Next Article in Journal
Valorization of Kraft Lignin from Black Liquor in the Production of Composite Materials with Poly(caprolactone) and Natural Stone Groundwood Fibers
Next Article in Special Issue
Electrospinning vs. Electro-Assisted Solution Blow Spinning for Fabrication of Fibrous Scaffolds for Tissue Engineering
Previous Article in Journal
Effect of Glass Filler Geometry on the Mechanical and Optical Properties of Highly Transparent Polymer Composite
Previous Article in Special Issue
Fabrication of Electrospun Cellulose Acetate/Nanoclay Composites for Pollutant Removal
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrothermal-Assisted Synthesis of Copper Nanoparticles-Decorated Titania Nanofibers for Methylene Blue Photodegradation and Catalyst for Sodium Borohydride Dehydrogenation

Department of Chemical Engineering, College of Engineering, Jazan University, Jazan 11451, Saudi Arabia
Polymers 2022, 14(23), 5180; https://doi.org/10.3390/polym14235180
Submission received: 8 October 2022 / Revised: 11 November 2022 / Accepted: 14 November 2022 / Published: 28 November 2022

Abstract

:
Simple and inexpensive electrospinning and hydrothermal techniques were used to synthesize titania nanofibers (TiO2 NFs) (composite NFs) decorated with copper nanoparticle (Cu NPs). The fabricated composite NFs have been tested as a photocatalytic material to degrade methylene blue (MB) as a model dye under visible light. The introduced composite NFs have shown good photocatalytic activity compared with pristine TiO2 NFs; 100% and 50% of dye were degraded in 120 min for composite NFs and pristine TiO2 NFs, respectively. Furthermore, composite NFs demonstrated good stability for four cycles. In addition, the fabricated Cu-TiO2 NFs have shown good photocatalytic activity for the production of H2 from sodium borohydride.

1. Introduction

The increase in water pollution, due to toxic organic pollutants, discharged from various industries, is harmful to human health and aquatic life systems, and causes aesthetic pollution problems [1,2,3]. In particular, the textile industry releases a large volume of water after dying and washing cotton fibers to remove any unfixed residual dyes [4,5,6]. The draining of colored wastewater into water resources such as rivers inhibits light passing and decreases dissolved oxygen in the water. These impacts on water negatively affect aquatic life [7,8]. Different traditional techniques have been used to remove the color of dyes (e.g., adsorption, flocculation, and coagulation) [6,7,8]. These processes cannot completely remove the dyes, but can only separate the dyes as sludge. This creates a secondary pollution problem [9,10]. Thus, a further treatment step is needed to remove the sludge from the purified water. The photocatalytic reaction approach, photodegradation, is widely recognized as a reliable strategy for de-colorizing wastewater without causing any further environmental harm [10,11,12]. Photodegradation is potent in environmental remediation since it converts toxic organic pollutants in water and air into non-toxic products. It is necessary to run the photodegradation process in the presence of a catalyst for heterogeneous metal oxides and a light source (visible or UV light) [13,14,15]. It is well known that titanium dioxide (TiO2) photocatalyst is the most applied catalyst in organic pollutants decomposition [16]. TiO2 possesses many advantages, such as its resistance to biological and chemical attacks, high oxidizing potential, inexpensiveness, environmental conscientiousness, and long-term resistance against chemical and photo corrosion [16,17]. Since TiO2 has a very large band gap (3.3 eV), photogenerated electrons and holes quickly recombine during the photodegradation process, causing the material to degrade [18,19]. As a result, its photodegradation processes have a poor quantum yield, reducing its photocatalytic efficacy [10,20]. To address this challenge and boost the commercialization of the photocatalytic method for degrading hazardous organic pollutants, several co-catalyst metal NPs are combined with TiO2 [21,22,23]. One way to improve the separation of holes and photogenerated electrons is to couple metal NPs with TiO2, which leads to band gap tuning [10]. Electrons are easily trapped by metal NPs due to their lower Fermi level compared to TiO2, whereas the holes are set at the valance band of TiO2, which produces highly intensive free radicals that degrade the organic pollutants in the solutions [10,24]. Furthermore, the fabrication protocol has a direct effect on the photocatalytic process. In this study, metallic copper NPs are incorporated in TiO2 nanofibers (NFs) as an efficient photocatalyst for the degradation of methylene blue. The photocatalyst is prepared using electrospinning and hydrothermal processes. The introduced photocatalyst showed complete degradation of MB as a model dye under the sunlight within 60 min. In addition, the prepared photocatalyst showed good catalytic activity towards hydrogen release from sodium borohydride.

2. Materials and Methods

2.1. Materials

Titanium isopropoxide (TIPP, 97%), poly (vinyl pyrrolidone) (PVP, MW = 13,000.00 g·mol−1), acetic acid (AA), and ethanol were purchased from Sigma Aldrich, St. Louis, MO, USA, and used as purchased, and used as purchased [14,17]. Copper (II) nitrate trihydrate solution (MW) was purchased from Junsei Chemical Co., Ltd., Chuo-ku, Tokyo.

2.2. TiO2 Nanofibers Preparation

TiO2 NFs were prepared using electrospinning followed by calcination at a high temperature. The catalytic fabrication process includes the following steps: (1) preparation of electrospinning homogenous solution, (2) electrospinning of the prepared solution, and (3) calcination of the polymeric electrospun NFs.
To prepare the electrospinning solution, first, a 15 wt% PVP polymer solution was prepared by dissolving 1.5 g PVP powder in a solvent mixture of AA and ethanol (50:50 wt%). The PVP solution was perfectly mixed at room temperature using a magnetic stirrer until a homogenous polymer solution was obtained. Then, 1 g of TIPP was added to the previous homogenous PVP solution. Continuous mixing of the (TIIP-PVP) electrospinning solution at room temperature was performed until a homogenous yellow transparent (TIIP-PVP) sol–gel was achieved.
Second, the electrospinning solution was converted into polymeric NFs using lab scale electrospinning machine. This process was started by injecting the prepared yellow sol–gel into a plastic syringe. The stainless-steel tip of the syringe was connected to the positive pole of a power supply, whereas the negative pole was connected to a rotating cylinder covered with a poly(ethylene) sheet. The electrospinning operating condition was fixed at a 15 cm working distance between the rotating drum and the needle tip, room temperature, 20 kV, and 0.8 mL h−1 flow rate. The obtained TIIP-PVP NFs sheet was vacuum-dried at 60 °C for a full day to ensure the evaporation of the residual solvent.
Finally, the calcination process was performed for 2 h at 700 °C to remove the PVP polymer and convert TIIP into TiO2 NFs.

2.3. Cu-Doped TiO2 Nanofibers Preparation

To prepare Cu-doped TiO2 NFs, 1 g of prepared TiO2 NFs was added to 0.025 M copper (II) nitrate trihydrate solution. The solution was stirred for 1 h at room temperature to disperse the Cu precursor particles on the surface of the TiO2 NFs. Finally, a reduction process was performed to convert the Cu precursor to Cu nanoparticles (NPs). Hydrazine hydrate (750 µL) and formic acid (750 µL) were drowsily added to reduce the previous solution [25,26]. The solution was stirred for 30 min at room temperature. The solution was finally transferred to an autoclave (Teflon-lined stainless steel) that was heated at 170 °C for 6 h in a muffle furnace. Finally, to clean the formed catalyst from impurities, it was filtrated and washed with ethanol and water, then dried at 80 °C for 24 h.

2.4. Characterization

The morphology of the fabricated electrospun catalyst was tested using a field emission scanning electron microscope (FESEM, Hitachi S-7400, Tokyo, Japan). A transmission electron microscope (TEM, JEOL Ltd., Tokyo, Japan) operating at 200 kV with EDX. UV-visible spectroscopy was also used to examine the concentration of dyes during the photodegradation process (HP 8453, Berlin, Germany).

2.5. Photocatalytic Activity Test

The photodegradation study of MB was executed in a batch reactor (glass bottle with 100 mL capacity). In total, 25 mg from the prepared photocatalysts was added to 50 mL of the 10 ppm MB solution and exposed for sunlight irradiation. The temperature of the solar irradiation was 29 ± 1 °C, with continuous stirring. In total, 2 mL from the solution was taken out at specific intervals and centrifuged to separate the residual NF catalyst. The absorbance of the separated MB solution was measured using a UV-visible spectrophotometer at λmax = 664 nm. A calibration curve was constructed between different concentrations and absorbance to measure unknown concentrations at defined absorbance.

2.6. Catalytic Hydrolysis of SBH

Using a lab-scale reactor, as reported in previous studies [27,28,29,30,31,32], hydrogen was released from SBH. An alkaline solution of 100 mM SBH was introduced to the reactor, along with a determined amount of Cu-TiO2 catalyst (37.83 mg, 1 mmol in 10 mL solution). The reactor was hooked up to a graduated cylinder containing water, and a magnet was used to agitate the solution at 900 rpm. The quantity of H2 released from SBH was calculated by monitoring the fall in water level in a graduated cylinder.

3. Results and Discussion

Based on its ease of use, cheap cost, high yield, and high efficiency, electrospinning stands out as the best method for producing NFs [25,33,34,35,36,37,38,39,40,41,42,43,44,45]. As known, size and morphology have direct effects on the chemical and physical properties of materials. The produced NFs have a diameter in the ranges of 50 to 1000 nm. The nanofibrous structure has a high axial ratio which could improve the catalytic performance in different reactions. After undergoing hydrothermal treatment, the NFs are shown at low and high magnification in FSEM images (Figure 1A,B, respectively). As a result of the process’s vigorous chemical reactions, a good nanofibrous structure is preserved. In addition, the high temperature, pressure, and reduction reaction that occur throughout the procedure allow nanopores to be seen extremely clearly in the resulting images. More morphological details of the prepared NFs were investigated by TEM and HRTEM images (Figure 2A,B). A normal TEM image (Figure 2A) shows the formation of hetero-structure NFs. According to the deposition process, the Cu NPs covered the surface of TiO2 NFs. The lattice fringes of NFs were determined to be 0.36 and 0.21 nm (Figure 2B), whose obtained values correspond to the TiO2 (101) and metallic Cu (111) planes, respectively [5,6]. Furthermore, few Cu species might be interstitial to the TiO2 matrix due to the thermally enhanced diffusion process. The respective HRTEM images (Figure 2B) indicated the formation of highly crystalline material. The image shows the formation of an interfacial region between Cu NPs, and the matrix is clear in the image. The inset in Figure 2B represents the SAED pattern which confirmed the formation of highly crystalline composite material without any defects. Figure 3 displays the TEM-EDX image for one selected NF; it confirmed the formation of hetero-structure NF with the rough and nano-porous surface.
Figure 3B–D indicate the EDX analysis of the drawn line in Figure 2A. The EDX analysis shows the formation of Cu, Ti, and O elements only. The copper NPs are homogeneity distributed along the Titania NF and its surface. In other words, copper NPs are grown at Titania matrix-like core–shell structure due to the deposition process; the interfacial region confirms this in the HRTEM image (Figure 2B).
The PL Emission spectra were used to investigate the semiconductor lifetime and charge separation in which PL indicates the electron/hole recombination rate in the semiconductor materials. The comparison of the PL spectrum (applied ʎ = 320 nm) of Titania NFs and copper-doped Titania NFs shows a similar spectrum at emission peaks 422 and 468 nm (Figure 4A). As seen in figure, copper-doped Titania NFs showed a lower intensity peak compared to Titania NFs which demonstrates a lower electrons/holes recombination rate and low defects in the copper-doped Titania NFs [46]. This might be attributed to the excited electrons from the valance band of Titania NFs to their conduction band and finally transfer to co-catalytic Cu NPs, which inhibit the electrons/holes recombination. This is desirable in exploiting the materials in the photocatalytic reaction.
Figure 4B shows the photodegradation study of MB used to prepare NFs to evaluate their activity. The copper-doped Titania NFs have shown a good photocatalytic performance compared with Titania NFs as 100% and 52% of dye have been removed at 120 min, respectively, under sunlight irradiation. It is evident from the PL result that copper-doped Titania NFs can be effectively used under visible light; however, pure Titania NFs can only be used as a photocatalyst in the ultraviolet spectrum [13]. For comparison, the adsorption effect of prepared NFs and photocatalytic activity without photocatalyst were studied. They have not shown any observable effect in dye degradation
To study the long performance of fabricated copper-doped Titania NFs, photocatalytic NFs were used for four cycles (Figure 5). As shown in Figure 4 below, good a photocatalytic response with little change in the catalytic performance was observed, which demonstrated that the introduced photocatalytic NFs are robust in the catalytic reaction.

Hydrolytic Dehydrogenation of NaBH4 Using Cu-TiO2 Catalyst

Metal hydrides such as ammonia borane (NH3BH3), hydrazine hydrate (N2H4.H2O), and sodium borohydride (NaBH4, SB) are very promising H2 storage materials due to their gravimetric density and high capacity for H2 storage [47,48]. SB is considered one of the best metal hydride H2 storage materials with 10.8% H2 capacity by weight. The hydrolysis of SB with water in the presence of an effective catalyst resulted in an exothermic reaction that produces H2. The reaction can occur at a low temperature to produce pure and controllable H2. SB hydrolysis reaction is irreversible and generates four moles of H2. Recently, Cu@TiO2 nanostructures have demonstrated excellent catalytic performance towards H2 generation from hydrogen storage materials [49,50]. Here, first, a controlled experiment was performed to demonstrate that at 30 ± 1 °C and 55 ± 1 °C, bare TiO2 had no catalytic activity in the hydrolysis process. Figure 6A depicts the percentage of H2 gas produced from the NaBH4 hydrolysis as a function of the reaction time in the presence of different amounts of the fabricated catalyst (Cu-TiO2) (75 mg, 100 mg, and 200 mg) and 1 mmol NaBH4 at 30 ± 1 °C. When Cu-TiO2 was added to the reactor, the NaBH4 hydrolysis began immediately. Increases in catalyst amount resulted in sustained hydrogen release for longer times. Accordingly, the rate at which hydrogen is produced is proportional to the amount of the catalyst. NaBH4 hydrolysis follows half-order kinetics with regard to the amount of Cu-TiO2 as shown by the straight line with a slope of 0.4 in Figure 6B. Figure 6B plots the rate of H2 production against the logarithmic values of Cu-TiO2 amount. Figure 7A shows the NaBH4 concentration (1 mmol, 2 mmol, and 3 mmol) and its influence on hydrogen production rate. Both the catalyst dose (75 mg) and the temperature (30 ± 1 °C) were kept constant. As shown in Figure 7A, as the concentration of NaBH4 increases, so does the rate at which hydrogen is produced. NaBH4 hydrolysis is first order in its dependence on [NaBH4], as seen by the straight line with a slope of 1.1 in a plot of hydrogen production rate vs. logarithmic values of [NaBH4] (Figure 7B). The activation energy (Ea) of Cu-TiO2 in the NaBH4 hydrolysis process was measured by catalyzing the reaction using Cu-TiO2 = 75 mg and 1 mmole NaBH4 at different temperatures (Figure 8A). Increasing the reaction temperature results in a relatively larger hydrogen generation. By plotting the logarithmic k and KD values as a function of the inverse temperature (1/T), where k is the rate constant for hydrogen generation, the Arrhenius and Eyring plots (shown in Figure 8B,C) can be constructed using Equations (1) and (2), respectively.
k = A e E a R T
ln K D = ( Δ S ° R ) ( Δ H ° R T )
From the Arrhenius and Eyring plots, we determine that the reaction thermodynamic parameters (Ea, S, and H) are 19.03, 0.02894, and 16.48 kJ/mol.

4. Conclusions

Cu-NP-decorated TiO2 NFs were successfully prepared using a simple electrospinning technique followed by a hydrothermal process. The characterization technique confirmed the deposition of CuNPs on the surface of TiO2 NFs. PL data showed the lower recombination of electrons and holes compared to TiO2 NFs. Accordingly, composite NFs have been shown to have good photocatalytic performance compared with TiO2 NFs. The introduced composite NFs have shown good photocatalytic activity compared with pristine TiO2 NFs; 100% and 50% of dye are degraded in 120 min for Composite NFs and pristine TiO2 NFs, respectively. In addition, the fabricated Cu-TiO2 NFs have shown good photocatalytic activity for the production of H2 from sodium borohydride. The H2 generation yield has been increased from 63% to 90% with an increase in the catalyst amount from 75 mg to 200 mg in the presence of 1 mmol sodium borohydride and 30 °C. This indicated that the reaction is catalyst-dependent. The kinetics study showed that the reaction followed the pseudo first-order reaction with respect to the concentration of sodium borohydride. The hydrogen generation has been increased with the increase in reaction temperature. A low activation energy (19.03 kJ mol−1) is obtained.

Funding

The author extends his appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia for funding this research work through the project number ISP20-30.

Acknowledgments

The author acknowledges the support from Jazan University.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Schwarzenbach, R.P.; Egli, T.; Hofstetter, T.B.; Von Gunten, U.; Wehrli, B. Global water pollution and human health. Annu. Rev. Environ. Resour. 2010, 35, 109–136. [Google Scholar] [CrossRef]
  2. Carmen, Z.; Daniela, S. Textile Organic Dyes–Characteristics, Polluting Effects and Separation/Elimination Procedures from Industrial Effluents–A Critical Overview, Organic Pollutants Ten Years after the Stockholm Convention-Environmental and Analytical Update; IntechOpen: London, UK, 2012; p. 32373. [Google Scholar]
  3. Khan, S.; Malik, A. Environmental and Health Effects of Textile Industry Wastewater, Environmental Deterioration and Human Health; Springer: Berlin/Heidelberg, Germany, 2014; pp. 55–71. [Google Scholar]
  4. Kant, R. Textile Dyeing Industry An Environmental Hazard. Nat. Sci. 2011, 4, 1. [Google Scholar] [CrossRef] [Green Version]
  5. Saini, R.D. Textile organic dyes: Polluting effects and elimination methods from textile waste water. Int. J. Chem. Eng. Res. 2017, 9, 121–136. [Google Scholar]
  6. Verma, A.K.; Dash, R.R.; Bhunia, P. A review on chemical coagulation/flocculation technologies for removal of colour from textile wastewaters. J. Environ. Manag. 2012, 93, 154–168. [Google Scholar] [CrossRef]
  7. Yagub, M.T.; Sen, T.K.; Afroze, S.; Ang, H.M. Dye and its removal from aqueous solution by adsorption: A review. Adv. Colloid Interface Sci. 2014, 209, 172–184. [Google Scholar] [CrossRef]
  8. Beluci, N.d.C.L.; Mateus, G.A.P.; Miyashiro, C.S.; Homem, N.C.; Gomes, R.G.; Fagundes-Klen, M.R.; Bergamasco, R.; Vieira, A.M.S. Hybrid treatment of coagulation/flocculation process followed by ultrafiltration in TIO2-modified membranes to improve the removal of reactive black 5 dye. Sci. Total Environ. 2019, 664, 222–229. [Google Scholar] [CrossRef] [PubMed]
  9. Jeon, S.; Yun, J.; Lee, Y.-S.; Kim, H.-I. Preparation of poly (vinyl alcohol)/poly (acrylic acid)/TiO2/carbon nanotube composite nanofibers and their photobleaching properties. J. Ind. Eng. Chem. 2012, 18, 487–491. [Google Scholar] [CrossRef]
  10. Yousef, A.; El-Halwany, M.; Barakat, N.A.; Al-Maghrabi, M.N.; Kim, H.Y. Cu0-doped TiO2 nanofibers as potential photocatalyst and antimicrobial agent. J. Ind. Eng. Chem. 2015, 26, 251–258. [Google Scholar] [CrossRef]
  11. Xie, M.; Liu, X.; Wang, S. Degradation of methylene blue through Fenton-like reaction catalyzed by MoS2-doped sodium alginate/Fe hydrogel. Colloids Surf. B Biointerfaces 2022, 214, 112443. [Google Scholar] [CrossRef] [PubMed]
  12. Panthi, G.; Park, M.; Kim, H.-Y.; Lee, S.-Y.; Park, S.-J. Electrospun ZnO hybrid nanofibers for photodegradation of wastewater containing organic dyes: A review. J. Ind. Eng. Chem. 2015, 21, 26–35. [Google Scholar] [CrossRef]
  13. Yousef, A.; Barakat, N.A.; Amna, T.; Unnithan, A.R.; Al-Deyab, S.S.; Kim, H.Y. Influence of CdO-doping on the photoluminescence properties of ZnO nanofibers: Effective visible light photocatalyst for waste water treatment. J. Lumin. 2012, 132, 1668–1677. [Google Scholar] [CrossRef]
  14. Yousef, A.; Barakat, N.A.; Kim, H.Y. Electrospun Cu-doped titania nanofibers for photocatalytic hydrolysis of ammonia borane. Appl. Catal. A Gen. 2013, 467, 98–106. [Google Scholar] [CrossRef]
  15. Yousef, A.; Barakat, N.A.; Al-Deyab, S.S.; Nirmala, R.; Pant, B.; Kim, H.Y. Encapsulation of CdO/ZnO NPs in PU electrospun nanofibers as novel strategy for effective immobilization of the photocatalysts. Colloids Surf. A Physicochem. Eng. Asp. 2012, 401, 8–16. [Google Scholar] [CrossRef]
  16. Panthi, G.; Yousef, A.; Barakat, N.A.; Khalil, K.A.; Akhter, S.; Choi, Y.R.; Kim, H.Y. Mn2O3/TiO2 nanofibers with broad-spectrum antibiotics effect and photocatalytic activity for preliminary stage of water desalination. Ceram. Int. 2013, 39, 2239–2246. [Google Scholar] [CrossRef]
  17. Yousef, A.; Brooks, R.M.; El-Halwany, M.M.; EL-Newehy, M.H.; Al-Deyab, S.S.; Barakat, N.A. Cu0/S-doped TiO2 nanoparticles-decorated carbon nanofibers as novel and efficient photocatalyst for hydrogen generation from ammonia borane. Ceram. Int. 2016, 42, 1507–1512. [Google Scholar] [CrossRef]
  18. Lee, S.S.; Bai, H.; Liu, Z.; Sun, D.D. Novel-structured electrospun TiO2/CuO composite nanofibers for high efficient photocatalytic cogeneration of clean water and energy from dye wastewater. Water Res. 2013, 47, 4059–4073. [Google Scholar] [CrossRef]
  19. Chong, M.N.; Jin, B.; Chow, C.W.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef]
  20. Zhu, H.-Y.; Xiao, L.; Jiang, R.; Zeng, G.-M.; Liu, L. Efficient decolorization of azo dye solution by visible light-induced photocatalytic process using SnO2/ZnO heterojunction immobilized in chitosan matrix. Chem. Eng. J. 2011, 172, 746–753. [Google Scholar] [CrossRef]
  21. Doustkhah, E.; Assadi, M.H.N.; Komaguchi, K.; Tsunoji, N.; Esmat, M.; Fukata, N.; Tomita, O.; Abe, R.; Ohtani, B.; Ide, Y. In situ Blue titania via band shape engineering for exceptional solar H2 production in rutile TiO2. Appl. Catal. B Environ. 2021, 297, 120380. [Google Scholar] [CrossRef]
  22. Kerkez-Kuyumcu, Ö.; Kibar, E.; Dayıoğlu, K.; Gedik, F.; Akın, A.N.; Özkara-Aydınoğlu, Ş. A comparative study for removal of different dyes over M/TiO2 (M= Cu, Ni, Co, Fe, Mn and Cr) photocatalysts under visible light irradiation. J. Photochem. Photobiol. A Chem. 2015, 311, 176–185. [Google Scholar] [CrossRef]
  23. Tayade, R.J.; Kulkarni, R.G.; Jasra, R.V. Transition metal ion impregnated mesoporous TiO2 for photocatalytic degradation of organic contaminants in water. Ind. Eng. Chem. Res. 2006, 45, 5231–5238. [Google Scholar] [CrossRef]
  24. Carp, O.; Huisman, C.L.; Reller, A. Photoinduced reactivity of titanium dioxide. Prog. Solid State Chem. 2004, 32, 33–177. [Google Scholar] [CrossRef]
  25. Yousef, A.; Barakat, N.A.; El-Newehy, M.H.; Ahmed, M.; Kim, H.Y. Catalytic hydrolysis of ammonia borane for hydrogen generation using Cu (0) nanoparticles supported on TiO2 nanofibers. Colloids Surf. A Physicochem. Eng. Asp. 2015, 470, 194–201. [Google Scholar] [CrossRef]
  26. Giannousi, K.; Lafazanis, K.; Arvanitidis, J.; Pantazaki, A.; Dendrinou-Samara, C. Hydrothermal synthesis of copper based nanoparticles: Antimicrobial screening and interaction with DNA. J. Inorg. Biochem. 2014, 133, 24–32. [Google Scholar] [CrossRef]
  27. Al-Enizi, A.M.; Brooks, R.M.; Abutaleb, A.; El-Halwany, M.; El-Newehy, M.H.; Yousef, A.J.C.I. Electrospun carbon nanofibers containing Co-TiC nanoparticles-like superficial protrusions as a catalyst for H2 gas production from ammonia borane complex. Ceram. Int. 2017, 43, 15735–15742. [Google Scholar] [CrossRef]
  28. Yousef, A.; Brooks, R.M.; El-Halwany, M.M.; Abutaleb, A.; El-Newehy, M.H.; Al-Deyab, S.S.; Kim, H.Y. Electrospun CoCr7C3-supported C nanofibers: Effective, durable, and chemically stable catalyst for H2 gas generation from ammonia borane. Mol. Catal. 2017, 434, 32–38. [Google Scholar] [CrossRef]
  29. Yousef, A.; Barakat, N.A.; Khalil, K.A.; Unnithan, A.R.; Panthi, G.; Pant, B.; Kim, H.Y.J.C.; Physicochemical, S.A.; Aspects, E. Photocatalytic release of hydrogen from ammonia borane-complex using Ni (0)-doped TiO2/C electrospun nanofibers. Colloids Surf. A Physicochem. Eng. Asp. 2012, 410, 59–65. [Google Scholar] [CrossRef]
  30. Yousef, A.; Brooks, R.M.; El-Halwany, M.; Obaid, M.; El-Newehy, M.H.; Al-Deyab, S.S.; Barakat, N.A. A novel and chemical stable Co–B nanoflakes-like structure supported over titanium dioxide nanofibers used as catalyst for hydrogen generation from ammonia borane complex. Int. J. Hydrogen Energy 2016, 41, 285–293. [Google Scholar] [CrossRef]
  31. Yousef, A.; El-Halwany, M.; Barakat, N.A.; Kim, H.Y. One pot synthesis of Cu-doped TiO2 carbon nanofibers for dehydrogenation of ammonia borane. Ceram. Int. 2015, 41, 6137–6140. [Google Scholar] [CrossRef]
  32. Abutaleb, A.; Zouli, N.; El-Halwany, M.M.; Ubaidullah, M.; Yousef, A. Graphitic nanofibers supported NiMn bimetallic nanoalloys as catalysts for H2 generation from ammonia borane. Int. J. Hydrogen Energy 2021, 46, 35248–35260. [Google Scholar] [CrossRef]
  33. Yousef, A.; Akhtar, M.S.; Barakat, N.A.; Motlak, M.; Yang, O.-B.; Kim, H.Y. Effective NiCu NPs-doped carbon nanofibers as counter electrodes for dye-sensitized solar cells. Electrochim. Acta 2013, 102, 142–148. [Google Scholar] [CrossRef]
  34. Al-Enizi, A.M.; Nafady, A.; El-Halwany, M.M.; Brooks, R.M.; Abutaleb, A.; Yousef, A. Electrospun carbon nanofiber-encapsulated NiS nanoparticles as an efficient catalyst for hydrogen production from hydrolysis of sodium borohydride. Int. J. Hydrog. Energy 2019, 44, 21716–21725. [Google Scholar] [CrossRef]
  35. Yousef, A.; Brooks, R.M.; Abutaleb, A.; El-Newehy, M.H.; Al-Deyab, S.S.; Kim, H.Y. One-step synthesis of Co-TiC-carbon composite nanofibers at low temperature. Ceram. Int. 2017, 43, 5828–5831. [Google Scholar] [CrossRef]
  36. Al-Enizi, A.M.; El-Halwany, M.M.; Al-Abdrabalnabi, M.A.; Bakrey, M.; Ubaidullah, M.; Yousef, A. Novel Low Temperature Route to Produce CdS/ZnO Composite Nanofibers as Effective Photocatalysts. Catalysts 2020, 10, 417. [Google Scholar] [CrossRef] [Green Version]
  37. Jiang, W.; Zhao, P.; Song, W.; Wang, M.; Yu, D.-G. Electrospun zein/polyoxyethylene core-sheath ultrathin fibers and their antibacterial food packaging applications. Biomolecules 2022, 12, 1110. [Google Scholar] [CrossRef] [PubMed]
  38. Jiang, W.; Zhang, X.; Liu, P.; Zhang, Y.; Song, W.; Yu, D.-G.; Lu, X. Electrospun healthcare nanofibers from medicinal liquor of Phellinus igniarius. Adv. Compos. Hybrid Mater. 2022, 5, 3045–3056. [Google Scholar] [CrossRef]
  39. Abbas, M.; Hameed, R.A.; Al-Enizi, A.M.; Thamer, B.M.; Yousef, A.; El-Newehy, M.H. Decorated carbon nanofibers with mixed nickel− manganese carbides for methanol electro-oxidation in alkaline solution. Int. J. Hydrogen Energy 2021, 46, 6494–6512. [Google Scholar] [CrossRef]
  40. Al-Enizi, A.M.; Brooks, R.M.; El-Halwany, M.; Yousef, A.; Nafady, A.; Hameed, R.A. CoCr7C3-like nanorods embedded on carbon nanofibers as effective electrocatalyst for methanol electro-oxidation. Int. J. Hydrogen Energy 2018, 43, 9943–9953. [Google Scholar] [CrossRef]
  41. Maafa, I.M.; Al-Enizi, A.M.; Abutaleb, A.; Zouli, N.I.; Ubaidullah, M.; Shaikh, S.F.; Al-Abdrabalnabi, M.A.; Yousef, A. One-pot preparation of CdO/ZnO core/shell nanofibers: An efficient photocatalyst. Alex. Eng. J. 2021, 60, 1819–1826. [Google Scholar] [CrossRef]
  42. Yousef, A.; Brooks, R.M.; El-Newehy, M.H.; Al-Deyab, S.S.; Kim, H.Y. Electrospun Co-TiC nanoparticles embedded on carbon nanofibers: Active and chemically stable counter electrode for methanol fuel cells and dye-sensitized solar cells. Int. J. Hydrogen Energy 2017, 42, 10407–10415. [Google Scholar] [CrossRef]
  43. Zouli, N.; Hameed, R.A.; Abutaleb, A.; El-Halwany, M.M.; El-Newehy, M.H.; Yousef, A. Insights on the role of supporting electrospun carbon nanofibers with binary metallic carbides for enhancing their capacitive deionization performance. J. Mater. Res. Technol. 2021, 15, 3795–3806. [Google Scholar] [CrossRef]
  44. Al-Enizi, A.M.; Karim, A.; Yousef, A. A novel method for fabrication of electrospun cadmium sulfide nanoparticles-decorated zinc oxide nanofibers as effective photocatalyst for water photosplitting. Alex. Eng. J. 2022, in press. [CrossRef]
  45. Hameed, R.M.A.; Zouli, N.; Abutaleb, A.; El-Halwany, M.M.; El-Newehy, M.H.; Yousef, A. Improving water desalination performance of electrospun carbon nanofibers by supporting with binary metallic carbide nanoparticles. Ceram. Int. 2022, 48, 4741–4753. [Google Scholar] [CrossRef]
  46. Xu, S.; Sun, D.D. Significant improvement of photocatalytic hydrogen generation rate over TiO2 with deposited CuO. Int. J. Hydrogen Energy 2009, 34, 6096–6104. [Google Scholar] [CrossRef]
  47. Yu, J.; Hai, Y.; Jaroniec, M. Photocatalytic hydrogen production over CuO-modified titania. J. Colloid Interface Sci. 2011, 357, 223–228. [Google Scholar] [CrossRef]
  48. Al-Enizi, A.M.; El-Halwany, M.M.; Shaikh, S.F.; Pandit, B.; Yousef, A. Electrospun nickel nanoparticles@ poly (vinylidene fluoride-hexafluoropropylene) nanofibers as effective and reusable catalyst for H2 generation from sodium borohydride. Arab. J. Chem. 2022, 15, 104207. [Google Scholar] [CrossRef]
  49. Esmat, M.; Doustkhah, E.; Abdelbar, M.; Tahawy, R.; El-Hosainy, H.; Abdelhameed, M.; Ide, Y.; Fukata, N. Structural Conversion of Cu-Titanate into Photoactive Plasmonic Cu-TiO2 for H2 Generation in Visible Light. ACS Sustain. Chem. Eng. 2022, 10, 4143–4151. [Google Scholar] [CrossRef]
  50. Kılınç, D.; Şahin, Ö. Effective TiO2 supported Cu-Complex catalyst in NaBH4 hydrolysis reaction to hydrogen generation. Int. J. Hydrogen Energy 2019, 44, 18858–18865. [Google Scholar] [CrossRef]
Figure 1. Low and high magnification FE-SEM images of the produced powder after hydrothermal process.
Figure 1. Low and high magnification FE-SEM images of the produced powder after hydrothermal process.
Polymers 14 05180 g001
Figure 2. (A) Normal TEM image and (B) HR-TEM image of produced powder after the hydrothermal process (Inset in b shows the SAED image).
Figure 2. (A) Normal TEM image and (B) HR-TEM image of produced powder after the hydrothermal process (Inset in b shows the SAED image).
Polymers 14 05180 g002
Figure 3. (A) STEM image along with the line EDX analysis of the produced powder after the hydrothermal process; (BD) indicate line analysis EDX for the line shown in (A).
Figure 3. (A) STEM image along with the line EDX analysis of the produced powder after the hydrothermal process; (BD) indicate line analysis EDX for the line shown in (A).
Polymers 14 05180 g003
Figure 4. Photoluminescence spectrum of composite NFs and TiO2 NFs (A) and photodegradation profiles of the MB under sunlight irradiation (B).
Figure 4. Photoluminescence spectrum of composite NFs and TiO2 NFs (A) and photodegradation profiles of the MB under sunlight irradiation (B).
Polymers 14 05180 g004
Figure 5. Reusability test of composite NFs.
Figure 5. Reusability test of composite NFs.
Polymers 14 05180 g005
Figure 6. (A) Effect of Cu−TiO2 amount on hydrogen generation from aqueous NaBH4; (B) plot of logarithmic hydrogen generation rate vs. Cu−TiO2 amount.
Figure 6. (A) Effect of Cu−TiO2 amount on hydrogen generation from aqueous NaBH4; (B) plot of logarithmic hydrogen generation rate vs. Cu−TiO2 amount.
Polymers 14 05180 g006
Figure 7. (A) Effect of NaBH4 concentration on the catalytic performance; (B) plot of logarithmic of hydrogen generation rate vs. [NaBH4].
Figure 7. (A) Effect of NaBH4 concentration on the catalytic performance; (B) plot of logarithmic of hydrogen generation rate vs. [NaBH4].
Polymers 14 05180 g007
Figure 8. (A) Effect of temperature on hydrogen released from aqueous NaBH4 in the presence of Cu-TiO2; (B) plot of ln K vs. 1/T; (C) plot of ln (K/T) vs. 1/T.
Figure 8. (A) Effect of temperature on hydrogen released from aqueous NaBH4 in the presence of Cu-TiO2; (B) plot of ln K vs. 1/T; (C) plot of ln (K/T) vs. 1/T.
Polymers 14 05180 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Abutaleb, A. Hydrothermal-Assisted Synthesis of Copper Nanoparticles-Decorated Titania Nanofibers for Methylene Blue Photodegradation and Catalyst for Sodium Borohydride Dehydrogenation. Polymers 2022, 14, 5180. https://doi.org/10.3390/polym14235180

AMA Style

Abutaleb A. Hydrothermal-Assisted Synthesis of Copper Nanoparticles-Decorated Titania Nanofibers for Methylene Blue Photodegradation and Catalyst for Sodium Borohydride Dehydrogenation. Polymers. 2022; 14(23):5180. https://doi.org/10.3390/polym14235180

Chicago/Turabian Style

Abutaleb, Ahmed. 2022. "Hydrothermal-Assisted Synthesis of Copper Nanoparticles-Decorated Titania Nanofibers for Methylene Blue Photodegradation and Catalyst for Sodium Borohydride Dehydrogenation" Polymers 14, no. 23: 5180. https://doi.org/10.3390/polym14235180

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