Next Article in Journal
Influence of Aging Treatment Regimes on Microstructure and Mechanical Properties of Selective Laser Melted 17-4 PH Steel
Previous Article in Journal
Research on a Method to Improve the Temperature Performance of an All-Silicon Accelerometer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Photodegradation of Rhodamine B by Fiber-like Nitrogen-Doped TiO2/Ni(OH)2 Nanocomposite under Visible Light Irradiation

1
Key Laboratory for Comprehensive Energy Saving of Cold Regions Architecture of Ministry of Education, Jilin Jianzhu University, Changchun 130118, China
2
Department of Materials Science, Jilin Jianzhu University, Changchun 130118, China
3
Department of Chemistry, Jilin Normal University, Siping 136000, China
*
Authors to whom correspondence should be addressed.
Micromachines 2023, 14(4), 870; https://doi.org/10.3390/mi14040870
Submission received: 23 March 2023 / Revised: 12 April 2023 / Accepted: 14 April 2023 / Published: 18 April 2023

Abstract

:
N-TiO2/Ni(OH)2 nanofiber was successfully prepared by combining the electrospinning and solvothermal method. It has been found that under visible light irradiation, the as-obtained nanofiber exhibits excellent activity for the photodegradation of rhodamine B, and the average degradation rate reaches 3.1%/min−1. Further insight investigations reveal that such a high activity was mainly due to the heterostructure-induced increase in the charge transfer rate and separation efficiency.

1. Introduction

Back in 2001, Asahi et al. [1,2] identified that the tactic of non-metal-doping can narrow the bandgap of TiO2 (anatase, 3.2 eV). Both experimental results and the relevant density functional theory (DFT) theoretical calculations revealed that nitrogen doping could significantly extend the absorption range of TiO2 into the visible region. Since then, numerous investigations focus on non-metal-doped (N, S, P, B) TiO2 photocatalysts [3,4]. Through the non-metallic dopants and their combinations, the corresponding photocatalytic activity of derived TiO2 could be improved, thus exhibiting outstanding photocatalytic performances in various applications, such as photocatalytic water splitting [5,6], CO2 reduction [7,8], removal of organic compounds [9,10], sterilization [11], and anti-viral applications [12]. Owing to these excellent properties, N-doped TiO2 and cocatalyst-supported TiO2 have been commercialized and widely used in many practical applications [13]. However, the photocatalytic reaction rate of N-doped TiO2 is still very low, which is believed to be caused by the following two reasons: one is its weak absorption in the visible range, while the other is the limited rate of the photocatalytic reactions by the nitrogen-induced oxygen vacancies, which is supposed to act as the carrier traps and/or recombination centers.
To enhance the photocatalytic performance of nitrogen-doped TiO2 (N-TiO2) in the visible range, tremendous efforts have been devoted to it during the past decades [14,15,16]. It has been revealed that transition metal nickel hydroxide and NiS could act as cocatalysts to significantly strengthen the photocatalytic activity of TiO2, which is believed to be attributed to the inhibition of charge carriers’ recombination [17,18,19]. Zhang et al. [20] prepared three-dimensional flower-like TiO2@Ni(OH)2 core–shell heterostructures, exhibiting a six-fold activity enhancement for hydrogen photogeneration. Meng et al. [21] synthesized a hierarchical TiO2/Ni(OH)2 composite photocatalyst via decorating Ni(OH)2 nanosheets on the electrospun TiO2 nanofibers by a simple wet method, and the results indicate that, compared to pristine TiO2 fibers, its activity of CO2 photoreduction was significantly improved, which was ascribed to the increased charge separation efficiency and higher CO2 capture capacity owing to the presence of Ni(OH)2. Chen et al. [22] reported the construction of a photo-supercapacitor with high specific capacitance and ultra-fast charge–discharge response, in which TiO2 was used as a photoelectric conversion material and Ni(OH)2 was used as an energy storage material in the TiO2/Ni(OH)2 composite material, offering a simple fabrication process for efficient direct solar energy storage. These findings inspired us to develop a heterojunction between N-TiO2 and nickel hydroxide to increase the charge separation efficiency thereby improving its corresponding photocatalytic activity.
To our knowledge, only a few studies were focused on the photocatalytic applications of N-TiO2/Ni(OH)2 composite. In this work, we first prepared N-TiO2 nanofibers by electrospinning technique followed by decoration of Ni(OH)2 via the hydrothermal method to obtain the high-quality N-TiO2/Ni(OH)2 heterostructures. The designated composite exhibits a much-enhanced photocatalytic ability for the degradation of rhodamine B (RhB) compared to N-TiO2 under visible light irradiation. Further detailed analyses reveal the relevant enhanced activity is ascribed to the increased charge separation efficiency induced by the delicate heterostructure.

2. Experimental Section

2.1. Chemicals and Reagents

Acetic acid (99.5%), butyl titanate (99.0%), ethanol (99.0%), urea (99.0%), polyvinylpyrrolidone (99.0%), Rhodamine B (99.0%), nickel nitrate hexahydrate (99.0%), and hexamethylenetetramine (99.0%) were purchased from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). All chemicals were used as received without further purification.

2.2. Synthesis of Ni(OH)2 Nanosheets

N-TiO2 nanofibers were prepared by the electrospinning method according to the previous work [23]. Typically, 5 mL of acetic acid and 5 mL of butyl titanate were added into 10 mL ethanol under vigorous stirring at room temperature (RT). After ~10 min., 0.2 g of urea was introduced, followed by another 20 min stirring. Meanwhile, another solution was obtained by introducing 1.5 g of polyvinylpyrrolidone into 10 mL ethanol. Then, the two above solutions were homogeneously mixed and further heated at 70 °C for 1 h under vigorous stirring to obtain the precursor solution for electrospinning. At a static voltage of 15 kV, a capillary tube with an inner diameter of 0.5 mm at a 15 cm distance from the collector was used to prepare the electrospun film. The flow rate was set at 1 mL/h. Afterward, the fiber film was peeled off and placed in a crucible that was sealed with tin foil and kept at 550 °C for 3 h in a muffle furnace to obtain N-TiO2 fibers. For comparison, TiO2 nanofibers were prepared by a similar procedure and the only exception is without introducing urea.

2.3. Preparation of N-TiO2/Ni(OH)2 Nanocomposite

A total of 100 mg of the as-prepared N-TiO2 nanofibers were added into 40 mL ultrapure water followed by sonication for 10 min to obtain a homogeneous suspension. Next, 3.1 mg nickel nitrate hexahydrate and 1.7 mg hexamethylenetetramine were introduced and the corresponding solution was further stirred for 1 h at RT. Then, the solution was transferred to the Teflon-lined autoclave and heated at 120 °C for another 6 h. The obtained precipitation was alternatively washed with ethanol and water several times and subsequently dried in a vacuum oven at 80 °C for 4 h. Samples with 0.5 and 1.5 wt% Ni(OH)2 contents were prepared by the same protocol with different amounts of nickel nitrate hexahydrate and hexamethylenetetramine introduced.

2.4. Characterization

The crystal structure of samples was investigated on an X-ray diffractometer (Rigaku D/Max 2550, Japan Rigaku Co., Ltd., Tokyo, Japan) at 40 kV and 40 mA with copper Kα radiation (λ = 0.154056 nm). The morphological and structural information was characterized via field-emission scanning electron microscopy (SEM, JSM-7610F, Tokyo, Japan) and transmission electron microscopy (TEM, Tecnai G2 F20 S-TWIN, FEI, Valley City, ND, USA). X-ray photoelectron spectroscopy (XPS) measurements were recorded by a photoelectron spectrometer (ESCALAB MKII, VG Scientific, Waltham, MA, USA). The relevant photoluminescence results were obtained on an FLS-1000 fluorescence spectrophotometer (Edinburgh Instruments Ltd., Livingston, UK). The transient photocurrent signals were measured by an electrochemical workstation (CHI-660, CH Instruments, Austin, TX, USA). UV-Vis diffuse reflectance spectra of the samples, and the absorbance of Rh B solution was measured on a Shimadzu-2600 spectrophotometer (Shimadzu, Kyoto, Japan).

2.5. Measurement of Photocatalytic Reactions

The RhB photodegradation experiments were performed in a glass flask. Typically, 25 mg of the relevant photocatalyst was dispersed into a 100 mL Rh B aqueous solution (10 mg/L). Before the photoreaction, the solution was stirred for 30 min in dark to exclude the adsorption effect of the sample. Then, the solution was illuminated by a 300-W Xenon lamp with a filter (>420 nm). A total of 2 mL of the solution was taken at regular intervals (30 min), then subjected to UV-Vis measurements after filtrating the photocatalyst.

3. Results

The crystal structural information of TiO2, N-TiO2, and N-TiO2/Ni(OH)2 nanofibers were obtained by X-ray diffraction (XRD) technique. As shown (Figure 1), for TiO2, the clear diffraction peaks at 2θ = 25.2°, 37.8°, 47.8°, 55.2°, 62.7°, 68.9°, 75.1°, and 82.6° ascribed to (101), (004), (200), (105), (204), (116), and (215) crystal planes of anatase (JCPDS card number 21-1272) were observed, respectively. Meanwhile, peaks at 2θ = 27.4°, 36.1°, and 41.1°, which is consistent with the (110), (101), and (111) crystal planes of rutile, respectively, TiO2 (JCPDS card number 21-1276) are also detected. The results demonstrate that under the current calcination condition, the obtained TiO2 nanofibers are in the mixed phase of anatase and rutile. Interestingly, when nitrogen is introduced into TiO2, the rutile phase of TiO2 seems to be significantly inhibited. Furthermore, for N-TiO2/Ni(OH)2 nanocomposite, no peak belonging to Ni(OH)2 was observed, which is considered to be owing to the low Ni(OH)2 content [18].
The morphological and structural information of the sample was obtained by SEM and TEM measurements. Figure 2a displays the SEM result of TiO2 nanofibers. It is seen that the TiO2 was in a fibrous structure with a diameter of ~500 nm and length over 10 µm, while the introduction of nitrogen did not induce any observable morphological change (Figure 2b). Further decoration of Ni(OH)2 onto N-TiO2 nanofibers could induce the two-dimensional structural growth on the surface of the corresponding samples, (Figure 2c), which manifests the successful deposition of Ni(OH)2. TEM image of the N-TiO2/1.0 wt% Ni(OH)2 nanofibers indicates the diameter of the relevant sample in ~500 nm (Figure 2d), and higher magnification TEM result clearly demonstrates a 50-nm-thick layer was homogenously coated onto the outside surface of the N-TiO2 fibers, which is considered to be Ni(OH)2 (Figure 2e). Furthermore, HRTEM measurement confirms our assumption. As is shown (Figure 2f), the clear granular nanostructure ascribed to (110) crystal plane of Ni(OH)2 was found in the corresponding outside layer area of the sample, and the green circle represents Ni(OH)2 nanocrystalline particles in Figure 2f.
To acquire the chemical state of each element in the samples, XPS measurements were performed. Figure 3 shows the high-resolution Ti 2p, O 1s, N 1s, and Ni 2p XPS spectra of the N-TiO2/1.0 wt% Ni(OH)2. As is shown for Ti 2p, peaks at 458.5 and 464.3 eV could be ascribed to Ti 2p3/2 and Ti 2p1/2 of TiO2, respectively, indicating the existence of Ti4+ in the sample [24,25,26]; while for O 1s, signals of lattice oxygen (titanium–oxygen–titanium, 529.7 eV) and hydroxyl oxygen (Ti-OH and Ni-OH, 531.4 eV) are observed [27] (Figure 3b). A weak characteristic peak (N1s) at 399.9 eV attributes to the signal combination of weakly charged nitrogen species with C, H, or O atoms [28] (Figure 3c), and demonstrates the successful diffusion of nitrogen into the TiO2 lattice at interstitial positions [29,30]. It is noted that nitrogen had a weak positive charge in the Ni 2p region (Figure 3d) due to the bonding with O atoms of TiO2 [30,31,32]. In addition, the peaks at 855.3 and 873.0 eV correspond to Ni 2p3/2 and Ni 2p1/2, respectively, indicating the existence of nickel hydroxide [33]. In addition, the two shoulder peaks were ascribed to the satellite peaks of Ni 2p3/2 and Ni 2p1/2 in the vicinity of 861.7 and 880.3 eV, respectively [20].
The absorption property of different samples was measured by UV-Vis diffuse reflectance absorption spectroscopy. As indicated (Figure 4), Ni(OH)2 control sample exhibits an obvious absorption at 450 and 600–800 nm, which is due to the d–d transition of Ni [34]. In addition, compared to TiO2, the absorption of N-TiO2 and different N-TiO2/Ni(OH)2 samples in the visible region are significantly enhanced. It should be noted that, in the range of 400–550 nm, the absorption intensity of N-TiO2/Ni(OH)2 nanofibers is increased with the increase in Ni(OH)2 content.
The photocatalytic activity on the degradation of RhB of different samples was then evaluated. As shown (Figure 5a), all the employed samples exhibit the similar physical adsorption property for RhB. However, interestingly, once illuminated, all the N-TiO2/Ni(OH)2 nanocomposites show a much higher ability for the photodegradation of RhB compared to that of a relevant single component. Especially, N-TiO2/Ni(OH)2 nanocomposite possesses the best performance, and ~97% of RhB would be photodegraded in 90 min. To illustrate the decomposition rate more clearly, the Langmuir-Hinshelwood kinetics equation was employed [35]. As shown in Figure 5b, the activities of all the nanocomposites were higher than those of a single component, indicating an apparent synergistic effect between N-TiO2 and Ni(OH)2, which could also be reflected by the rate constant (Figure 5c). Our results unambiguously manifest the vital role of forming the heterostructure for significantly enhancing the RhB photodegradation ability. Further optimization of the addition of Ni(OH)2 in the composite showed that when the amount of Ni(OH)2 in the composite is 1.0-wt%, the best performance of 0.0310 min−1 could be achieved (Figure 5b), which was comparable to the recent benchmark test results (Table 1).
The stability of photocatalyst is a critical issue for the potential real application. Hence, the relevant stability of N-TiO2/1.0-wt% Ni(OH)2 was evaluated (Figure 5d). After each round of photocatalytic reaction, the sample was collected via centrifugation and washed with ultrapure water several times before further usage. As shown, only a slight activity decrease (~3.7%) was observed after five cycles demonstrating its excellent stability for long-term use.
To elucidate the underlying mechanism of the heterostructure-induced photocatalytic activity enhancement, photoluminescence (PL) measurements were first employed. As shown in Figure 6a, N-TiO2 gives an obvious PL emission in the range of 400–550 nm, while for that of Ni(OH)2, hardly any PL signal could be observed. Interestingly, for the nanocomposite sample, the relevant PL intensity is gradually decreased with the gradual increase in Ni(OH)2 content, which is considered to be ascribed to charge transfer from N-TiO2 to Ni(OH)2 [35]. To further verify the separation and migration behavior of charge carriers, photocurrent response tests were performed. As shown in Figure 6b, all the composite samples give higher photocurrent responses than that of the N-TiO2 or Ni(OH)2, and N-TiO2/1.0 wt% Ni(OH)2 exhibits the highest response, which is consistent with the trend of their photocatalytic performance. This strongly signifies the introduction of Ni(OH)2 is beneficial for increasing charge separation efficiency, and thus enhances the relevant photocatalytic activity.
Based on the above results, the plausible underlying mechanism for the enhanced photocatalytic activity of N-TiO2/Ni(OH)2 composite has been proposed (Figure 7). After the absorption of the incident light, both N-TiO2 and Ni(OH)2 were excited. Thereafter, the electrons and holes are photogenerated in each component, and electrons are excited into the conduction band (CB) while the holes are left at the valence band (VB). Owing to the potential of Ni2+/Ni (0.23 eV) being slightly lower than that of anatase TiO2 (~0.26 eV), photogenerated electrons will transfer from CB of TiO2 into CB of Ni(OH)2, while holes remain in VB of TiO2. This will significantly promote charge separation and inhibit the recombination of photogenerated electrons and holes. Subsequently, electrons could react with adsorbed oxygen to generate superoxide free radicals. It is noted the formation of free radicals can not only effectively inhibit electron-hole recombination, but also destroy the bonds of RhB and therefore degrade them, which finally exhibits enhanced photocatalytic activity.

4. Conclusions

In this work, the heterostructured N-TiO2/Ni(OH)2 nanofibers were successfully constructed and further applied for the enhanced photodegradation of RhB compared to that of N-TiO2 under visible light irradiation (>420 nm). Our results indicate that the elaborated design could significantly enhance the corresponding photocatalytic activity and the one with 1.0 wt% Ni(OH)2 exhibits the best performance. Further detailed investigations revealed that the enhanced activity is mainly ascribed to the efficient charge transfer and separation in N-TiO2 induced by the integration of Ni(OH)2. It is anticipated that our tactics and results would be beneficial for the future design of high-performance photocatalysts.

Author Contributions

Conceptualization, H.W. and M.D.; methodology, H.W. and B.S.; validation, Y.C. and C.W.; formal analysis, H.W. and S.L.; investigation, H.W. and X.Y.; writing—original draft preparation, H.W. and R.D.; writing—review and editing, H.W., B.W., Y.C. and X.Y.; visualization, M.D. and B.S.; supervision, X.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the science and technology development project of Jilin province, China (No. 20230508059RC, No. 20210203098SF, and No. 20210203021SF), the financial support from Key Laboratory for Comprehensive Energy Saving of Cold Regions Architecture of Ministry of Education (No. JLZHKF022021008).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  2. Morikawa, T.; Asahi, R.; Ohwaki, T.; Aoki, K.; Taga, Y. Band-Gap Narrowing of Titanium Dioxide by Nitrogen Doping. Jpn. J. Appl. Phys. 2001, 40, L561. [Google Scholar] [CrossRef]
  3. Devi, L.G.; Kavitha, R. A review on nonmetal ion doped titania for the photocatalytic degradation of organic pollutants under UV/solar light: Role of photogenerated charge carrier dynamics in enhancing the activity. Appl. Catal. B Environ. 2013, 140–141, 559–587. [Google Scholar] [CrossRef]
  4. Yang, Y.; Zhong, H.; Tian, C. Photocatalytic mechanisms of modified titania under visible light. Res. Chem. Intermed. 2011, 37, 91–102. [Google Scholar] [CrossRef]
  5. Reddy, P.A.K.; Reddy, P.V.L.; Kim, K.-H.; Kumar, M.K.; Manvitha, C.; Shimn, J.-J. Novel Approach for the Synthesis of Nitrogen-Doped Titania with Variable Phase Composition and Enhanced Production of Hydrogen under Solar Irradiation. J. Ind. Eng. Chem. 2017, 53, 253–260. [Google Scholar] [CrossRef]
  6. Preethi, L.K.; Antony, R.P.; Mathews, T.; Loo, S.C.J.; Wong, L.H.; Dash, S.; Tyagi, A.K. Nitrogen Doped Anatase-Rutile Heterostructured Nanotubes for Enhanced Photocatalytic Hydrogen Production: Promising Structure for Sustainable Fuel Production. Int. J. Hydrog. Energ. 2016, 41, 5865–5877. [Google Scholar] [CrossRef]
  7. Akple, M.S.; Low, J.; Qin, Z.; Wageh, S.; Al-Ghamdi, A.A.; Yu, J.; Liu, S. Nitrogen-Doped TiO2 Microsheets with Enhanced Visible Light Photocatalytic Activity for CO2 Reduction. Chin. J. Catal. 2015, 36, 2127–2134. [Google Scholar] [CrossRef]
  8. Delavari, S.; Amin, N.A.S.; Ghaedi, M. Photocatalytic Conversion and Kinetic Study of CO2 and CH4 over Nitrogen-doped Titania Nanotube Arrays. J. Clean. Prod. 2016, 111, 143–154. [Google Scholar] [CrossRef]
  9. Kamaludin, R.; Othman, M.H.D.; Kadir, S.H.S.A.; Ismail, A.F.; Rahman, M.A.; Jaafar, J. Visible-Light-Driven Photocatalytic N-doped TiO2 for Degradation of Bisphenol A (BPA) and Reactive Black 5 (RB5) Dye. Water Air Soil Pollut. 2018, 229, 363. [Google Scholar] [CrossRef]
  10. Le, P.; Hieu, L.; Lam, T.-N.; Hang, N.; Truong, N.; Tuyen, L.; Phong, P.; Leu, J. Enhanced Photocatalytic Performance of Nitrogen-doped TiO2 Nanotube Arrays Using a Simple Annealing Process. Micromachines 2018, 9, 618. [Google Scholar] [CrossRef]
  11. Ananpattarachai, J.; Boonto, Y.; Kajitvichyanukul, P. Visible Light Photocatalytic Antibacterial Activity of Ni-doped and N-doped TiO2 on Staphylococcus aureus and Escherichia coli Bacteria. Environ. Sci. Pollut. Res. 2015, 23, 4111–4119. [Google Scholar] [CrossRef] [PubMed]
  12. Asahi, R.; Hiroshi, T.; Vequizo, J.-M.; Yamakata, A. Improvement of photocatalytic activity under visible-light irradiation by heterojunction of Cu ion loaded WO3 and Cu ion loaded N-TiO2. Appl. Catal. B Environ. 2019, 248, 249–254. [Google Scholar]
  13. Asahi, R.; Morikawa, T.; Irie, H.; Ohwaki, T. Nitrogen-Doped Titanium Dioxide as Visible-Light-Sensitive Photocatalyst: Designs, Developments, and Prospects. Chem. Rev. 2014, 114, 9824–9852. [Google Scholar] [CrossRef]
  14. Zhao, Y.; Qiu, X.; Burda, C. The Effects of Sintering on the Photocatalytic Activity of N-Doped TiO2 Nanoparticles. Chem. Mater. 2008, 20, 2629–2636. [Google Scholar] [CrossRef]
  15. Kumar, M.P.; Jagannathan, R.; Ravichandran, S. Photoelectrochemical System for Unassisted High-Efficiency Water-Splitting Reactions Using N-Doped TiO2 Nanotubes. Energy Fuels 2020, 34, 9030–9036. [Google Scholar] [CrossRef]
  16. Zhang, J.; Li, Y.; Li, L.; Li, W.; Yang, C. Dual Functional N-Doped TiO2-Carbon Composite Fibers for Efficient Removal of Water Pollutants. ACS Sustain. Chem. Eng. 2018, 6, 12893–12905. [Google Scholar] [CrossRef]
  17. Ong, W.L.; Ng, S.W.L.; Zhang, C.; Hong, M.; Ho, G.W. 2D hydrated layered Ni(OH)2 structure with hollow TiO2 nanocomposite directed chromogenic and catalysis capabilities. J. Mater. Chem. A 2016, 4, 13307–13315. [Google Scholar] [CrossRef]
  18. Yu, J.; Hai, Y.; Cheng, B. Enhanced Photocatalytic H2-Production Activity of TiO2 by Ni(OH)2 Cluster Modification. J. Phys. Chem. C 2011, 115, 4953–4958. [Google Scholar] [CrossRef]
  19. Luo, Y.; Dong, J.; Jiang, Z.; Zhan, X.; Li, Y.; Wang, C. β-Ni(OH)2/NiS/TiO2 3D flower-like p-n-p heterostructural photocatalysts for high-efficiency removal of soluble anionic dyes and hydrogen releasing. Opt. Mater. 2021, 114, 110951. [Google Scholar] [CrossRef]
  20. Zhang, W.; Zhang, H.; Xu, J.; Zhuang, H.; Long, J. 3D flower-like heterostructured TiO2@Ni(OH)2 microspheres for solar photocatalytic hydrogen production. Chin. J. Catal. 2019, 40, 320–325. [Google Scholar] [CrossRef]
  21. Meng, A.; Wu, S.; Cheng, B.; Yu, J.; Xu, J. Hierarchical TiO2/Ni(OH)2 composite fibers with enhanced photocatalytic CO2 reduction performance. J. Mater. Chem. A 2018, 6, 4729–4736. [Google Scholar] [CrossRef]
  22. Chen, P.; Cao, C.; Ding, C.; Yin, Z.; Qi, S.; Guo, J.; Zhang, M.; Sun, Z. One-body style photo-supercapacitors based on Ni(OH)2/TiO2 heterojunction array: High specific capacitance and ultra-fast charge/discharge response. J. Power Sources 2022, 521, 230920. [Google Scholar] [CrossRef]
  23. Gao, M.; Sheng, W.; Zhuang, Z.; Fang, Q.; Gu, S.; Jiang, J.; Yan, Y. Efficient Water Oxidation Using Nanostructured α-Nickel-Hydroxide as an Electrocatalyst. J. Am. Chem. Soc. 2014, 136, 7077–7084. [Google Scholar] [CrossRef]
  24. Wang, H.; Xiao, L.; Wang, C.; Lin, B.; Lv, S.; Chu, X.F.; Chi, Y.D.; Yang, X.T. Pd/TiO2 Nanospheres with Three-dimensional Hyperstructure for Enhanced Photodegradation of Organic Dye. Chem. Res. Chin. Univ. 2019, 35, 667–673. [Google Scholar] [CrossRef]
  25. Reddy, N.L.; Cheralathan, K.K.; Kumari, V.D.; Neppolian, B.; Venkatakrishnan, S.M. Photocatalytic Reforming of Biomass Derived Crude Glycerol in Water: A Sustainable Approach for Improved Hydrogen GenerationUsing Ni(OH)2 Decorated TiO2 Nanotubes under Solar Light Irradiation. ACS Sustain. Chem. Eng. 2018, 6, 3754–3764. [Google Scholar] [CrossRef]
  26. Cipagauta-Díaz, S.; Estrella-Gonz’alez, A.; Navarrete-Maga’na, M.; G’omez, R. N doped-TiO2 coupled to BiVO4 with high performance in photodegradation of Ofloxacin antibiotic and Rhodamine B dye under visible light. Catal. Today 2022, 394, 445–457. [Google Scholar] [CrossRef]
  27. Liu, F.; Xie, Y.; Yu, C.; Liu, X.; Dai, Y.; Liu, L.; Ling, Y. Novel hybrid Sr-doped TiO2/magneticNi0.6Zn0.4Fe2O4 for enhanced separation and photodegradation of organics under visible light. RSC Adv. 2015, 5, 24056–24063. [Google Scholar] [CrossRef]
  28. Kovalevskiy, N.; Selishchev, D.; Svintsitskiy, D.; Selishcheva, S.; Berezin, A.; Kozlov, D. Synergistic effect of polychromatic radiation on visible light activity of Ndoped TiO2 photocatalyst. Catal. Commun. 2020, 134, 105841. [Google Scholar] [CrossRef]
  29. Motlak, M.; Akhtar, M.S.; Barakat, N.A.M.; Hamza, A.M.; Yang, O.B.; Kim, H.Y. High-Efficiency Electrode Based on Nitrogen-Doped TiO2 Nanofibers for Dye-Sensitized Solar Cells. Electrochim. Acta 2014, 115, 493–498. [Google Scholar] [CrossRef]
  30. Liu, S.J.; Ma, Q.; Gao, F.; Song, S.H.; Gao, S. Relationship between N-doping induced point defects by annealing in ammonia and enhanced thermal stability for anodized titania nanotube arrays. J. Alloys Compd. 2012, 543, 71–78. [Google Scholar] [CrossRef]
  31. Peng, F.; Cai, L.; Yu, H.; Wang, H.; Yang, J. Synthesis and characterization of substitutional and interstitial nitrogen-doped titanium dioxides with visible light photocatalytic activity. J. Solid State Chem. 2008, 181, 130–136. [Google Scholar] [CrossRef]
  32. Chen, L.; Gu, Q.; Hou, L.; Zhang, C.; Lu, Y.; Wang, X.; Long, J. Molecular p-n heterojunction-enhanced visible light hydrogen evolution over a N-doped TiO2 photocatalyst. Catal. Sci. Technol. 2017, 7, 2039–2049. [Google Scholar] [CrossRef]
  33. Ran, J.; Yu, J.; Jaroniec, M. Ni(OH)2 modified CdS nanorods for highly efficient visible-light-driven photocatalytic H2 generation. Green Chem. 2011, 13, 2708–2713. [Google Scholar] [CrossRef]
  34. Mao, L.; Ba, Q.; Jia, X.; Liu, S.; Liu, H.; Zhang, J.; Li, X.; Chen, W. Ultrathin Ni(OH)2 nanosheets: A new strategy for cocatalyst design on CdS surfaces for photocatalytic hydrogen generation. RSC Adv. 2019, 9, 1260–1269. [Google Scholar] [CrossRef] [PubMed]
  35. Mou, J.; Xu, Y.; Zhong, D.; Chang, H.; Xu, C.; Wang, H.; Shen, H. Efcient Visible-Light-Responsive Photocatalytic Fuel Cell with a ZnIn2S4/UiO-66/TiO2/Ti Photoanode for Simultaneous RhB Degradation and Electricity Generation. J. Electron. Mater. 2022, 51, 6121–6133. [Google Scholar] [CrossRef]
  36. Nagajyothi, P.C.; Deyarayapalli, K.C.; Tettey, C.O.; Vattikuti, S.V.P.; Shim, J. Eco-friendly green synthesis: Catalytic activity of nickel hydroxide nanoparticles. Mater. Res. Express 2019, 6, 055036. [Google Scholar] [CrossRef]
  37. Shafi, A.; Ahmad, N.; Sultana, S.; Sabir, S.; Khan, M.Z. Ag2S-Sensitized NiO–ZnO Heterostructures with Enhanced Visible Light Photocatalytic Activity and Acetone Sensing Property. ACS Omega 2019, 4, 12905–12918. [Google Scholar] [CrossRef]
  38. Yang, M.; Lu, D.; Wang, H.; Wu, P.; Fang, P. A facile one-pot hydrothermal synthesis of two-dimensional TiO2-based nanosheets loaded with surface-enriched NixOy nanoparticles for efficient visible-light-driven photocatalysis. Appl. Surf. Sci. 2019, 467–468, 124–134. [Google Scholar] [CrossRef]
  39. Zhu, Q.; Liu, N.; Zhang, N.; Song, Y.; Stanislaus, M.S.; Zhao, C.; Yang, Y. Efficient photocatalytic removal of RhB, MO and MB dyes by optimized Ni/NiO/TiO2 composite thin films under solar light irradiation. J. Environ. Chem. Eng. 2018, 6, 2724–2732. [Google Scholar] [CrossRef]
  40. Peng, Y.; Cai, J.; Shi, Y.; Jiang, H.; Li, G. Thin p-type NiO nanosheet modified peanut-shaped monoclinic BiVO4 for enhanced charge separation and photocatalytic activities. Catal. Sci. Technol. 2022, 12, 5162–5170. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of TiO2, N-TiO2, Ni(OH)2, and N-TiO2/Ni(OH)2 with different amount of Ni(OH)2. The rectangle indicates the rutile phase of TiO2.
Figure 1. X-ray diffraction (XRD) patterns of TiO2, N-TiO2, Ni(OH)2, and N-TiO2/Ni(OH)2 with different amount of Ni(OH)2. The rectangle indicates the rutile phase of TiO2.
Micromachines 14 00870 g001
Figure 2. SEM, TEM, and high-resolution TEM (HRTEM) images of (a) TiO2 fibers, (b) N-TiO2 fibers, and (cf) N-TiO2/1.0 wt% Ni(OH)2 composite fibers.
Figure 2. SEM, TEM, and high-resolution TEM (HRTEM) images of (a) TiO2 fibers, (b) N-TiO2 fibers, and (cf) N-TiO2/1.0 wt% Ni(OH)2 composite fibers.
Micromachines 14 00870 g002
Figure 3. XPS spectra of (a) Ti 2p, (b) O 1s, (c) N 1s, and (d) Ni 2p of N-TiO2/1.0 wt% Ni(OH)2 nanofibers. The red lines in (b,c) are the fitted curves of the corresponding data.
Figure 3. XPS spectra of (a) Ti 2p, (b) O 1s, (c) N 1s, and (d) Ni 2p of N-TiO2/1.0 wt% Ni(OH)2 nanofibers. The red lines in (b,c) are the fitted curves of the corresponding data.
Micromachines 14 00870 g003
Figure 4. UV-Vis diffuse reflectance absorption spectra of Ni(OH)2, TiO2, N-TiO2, N-TiO2/Ni(OH)2 with different Ni(OH)2 contents, respectively.
Figure 4. UV-Vis diffuse reflectance absorption spectra of Ni(OH)2, TiO2, N-TiO2, N-TiO2/Ni(OH)2 with different Ni(OH)2 contents, respectively.
Micromachines 14 00870 g004
Figure 5. (a) Photodegradation of RhB for TiO2, N-TiO2, Ni(OH)2, and N-TiO2/Ni(OH)2 of different Ni(OH)2 contents under the irradiation of visible-light (>420 nm); (b) Natural logarithm degradation curves of different samples; (c) Rate constants (kR) of different samples for photodegradation of RhB; (d) five consecutive photodegeration curves of RhB over N-TiO2/1.0-wt% Ni(OH)2 under the visible light irradiation.
Figure 5. (a) Photodegradation of RhB for TiO2, N-TiO2, Ni(OH)2, and N-TiO2/Ni(OH)2 of different Ni(OH)2 contents under the irradiation of visible-light (>420 nm); (b) Natural logarithm degradation curves of different samples; (c) Rate constants (kR) of different samples for photodegradation of RhB; (d) five consecutive photodegeration curves of RhB over N-TiO2/1.0-wt% Ni(OH)2 under the visible light irradiation.
Micromachines 14 00870 g005
Figure 6. (a) PL spectra and (b) photocurrent response (>420 nm)of N-TiO2, Ni(OH)2, and N-TiO2/Ni(OH)2 with different Ni(OH)2 content, respectively.
Figure 6. (a) PL spectra and (b) photocurrent response (>420 nm)of N-TiO2, Ni(OH)2, and N-TiO2/Ni(OH)2 with different Ni(OH)2 content, respectively.
Micromachines 14 00870 g006
Figure 7. Schematic diagram of the mechanism of photocatalytic degradation of Rh B by Ni(OH)2/N-TiO2 under visible light irradiation.
Figure 7. Schematic diagram of the mechanism of photocatalytic degradation of Rh B by Ni(OH)2/N-TiO2 under visible light irradiation.
Micromachines 14 00870 g007
Table 1. Comparison of the catalytic performance of N-TiO2/Ni(OH)2 with the previously reported data for the reduction in organic dyes of Rh B.
Table 1. Comparison of the catalytic performance of N-TiO2/Ni(OH)2 with the previously reported data for the reduction in organic dyes of Rh B.
Material UsedLight UsedDye Degradation Kapp
(min−1)
Removal (%)/
30 min
Ref.
N-TiO2/Ni(OH)2visible lightRhB0.031060%This work
Ni(OH)2UV-visibleRhB0.0130/[36]
NiO-ZnOUV-visibleRhB0.030250[37]
NixOy/TiO2UV-visibleRhB0.034956.5[38]
Ni/NiO/TiO2UV-visibleRhB/35[39]
NiO/BVOUV-visibleRhB/40[40]
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

Wang, H.; Dong, M.; Shao, B.; Chi, Y.; Wang, C.; Lv, S.; Duan, R.; Wu, B.; Yang, X. Efficient Photodegradation of Rhodamine B by Fiber-like Nitrogen-Doped TiO2/Ni(OH)2 Nanocomposite under Visible Light Irradiation. Micromachines 2023, 14, 870. https://doi.org/10.3390/mi14040870

AMA Style

Wang H, Dong M, Shao B, Chi Y, Wang C, Lv S, Duan R, Wu B, Yang X. Efficient Photodegradation of Rhodamine B by Fiber-like Nitrogen-Doped TiO2/Ni(OH)2 Nanocomposite under Visible Light Irradiation. Micromachines. 2023; 14(4):870. https://doi.org/10.3390/mi14040870

Chicago/Turabian Style

Wang, Huan, Mingxuan Dong, Baorui Shao, Yaodan Chi, Chao Wang, Sa Lv, Ran Duan, Boqi Wu, and Xiaotian Yang. 2023. "Efficient Photodegradation of Rhodamine B by Fiber-like Nitrogen-Doped TiO2/Ni(OH)2 Nanocomposite under Visible Light Irradiation" Micromachines 14, no. 4: 870. https://doi.org/10.3390/mi14040870

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