
Molecules 30 03085 g005
Figure 6. PDOS of ISNO+OH (a), TSNO+OH (b), and FSNO+OH (c).
Figure 6. PDOS of ISNO+OH (a), TSNO+OH (b), and FSNO+OH (c).
Molecules 30 03085 g006
Figure 7. PDOS of ISNO2+OH (a), TSNO2+OH (b), and FSNO2+OH (c).
Figure 7. PDOS of ISNO2+OH (a), TSNO2+OH (b), and FSNO2+OH (c).
Molecules 30 03085 g007
Figure 8. Band structure of pristine HNO2-Au1-Pene (a), HNO3-Au1-Pene (b), H3CNO-Au1-Pene (c), and H3CNO2-Au1-Pene (d). In (ad), energy “0” corresponds to calculated Fermi level of the system, and the bands of spins are in red and black, respectively.
Figure 8. Band structure of pristine HNO2-Au1-Pene (a), HNO3-Au1-Pene (b), H3CNO-Au1-Pene (c), and H3CNO2-Au1-Pene (d). In (ad), energy “0” corresponds to calculated Fermi level of the system, and the bands of spins are in red and black, respectively.
Molecules 30 03085 g008
Table 1. Sensing performance of HO-Au1-Pene with respect to reported 2D sensors.
Table 1. Sensing performance of HO-Au1-Pene with respect to reported 2D sensors.
SensorAnalyteΔEg (eV)Sensitivity (%)Detection LimitResponse Time (s)Reference
Pene-Au1-OHNO0.8199.9 This work
PeneNO0.2326.7 [59]
PeneNO20.5462.8 [59]
V-ε-PeneNO0.5075 [60]
V-ε-PeneNO20.3042.5 [60]
Au doped PeneNO0.78 [29]
Ag doped PeneNO0.80 [29]
B3CNO0.7399.6 [13]
B3CNO20.045.5 [13]
Pene FETNO2 67100 ppb75[19]
MoS2NO2 93 ~220[19]
GrapheneNO2 13 ~240[19]
PeneNO2 >95 ~300[46]
Pene/AuNPsNO2 1 [46]
Pene/PtNPsNO2 >90 [46]
Pene FETNO2 2640 ppb500[61]
Pene FETNO20.281600100 ppb600[20]
Pene FETNO2 88100 ppb290[7]
MWCNT/AuNPsNO2 70.5 ppm [58]
MoS2/PtNPsNO2 20.025 ppb [62]
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

Guo, H.; Liu, Y.; Liu, X. Phosphorene-Supported Au(I) Fragments for Highly Sensitive Detection of NO. Molecules 2025, 30, 3085. https://doi.org/10.3390/molecules30153085

AMA Style

Guo H, Liu Y, Liu X. Phosphorene-Supported Au(I) Fragments for Highly Sensitive Detection of NO. Molecules. 2025; 30(15):3085. https://doi.org/10.3390/molecules30153085

Chicago/Turabian Style

Guo, Huimin, Yuhan Liu, and Xin Liu. 2025. "Phosphorene-Supported Au(I) Fragments for Highly Sensitive Detection of NO" Molecules 30, no. 15: 3085. https://doi.org/10.3390/molecules30153085

APA Style

Guo, H., Liu, Y., & Liu, X. (2025). Phosphorene-Supported Au(I) Fragments for Highly Sensitive Detection of NO. Molecules, 30(15), 3085. https://doi.org/10.3390/molecules30153085

Article Metrics

Back to TopTop