Next Article in Journal
YBa2Cu3Oy Superconducting Ceramics Incorporated with Different Types of Oxide Materials as Promising Radiation Shielding Materials: Investigation of The Structure, Morphology, and Ionizing Radiations Shielding Performances
Next Article in Special Issue
A Review on Montmorillonite-Based Nanoantimicrobials: State of the Art
Previous Article in Journal
Special Issue “ALD Technique for Functional Coatings of Nanostructured Materials”
Previous Article in Special Issue
Molecular Fingerprinting of the Omicron Variant Genome of SARS-CoV-2 by SERS Spectroscopy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bismuth Sulfide Doped in Graphitic Carbon Nitride Degrades Nitric Oxide under Solar Irradiation

1
Institute of Aquatic Science and Technology, National Kaohsiung University of Science and Technology, Kaohsiung 811213, Taiwan
2
Ph.D. Program in Maritime Science and Technology, College of Maritime, National Kaohsiung University of Science and Technology, Kaohsiung 81157, Taiwan
3
Department of Marine Environmental Engineering, National Kaohsiung University of Science and Technology, Kaohsiung 81157, Taiwan
4
Super Micro Mass Research and Technology Center, Cheng Shiu University, Kaohsiung City 8333031, Taiwan
5
Center for Environmental Toxin and Emerging-Contaminant Research, Cheng Shiu University, Kaohsiung City 8333031, Taiwan
6
Department of Chemical Engineering and Materials Science, Environmental Technology Research Center, Yuan Ze University, Taoyuan City 32003, Taiwan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(19), 3482; https://doi.org/10.3390/nano12193482
Submission received: 16 September 2022 / Revised: 29 September 2022 / Accepted: 3 October 2022 / Published: 5 October 2022
(This article belongs to the Special Issue Nanomaterials and 2D Materials Based on Semiconductors and Metals)

Abstract

:
This study developed and examined the application of bismuth sulfide doped on graphitic carbon nitride (Bi2S3@g-C3N4) in the degradation of NO under solar irradiation. Bi2S3@g-C3N4 was prepared through the calcination method. The morphological structure and chemical properties of the synthesized photocatalyst were analyzed before the degradation tests. After doping with Bi2S3@g-C3N4, the bandgap was reduced to 2.76 eV, which increased the absorption of solar light. As a result, the Bi2S3@g-C3N4 achieved higher NO degradation (55%) compared to pure Bi2S3 (35%) and g-C3N4 (45%). The trapping test revealed that the electrons were the primary species responsible for most of the NO degradation. The photocatalyst was stable under repeated solar irradiation, maintaining degradation efficiencies of 50% after five consecutive recycling tests. The present work offers strong evidence that Bi2S3@g-C3N4 is a stable and efficient catalyst for the photocatalytic oxidation of NO over solar irradiation.

1. Introduction

Air pollutants that are highly reactive include nitrogen oxides (NOx), usually caused by anthropogenic activities, particularly fuel combustion [1]. There are two approaches to controlling NOx emissions: primary measures to prevent NOx formation and secondary measures to reduce the NOx already formed [2]. Common primary measures include staged combustion [3] and exhaust gas recirculation [4], while secondary measures include the adoption of pollution control devices including selective catalytic reduction [5] and photocatalysts [6].
Photocatalytic oxidation (PCO), one of the most effective and easiest secondary control technologies for the removal of NOx emissions, has been researched in the last twenty years [7,8,9]. Titanium dioxide (TiO2) has been frequently used as a photocatalyst in NO degradation [10] because of its long-term durability, non-toxicity, PCO activity, and long-term photostability [11,12]. However, it has several limitations including a large band gap of 3.0–3.2 eV, and it is only effective under UV light at a wavelength of <380 nm [13,14].
The graphitic carbon nitride (g-C3N4) semiconductor is a polymeric narrow-band-gap metal-free material that is functional under solar light and has gained increasing popularity [15,16] because of its unique characteristics including high thermal stability, low cost, optical properties, easy preparation, and electron transfer capability [17]. However, in g- C3N4, electrons and holes easily recombine with each other resulting in low photocatalytic activity [18]. The problem has been addressed through doping with: metals, e.g., Au [19], Fe [20], Pd [21], and Ag [22]; nonmetals, such as P [23], I [24], B [25], and S [26]; and heterojunction composites, e.g., BiPO4 and Bi5Nb3O15 [27,28].
This study investigates the outcome of doping bismuth sulfide (Bi2S3) into g-C3N4 and the application of the resultant photocatalyst, Bi2S3@g-C3N4, on the degradation of NO under solar irradiation. Bi2S3 is among the sulfide semiconductors with a characteristically large absorption coefficient and narrow bandgap energy [29,30,31]. This study shows that this new photocatalytic material has the potential for application in the degradation of NO in both indoor and outdoor environments.

2. Materials and Methods

2.1. Chemicals

The lab water purification system (ELGA Lab Water, UK) provided deionized water, bismuth (III) nitrate pentahydrate (Bi(NO3)3·5H2O) was obtained from FERAK® (Berlin, Germany), thiourea 99% (CH4N2S) and urea (CH4N2O) were obtained from Alfa Aesar (Lancaster, UK), and the Ming Yang special gas company supplied nitric oxide (100 ppm) compressed gas. All reagents were analytical grade.

2.2. Synthesis of Photocatalysts

2.2.1. Synthesis of Bi2S3

Bismuth sulfide (Bi2S3) was prepared by the calcination method. First, 0.2512 g of CH4N2S and 0.9711 g of Bi(NO3)3·5H2O were placed in a mortar and crushed by a pestle for 30 min; then, the mixture was calcinated in a muffle furnace for 2 h at 550 °C. At room temperature, after naturally cooling down, the resulting material was washed 5 times with purified water and ethyl alcohol to eliminate impurities and then dried for 3 h at 80 °C in the oven [32].

2.2.2. Synthesis of g-C3N4 and Bi2S3@g-C3N4

In order to prepare pure g-C3N4, we mixed and crushed 3.0121 g of CH4N2S and 10.0113 g of CH4N2O in a mortar for 30 min and calcinated the mixture in a muffled furnace oven for 2 h at 550 °C. The same procedure was followed while preparing Bi2S3@g-C3N4 by the addition of 0.4851 g of Bi(NO3)3·5H2O.

2.3. Characterization Procedure

Analyses of the crystallinity and composition of the synthesized materials were performed using X-ray diffractometers (Panalytical Empyrean diffractometer) under Cu Kα radiation (K = 1.5418 Å). A transmission electron microscope was used to inspect the morphological observations and structures (JEOL, JEM-2000FXII), along with a 10 kV scanning electron microscope (SEM). Fourier-transform infrared spectroscopy (FTIR) (JASCO-4700) was utilized to examine the chemical binding variations. The chemical or analytical and optical properties were determined by X-ray photoelectron spectroscopy (XPS) (Thermo Fisher Scientific, Waltham, MA, USA) using K-Alpha and energy dispersive X-ray spectrometry (EDS) (JEOL S4800), respectively.

2.4. Experiments Evaluating Photocatalytic Activity

2.4.1. Experimental Procedure

Figure 1 shows an illustration of the lab-scale experimental setup for the NO degradation. Under solar irradiation (λ > 300 nm), using a 300W Xenon lamp as the light source, the synthesized photocatalysts were used to study NO degradation by photocatalysis at ppb levels. The photocatalysts were placed in a 4.5 L stainless steel reaction chamber (10 cm × 15 cm × 30 cm) with a glass on top to allow light to reach the surface of the material. Above the reactor, the lamp was vertically positioned at a distance of 40 cm. The initial NO concentration (500 ppb) was adjusted using air supplied by a zero-air generator, and in the reaction chamber, the gas flowed at a rate of 3 L min−1 at 70% relative humidity before flowing into the reaction chamber. By streaming gas into the reaction chamber without light, an equilibrium was achieved between adsorption and desorption. After equilibrium, the degradation experiment started by switching on the light source. The initial and final concentrations of NO were determined by an NOx analyzer (Thermo Scientific, Model 42i). The photocatalytic activity was finished when the minimum level of NO concentration was achieved.

2.4.2. Photocatalytic Activity

In each experiment, 0.2 g of the prepared photocatalyst was dispersed into 20 mL of deionized water in a glass dish (d = 12 cm), rinsed for 5 min in an ultrasonic cleaning bath, and dried for 1 h at 60 °C to completely remove the water. The dishes containing the photocatalyst were placed at room temperature for cooling before being used in the NO degradation experiments. The degradation efficiency of the photocatalytic activity was evaluated by Equation (1):
η % = ( C 0 C ) C × 100
where η denotes the degradation efficiency, and C0 and C denote the initial and final concentrations of NO with respect to time.

2.4.3. Evaluation of Active Species

To identify the effective scavengers, trapping tests were conducted by introducing trapping scavengers such as K2Cr2O7, IPA, and KI to trap electrons (e−), hydroxyl radicals (OH), and holes (h+), respectively. The samples were dispersed in a glass dish (d = 12 cm) containing 20 mL of deionized water and the addition of 1%wt of the trapping agent and then sonicated for 20 min. The coated dishes were baked in an oven for 1 h at 60 °C in order to completely remove the water before the NO removal test was conducted.

3. Results and Discussion

3.1. Characterizations

A TEM and SEM study characterized the morphologies of the pure Bi2S3, g-C3N4, and Bi2S3@g-C3N4. The TEM images of the synthesized photocatalyst are shown in Figure 2a–c. Figure 2a shows the Bi2S3 nanostructure composed of nanorods, and Figure 2b shows pure g-C3N4 with an irregular shape and aggregated layers, but the surface of the flat layers was smooth. The TEM images of composites were investigated to explore the morphology and the structural combination of the photocatalysts. Figure 2c displays the TEM result of the prepared Bi2S3@g-C3N4 morphological structure. After calcination of the Bi2S3@g-C3N4, the Bi2S3 nanorods appeared to be well distributed and narrow, which appeared darker than the g-C3N4 because of its weighted atoms, which were well scattered on the g-C3N4 with a junction and appeared as a rough surface morphology. Further, a thin layer of the flexible g-C3N4 nanosheets covered the Bi2S3 nanorods, which gave rise to a core-shell structure formed between the two composites. As well as wrapping and covering Bi2S3 nanorods, the g-C3N4 nanosheets filled the space between them, thus greatly improving the linkage between composites. This unique structure made the Bi2S3 nanorods more efficient in electron transportation. The resulting images of the Bi2S3@g-C3N4 showed the internal structure of the photocatalyst and the presence of both composites, g-C3N4 and Bi2S3.
The SEM images identified the Bi2S3 nanorods over the g-C3N4, as shown in Figure 2d–f. As shown in Figure 2d, the surface of Bi2S3 revealed the number of nanorods on its surface with apparent aggregation. As shown in Figure 2e, the g-C3N4 was established by several typical folded nanosheets, which could be easily analyzed by its edge. Figure 2f shows an SEM image of the Bi2S3@g-C3N4, where evidence of holes and folded sheets were visible on its surface. The nanorods of the Bi2S3 dispersed on the g-C3N4 folded sheets consistently, which provided a greater surface area, enhanced the porous properties, and provided a greater surface permeability. As a result of these properties, the composite Bi2S3@g-C3N4 sample provided an effectively higher capacity for storing charge. Dramatically, the nanorods of Bi2S3 became narrower than before due to the strong interconnection. These results supported the successful synthesis of the photocatalyst.
The XRD patterns of the synthesized composite before the photocatalytic test are shown in Figure 3a. The diffraction patterns of the composite g-C3N4 at 2θ = 13° and 27° were well-indexed to the (100) and (002) planes for graphitic materials, respectively. It is common to find two diffraction peaks in all sulfur-based g-C3N4 products prepared by addition of thiourea [33]. The composite g-C3N4 showed the two diffraction peaks at 2θ = 13° and 27° were well-indexed to the (1 0 0) and (0 0 2) for graphitic planes, respectively. The observed peak position at 2θ = 13° was associated with the tri-s-triazine in-plane structural unit, and the other observed strong peak at 2θ = 27° was ascribed to the interplanar stacking of the conjugated aromatic system [34], a cyclic phase with the bonds that enhanced the stability and capability as compared to other arrangements. The g-C3N4 and Bi2S3 diffraction peaks were noticed in all the Bi2S3@g-C3N4 composites, as shown in Figure 3a. There were specific peaks of Bi2S3@g-C3N4 observed on the photocatalyst surface; diffraction peaks were observed at 23.1°, 27°, 12.5°, and 33.2°, corresponding to the (013), (143), (121), and (122) planes of the Bi2S3@g-C3N4 respectively. Both composites displayed strong and sharp diffraction peaks, which indicated the advantageous crystallinity. By adding Bi2S3, the intensity of the g-C3N4 diffraction peaks decreased, while the intensity of the Bi2S3 increased, demonstrating that the Bi2S3 was successfully fabricated in the heterojunction with g-C3N4. Figure 3b shows the XRD pattern of the Bi2S3@g-C3N4 composite after the cycle of photocatalytic reactions, where no other new diffraction peaks were produced. This demonstrated that the structure of the Bi2S3@g-C3N4 composite did not change the crystallinity. These results indicated that the Bi2S3@g-C3N4 composite had good reusability and stability. Further, there was no evidence of impurities, which proved the crystallinity and purity of the photocatalysts.
The Fourier Transform Infrared (FTIR) spectra of the synthesized composites are illustrated in Figure 3c; their surface bonding and chemical groups were examined. The spectrum of pure g-C3N4 showed absorptions bands between 1150 and 1750 cm−1 and 3250 and 3600 cm−1. The absorption band of pure g-C3N4 at 1620 cm−1 and 3400 cm−1 represented the bending and stretching vibrations of the adsorbed C-N bonds’ and O-H bonds’ residuals, respectively, while the sharp band between 750 and 800 cm−1 represented the characteristics of the triazine units. The Bi2S3@g-C3N4 and g-C3N4 did not exhibit any additional peaks in the FTIR spectrum, which indicated a purely physical interface existed between the g-C3N4 and Bi2S3.
The elemental mapping technique was used in order to assess the presence of the elements in the Bi2S3@g-C3N4. As shown in Figure 4a, by analyzing the EDS spectrum of the photocatalyst Bi2S3@g-C3N4, the chemical composition of the prepared nanoparticles was identified, and the peaks of the catalysts Bi2S3 and g-C3N4 were examined. The construction of the achieved nanoparticles consisted of Bi, S, C, and N elements. Bi was the dominant element in this catalyst, and elements C and N were also settled in the catalyst. According to the EDS analysis of the synthesized Bi2S3@g-C3N4, the ratios of the N, C, S, and Bi for this compound were almost 1:6:13:80. The EDS analysis of the solid-state synthesized Bi2S3@g-C3N4 showed that the Bi, S, C, and N ratios for the Bi2S3@g-C3N4 were almost 80:13:6:1, respectively. The loading of the Bi in the Bi2S3@g-C3N4 was the highest, while the loading of the N and C in the Bi2S3@g-C3N4 was the lowest, which affirmed that the synthesized composite was fabricated by both Bi2S3 and g-C3N4, which displayed a notable variation between the element content and the input material ratio in the sample.
Furthermore, Figure 4b–e displays the valence state and the chemical compositions of the synthesized material, which were inspected by X-ray photoelectron spectroscopy (XPS). Figure 4b shows the survey and high resolution XPS spectra of the observed elemental signals of the Bi, S, C, and N in the binary system. The elements in the composite Bi2S3@g-C3N4 elaborated into Gaussian–Lorentzian peaks. Figure 4c shows the N 1s spectrum of the Bi2S3@g-C3N4 presented two obvious peaks at 396.72 eV and 398.34 eV, which were attributed to the sp2 hybridized N atoms involved in the form of C–N–C and the tertiary bridging nitrogen atoms in (N–(C)3) due to the junction between the g-C3N4 and Bi2S3. Figure 4d shows the C 1s spectrum, which presented two peaks found at 282.4 eV and 286.8 eV associated with the sp2 hybridized of C―C and the sp2 hybridized of (N–C=N) bonds in the carbon graphitic structure; in Figure 4e, the high resolution XPS spectrum of Bi 4f displayed two distinct peaks approximately at 157.15 eV and 162.47 eV associated with Bi 4f7/2 and Bi 4f5/2, respectively [35], which showed the evidence of the Bi2S3 on the g-C3N4 nanosheets. So, a strong electronic operation was found between the Bi2S3 and g-C3N4; the strong interconnection at the interface caused the electrons to move from the Bi2S3 to the g-C3N4.

3.2. Photocatalytic Test

As shown in Figure 5a, the NO was photocatalytically degraded by the Bi2S3@g-C3N4, g-C3N4, and Bi2S3 for 30 min under solar irradiation. Due to their large surface areas and low bandgap energies, nanocomposites can display photocatalytic activity. The initial concentration of NO dropped rapidly within 5 min of the photocatalytic activity and eventually reached the minimum concentration over the rest of experiment. The degradation rate of the NO over the photocatalyst Bi2S3@g-C3N4 was 55%, which was significantly higher than the Bi2S3 (35%) and g-C3N4 (45%), respectively. This was because doping on the g-C3N4 with Bi2S3 improved the separation rate between the electrons and holes due to its high absorption capacity. The apparent quantum efficiency (AQE), shown in Figure 5b, is the measurement to compare the utilized photogenerated electrons when the solar light was applied during the photocatalytic degradation of the NO over Bi2S3@g-C3N4. However, the degradation efficiency was more than 30% for all the composites, and the AQE was less than 10 × 10−4 %, which meant a large number of photons were not operated in this photocatalytic reaction. Figure 5c shows the stability of the photocatalyst Bi2S3@g-C3N4 under solar repeated irradiation, studied for its functional operation. A practicable composite sustained the performance in such a way that the photocatalyst could be used several times. To examine the stability, five repetitions of consecutive photocatalytic activities were conducted for the degradation of NO; after the completion of five successive cycles, the photocatalytic degradation of the nitric oxide remained same, and it showed the stability and ability of the catalyst to degrade the nitric oxide under solar light. Moreover, Figure 5d shows the trapping test conducted under solar light to investigate the mechanism of the photocatalytic activity. The photocatalytic degradation of the NO on the materialized catalyst was primarily driven by electrons (e−), hydroxyl radicals (OH), and hole pairs (h+) [36]. The scavengers (KI), (IPA), and (Kr2Cr2O7) were used in this study to trap holes (h+), hydroxyl radicals (OH), and electrons (e−), respectively, in order to examine the active radical species. Furthermore, the electrons (e−) were inhibited significantly when the scavengers were added by 1 wt% to the photocatalyst; these results revealed that photoexcited electrons were the most significant active species undergoing NO degradation.
These results showed that the photogenerated electrons were the primary effective species associated with the photocatalytic reaction. These reactions revealed that the photocatalytic degradation of the NO over the Bi2S3@g-C3N4 was primarily driven by the photogenerated electrons and holes.

3.3. Photocatalytic Mechanism of NO Degradation

Figure 6 shows the optical analysis of the prepared composites. The band gap of the synthesized photocatalyst is shown in Figure 6a, which defined the electric conductivity of the photocatalysts. The Tauc method was introduced to determine the band gap of the photocatalyst. Based on the relationship between the Eg and the optical absorption, the band gap energies of the Bi2S3, g-C3N4, and Bi2S3@g-C3N4 were evaluated. A wider band gap is reduced through the doping of metal oxides. The band gap of the Bi2S3@g-C3N4, observed as 2.76 eV, was the gap between the bonding molecular orbitals and the antibonding molecular orbitals. The doping process, by adding Bi2S3 and g-C3N4, altered the Fermi level, replaced the CB in g-C3N4, and replaced the VB in Bi2S3, which narrowed the distance between the CB and VB. In addition, these dopant atom orbitals acted as intermediate states or bridged the gap between the valence and conduction band of electrons of the Bi2S3 and g-C3N4, some new bonds were formed, and the number of molecular orbitals increased resulting in shorter energy for electronic transitions (a lower bandgap). Furthermore, a smaller particle size led to enhanced electron transfer properties, preventing electron–hole recombination and reducing the band gap of the composite, thereby increasing the photocatalytic activity. When the band gap was reduced, excitation occurred at lower irradiation powers. The determined energy band gap was estimated as 2.86 eV and 3.1 eV for the g-C3N4 and Bi2S3, respectively, while for the Bi2S3@g-C3N4, it was estimated as 2.76 eV. Figure 6a shows the observed band gaps as compared to the Bi2S3 and g-C3N4; the Bi2S3@g-C3N4 showed an improvement in light absorption, electron level structure, surface area, particle size, and morphology. The Eg of the Bi2S3@g-C3N4 corresponded less to the Bi2S3 and g-C3N4, which showed the potential for photocatalytic activity. The nanojunction of the Bi2S3 and g-C3N4 was advantageous for utilizing and absorbing the solar light, which was the cause of the formation of more electrons and performed as a good photocatalyst. For the photocatalytic mechanism of the prepared heterojunction composites shaped by the junction of the Bi2S3 and g-C3N4 shown in Figure 6b, both composites contributed a supportive platform to carry the photogenerated charge carriers from the valence band (VB) to the conduction band (CB) of the Bi2S3 and g-C3N4. When the Bi2S3 and g-C3N4 became excited under solar light, a photoinduced electron–hole pair was generated. In the Bi2S3, the electrons at the CB recombined rapidly with the holes at the VB of the g-C3N4 due to the synergistic effects of an inner electric field, Coulomb interaction, and band bending; during this time, the functional electrons and holes remained attached to the CB and VB of the g-C3N4 and the Bi2S3, respectively, for participation in the redox reaction. Based on the result, the charge carriers were efficiently separated and utilized, and the heterojunction photocatalyst boosted the photocatalytic activity, which was notably improved.

4. Conclusions

The present study provided an effective method for preparing a photocatalyst for the degradation of NO under solar light based on Bi2S3@g-C3N4. The degradation efficiency of the NO by the Bi2S3@g-C3N4 achieved 55%. The photocatalytic activity of the Bi2S3@g-C3N4 was attributed to the synergistic effects that occurred between the two heterojunction composites of g-C3N4 and Bi2S3, which expanded the absorption of solar light and accelerated the electron–hole pair separation efficiency. In the future, this photocatalyst could be tested on other pollutants, such as volatile organic compounds, under solar or visible light or coated on the surface of tiles and bricks and used in indoor and outdoor environments for the degradation of atmospheric pollutants.

Author Contributions

Writing—original draft, Investigation, Synthesis of material, and Experimental work, A.H. (Adnan Hussain); Supervision and Validation, C.L.; Writing—Reviewing and Editing, N.K.C.; Visualization, W.-Y.H.; Data curation, K.-S.L.; Software, A.H. (Abrar Hussain). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shang, H.; Li, M.; Li, H.; Huang, S.; Mao, C.; Ai, Z.; Zhang, L. Oxygen vacancies promoted the selective photocatalytic removal of NO with blue TiO2 via simultaneous molecular oxygen activation and photogenerated hole annihilation. Environ. Sci. Technol. 2019, 53, 6444–6453. [Google Scholar] [CrossRef] [PubMed]
  2. Dvořák, R.; Chlápek, P.; Jecha, D.; Puchýř, R.; Stehlík, P. New approach to common removal of dioxins and NOx as a contribution to environmental protection. J. Clean. Prod. 2010, 18, 881–888. [Google Scholar] [CrossRef]
  3. Houshfar, E.; Lovas, T.; Skreiberg, O. Detailed chemical kinetics modeling of NOx reduction in combined staged fuel and staged air combustion of biomass. In Proceedings of the 18th European Biomass Conference and Exhibition, Lyon, France, 3–7 May 2010. [Google Scholar]
  4. Li, H.; ElKady, A.; Evulet, A. Effect of exhaust gas recirculation on NOx formation in premixed combustion system. In Proceedings of the 47th AIAA Aerospace Sciences Meeting Including the New Horizons Forum and Aerospace Exposition, Orlando, FL, USA, 5–8 January 2009. [Google Scholar]
  5. Charles, E.B. Industrial Combustion Pollution and Control; CRC Press: Boca Raton, FL, USA, 2003. [Google Scholar]
  6. Nikokavoura, A.; Trapalis, C. Graphene and g-C3N4 based photocatalysts for NOx removal: A review. Appl. Surf. Sci. 2018, 430, 18–52. [Google Scholar] [CrossRef]
  7. Dalton, J.S.; Janes, P.A.; Jones, N.; Nicholson, J.A.; Hallam, K.R.; Allen, G.C. Photocatalytic oxidation of NOx gases using TiO2: A surface spectroscopic approach. Environ. Pollut. 2002, 120, 415–422. [Google Scholar] [CrossRef]
  8. Devahasdin, S.; Fan, C., Jr.; Li, K.; Chen, D.H. TiO2 photocatalytic oxidation of nitric oxide: Transient behavior and reaction kinetics. J. Photochem. Photobiol. A 2003, 156, 161–170. [Google Scholar] [CrossRef]
  9. Poon, C.S.; Cheung, E. NO removal efficiency of photocatalytic paving blocks prepared with recycled materials. Constr. Build. Mater. 2007, 21, 1746–1753. [Google Scholar] [CrossRef]
  10. Ballari, M.; Hunger, M.; Hüsken, G.; Brouwers, H. NOx photocatalytic degradation employing concrete pavement containing titanium dioxide. Appl. Catal. B 2010, 95, 245–254. [Google Scholar] [CrossRef]
  11. Huang, Y.; Cao, J.-J.; Kang, F.; You, S.-J.; Chang, C.-W.; Wang, Y.-F. High selectivity of visible-light-driven La-doped TiO2 photocatalysts for NO removal. Aerosol Air Qual Res. 2017, 17, 2555–2565. [Google Scholar] [CrossRef] [Green Version]
  12. Li, Y.; Cui, W.; Liu, L.; Zong, R.; Yao, W.; Liang, Y.; Zhu, Y. Removal of Cr (VI) by 3D TiO2-graphene hydrogel via adsorption enriched with photocatalytic reduction. Appl. Catal. B 2016, 199, 412–423. [Google Scholar] [CrossRef]
  13. Wang, Y.; Cheng, H.; Zhang, L.; Hao, Y.; Ma, J.; Xu, B.; Li, W. The preparation, characterization, photoelectrochemical and photocatalytic properties of lanthanide metal-ion-doped TiO2 nanoparticles. J. Mol. Catal. A Chem. 2000, 151, 205–216. [Google Scholar] [CrossRef]
  14. Ranjit, K.; Willner, I.; Bossmann, S.; Braun, A. Lanthanide oxide-doped titanium dioxide photocatalysts: Novel photocatalysts for the enhanced degradation of p-chlorophenoxyacetic acid. Environ. Sci. Technol. 2001, 35, 1544–1549. [Google Scholar] [CrossRef] [PubMed]
  15. He, Y.; Zhang, L.; Teng, B.; Fan, M. New application of Z-scheme Ag3PO4/g-C3N4 composite in converting CO2 to fuel. Environ. Sci. Technol. 2015, 49, 649–656. [Google Scholar] [CrossRef] [PubMed]
  16. Mamba, G.; Mishra, A. Graphitic carbon nitride (g-C3N4) nanocomposites: A new and exciting generation of visible light driven photocatalysts for environmental pollution remediation. Appl. Catal. B 2016, 198, 347–377. [Google Scholar] [CrossRef]
  17. Wang, X.; Chen, X.; Thomas, A.; Fu, X.; Antonietti, M. Metal-containing carbon nitride compounds: A new functional organic–metal hybrid material. Adv. Mater. 2009, 21, 1609–1612. [Google Scholar] [CrossRef]
  18. Cao, S.; Low, J.; Yu, J.; Jaroniec, M. Polymeric photocatalysts based on graphitic carbon nitride. Adv. Mater. 2015, 27, 2150–2176. [Google Scholar] [CrossRef]
  19. Faisal, M.; Jalalah, M.; Harraz, F.A.; El-Toni, A.M.; Khan, A.; Al-Assiri, M. Au nanoparticles-doped g-C3N4 nanocomposites for enhanced photocatalytic performance under visible light illumination. Ceram. Int. 2020, 46, 22090–22101. [Google Scholar] [CrossRef]
  20. Li, Z.; Kong, C.; Lu, G. Visible photocatalytic water splitting and photocatalytic two-electron oxygen formation over Cu-and Fe-doped g-C3N4. J. Phys. Chem. C 2016, 120, 56–63. [Google Scholar] [CrossRef]
  21. Zhao, R.; Sun, X.; Jin, Y.; Han, J.; Wang, L.; Liu, F. Au/Pd/g-C3N4 nanocomposites for photocatalytic degradation of tetracycline hydrochloride. J. Mater. Sci. 2019, 54, 5445–5456. [Google Scholar] [CrossRef]
  22. Liu, Z.; Zhang, M.; Wu, J. Enhanced Visible-Light Photocatalytic and Antibacterial Activities of Ag-Doped g-C3N4 Nanocomposites. ChemistrySelect 2018, 3, 10630–10636. [Google Scholar]
  23. Chen, W.; Liu, T.; Huang, T.; Liu, X.; Yang, X. Novel mesoporous P-doped graphitic carbon nitride nanosheets coupled with ZnIn2S4 heterostructures with remarkably enhanced. Nanoscale 2016, 8, 3711–3719. [Google Scholar] [CrossRef]
  24. Han, Q.; Hu, C.; Zhao, F.; Zhang, Z.; Chen, N.; Qu, L. One-step preparation of iodine-doped graphitic carbon nitride nanosheets as efficient photocatalysts for visible light water splitting. J. Mater. Chem. 2015, 3, 4612–4619. [Google Scholar] [CrossRef]
  25. Sagara, N.; Kamimura, S.; Tsubota, T.; Ohno, T. Photoelectrochemical CO2 reduction by a p-type boron-doped g-C3N4 electrode under visible light. Appl. Catal. B 2016, 192, 193–198. [Google Scholar] [CrossRef] [Green Version]
  26. Xu, C.; Han, Q.; Zhao, Y.; Wang, L.; Li, Y.; Qu, L. Sulfur-doped graphitic carbon nitride decorated with graphene quantum dots for an efficient metal-free electrocatalyst. J. Mater. Chem. 2015, 3, 1841–1846. [Google Scholar] [CrossRef]
  27. Zhu, X.; Wang, Y.; Guo, Y.; Wan, J.; Yan, Y.; Zhou, Y.; Sun, C. Environmental-friendly synthesis of heterojunction photocatalysts g-C3N4/BiPO4 with enhanced photocatalytic performance. Appl. Surf. Sci. 2021, 544, 148–872. [Google Scholar] [CrossRef]
  28. Zhang, S.; Yang, Y.; Guo, Y.; Guo, W.; Wang, M.; Guo, Y.; Huo, M. Preparation and enhanced visible-light photocatalytic activity of graphitic carbon nitride/bismuth niobate heterojunctions. J. Hazard. Mater. 2013, 261, 235–245. [Google Scholar] [CrossRef]
  29. Luo, S.; Qin, F.; Zhao, H.; Liu, Y.; Chen, R. Fabrication uniform hollow Bi2S3 nanospheres via Kirkendall effect for photocatalytic reduction of Cr (VI) in electroplating industry wastewater. J. Hazard. Mater. 2017, 340, 253–262. [Google Scholar] [CrossRef]
  30. Shi, H.; Zhao, Y.; Fan, J.; Tang, Z. Construction of novel Z-scheme flower-like Bi2S3/SnIn4S8 heterojunctions with enhanced visible light photodegradation and bactericidal activity. Appl. Surf. Sci. 2019, 465, 212–222. [Google Scholar] [CrossRef]
  31. Qiao, X.-Q.; Zhang, Z.-W.; Li, Q.-H.; Hou, D.; Zhang, Q.; Zhang, J.; Bu, X. In situ synthesis of n–n Bi2MoO6 Bi2S3 heterojunctions for highly efficient photocatalytic removal of Cr (VI). J. Mater. Chem. 2018, 6, 22580–22589. [Google Scholar] [CrossRef]
  32. Hu, T.; Dai, K.; Zhang, J.; Zhu, G.; Liang, C. One-pot synthesis of step-scheme Bi2S3/porous g-C3N4 heterostructure for enhanced photocatalytic performance. Mater. Lett. 2019, 257, 126740. [Google Scholar] [CrossRef]
  33. An, T.D.; Phuc, N.V.; Tri, N.N.; Phu, H.T.; Hung, N.P.; Vo, V. Sulfur-doped g-C3N4 with enhanced visible-light photocatalytic activity. Appl. Mech. Mater. 2019, 889, 43–50. [Google Scholar] [CrossRef]
  34. Yu, W.; Chen, J.; Shang, T.; Chen, L.; Gu, L.; Peng, T. Direct Z-scheme g-C3N4/WO3 photocatalyst with atomically defined junction for H2 production. Appl. Catal. B 2017, 219, 693–704. [Google Scholar] [CrossRef]
  35. Moyseowicz, A. Scalable one-pot synthesis of bismuth sulfide nanorods as an electrode active material for energy storage applications. J. Solid State Electrochem. 2019, 23, 1191–1199. [Google Scholar] [CrossRef]
  36. Nguyen, M.T.; Tran, H.H.; You, S.-J.; Wang, Y.-F.; Van Viet, P. Green synthesis of Ag@SnO2 nanocomposites for enhancing photocatalysis of nitrogen monoxide removal under solar light irradiation. Catal. Commun. 2020, 136, 105902. [Google Scholar]
Figure 1. Illustration of the experimental layout in this study.
Figure 1. Illustration of the experimental layout in this study.
Nanomaterials 12 03482 g001
Figure 2. The TEM images of the synthesized (a) Bi2S3, (b) g-C3N4, and (c) Bi2S3@g-C3N4. The SEM images of the synthesized (d) Bi2S3, (e) g-C3N4, and (f) Bi2S3@g-C3N4.
Figure 2. The TEM images of the synthesized (a) Bi2S3, (b) g-C3N4, and (c) Bi2S3@g-C3N4. The SEM images of the synthesized (d) Bi2S3, (e) g-C3N4, and (f) Bi2S3@g-C3N4.
Nanomaterials 12 03482 g002
Figure 3. XRD and FTIR analysis results. (a) The XRD patterns of the synthesized Bi2S3@g-C3N4, Bi2S3, and g-C3N4 before the photocatalytic test. (b) The XRD analysis of the synthesized Bi2S3@g-C3N4 after the photocatalytic test and the (c) FTIR spectra analysis of the prepared Bi2S3@g-C3N4, Bi2S3, and g-C3N4.
Figure 3. XRD and FTIR analysis results. (a) The XRD patterns of the synthesized Bi2S3@g-C3N4, Bi2S3, and g-C3N4 before the photocatalytic test. (b) The XRD analysis of the synthesized Bi2S3@g-C3N4 after the photocatalytic test and the (c) FTIR spectra analysis of the prepared Bi2S3@g-C3N4, Bi2S3, and g-C3N4.
Nanomaterials 12 03482 g003
Figure 4. (a) The EDX-mapping of the synthesized Bi2S3@g-C3N4, (b) the XPS Survey, and the XPS high resolution spectra, (c) the N 1s spectrum, (d) the C 1s spectrum, and (e) the Bi 4f spectrum.
Figure 4. (a) The EDX-mapping of the synthesized Bi2S3@g-C3N4, (b) the XPS Survey, and the XPS high resolution spectra, (c) the N 1s spectrum, (d) the C 1s spectrum, and (e) the Bi 4f spectrum.
Nanomaterials 12 03482 g004
Figure 5. (a) The photocatalytic degradation, (b) apparent quantum efficiency, (c) recycling test, and (d) trapping test of the synthesized material under solar irradiation.
Figure 5. (a) The photocatalytic degradation, (b) apparent quantum efficiency, (c) recycling test, and (d) trapping test of the synthesized material under solar irradiation.
Nanomaterials 12 03482 g005
Figure 6. (a) Tauc plots of the synthesized materials and (b) the photocatalytic proposed mechanism of the Bi2S3@g-C3N4 under solar irradiation.
Figure 6. (a) Tauc plots of the synthesized materials and (b) the photocatalytic proposed mechanism of the Bi2S3@g-C3N4 under solar irradiation.
Nanomaterials 12 03482 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hussain, A.; Lin, C.; Cheruiyot, N.K.; Huang, W.-Y.; Lin, K.-S.; Hussain, A. Bismuth Sulfide Doped in Graphitic Carbon Nitride Degrades Nitric Oxide under Solar Irradiation. Nanomaterials 2022, 12, 3482. https://doi.org/10.3390/nano12193482

AMA Style

Hussain A, Lin C, Cheruiyot NK, Huang W-Y, Lin K-S, Hussain A. Bismuth Sulfide Doped in Graphitic Carbon Nitride Degrades Nitric Oxide under Solar Irradiation. Nanomaterials. 2022; 12(19):3482. https://doi.org/10.3390/nano12193482

Chicago/Turabian Style

Hussain, Adnan, Chitsan Lin, Nicholas Kiprotich Cheruiyot, Wen-Yen Huang, Kuen-Song Lin, and Abrar Hussain. 2022. "Bismuth Sulfide Doped in Graphitic Carbon Nitride Degrades Nitric Oxide under Solar Irradiation" Nanomaterials 12, no. 19: 3482. https://doi.org/10.3390/nano12193482

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