Next Article in Journal
Recent Trends in the Design, Synthesis and Biomedical Applications of Covalent Organic Frameworks
Previous Article in Journal
High-Temperature Response Polylactic Acid Composites by Tuning Double-Percolated Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fabrications of Electrospun Mesoporous TiO2 Nanofibers with Various Amounts of PVP and Photocatalytic Properties on Methylene Blue (MB) Photodegradation

1
Department of Advanced Materials Engineering, Gangneung-Wonju National University, 7 Jukheongil, Gangneung 25457, Republic of Korea
2
Department of Energy Engineering, Konkuk University, 120 Neungdong-ro, Gwangjin-gu, Seoul 05029, Republic of Korea
3
Research Institute for Dental Engineering, Gangneung-Wonju National University, Gangneung 25457, Republic of Korea
4
Smart Hydrogen Energy Center, Gangneung-Wonju National University, 7 Jukheongil, Gangneung 25457, Republic of Korea
*
Authors to whom correspondence should be addressed.
Polymers 2023, 15(1), 134; https://doi.org/10.3390/polym15010134
Submission received: 16 November 2022 / Revised: 17 December 2022 / Accepted: 26 December 2022 / Published: 28 December 2022
(This article belongs to the Section Polymer Applications)

Abstract

:
The superior chemical and electrical properties of TiO2 are considered to be suitable material for various applications, such as photoelectrodes, photocatalysts, and semiconductor gas sensors; however, it is difficult to commercialize the applications due to their low photoelectric conversion efficiency. Various solutions have been suggested and among them, the increase of active sites through surface modification is one of the most studied methods. A porous nanostructure with a large specific surface area is an attractive solution to increasing active sites, and in the electrospinning process, mesoporous nanofibers can be obtained by controlling the composition of the precursor solution. This study successfully carried out surface modification of TiO2 nanofibers by mixing polyvinylpyrrolidone with different molecular weights and using diisopropyl azodicarboxylate (DIPA). The morphology and crystallographic properties of the TiO2 samples were analyzed using a field emission electron microscope and X-ray diffraction method. The specific surface area and pore properties of the nanofiber samples were compared using the Brunauer-Emmett-Teller method. The TiO2 nanofibers fabricated by the precursor with K-30 polyvinyl pyrrolidone and diisopropyl azodicarboxylate were more porous than the TiO2 nanofibers without them. The modified nanofibers with K-30 and DIPA had a photocatalytic efficiency of 150% compared to TiO2 nanofibers. Their X-ray diffraction patterns revealed anatase peaks. The average crystallite size of the modified nanofibers was calculated to be 6.27–9.27 nm, and the specific surface area was 23.5–27.4 m2/g, which was more than 150% larger than the 17.2 m2/g of ordinary TiO2 nanofibers.

Graphical Abstract

1. Introduction

In the last decades, environmental pollution due to the increase in population and resource consumption has become a critical issue. Environmental pollution caused by industrial waste disrupts sustainable development and causes multiple problems. The use of fossil fuels, such as coal, oil, and natural gas, increases the amount of carbon dioxide and fine dust in the atmosphere, and causes air pollution [1,2,3,4,5,6,7]; in addition, various types of domestic and industrial wastewater cause water pollution [8,9]. Many research studies on the capture, storage, and removal of pollutants have been conducted and solutions proposed for sustainable development [10,11,12,13,14]. Among them, the degradation of pollutants using semiconductor photocatalysts is one of the important ideas [15,16,17,18,19,20,21]. A photocatalyst for the degradation of water pollutants requires various properties, such as non-toxicity, chemical resistance, and an electronic structure that can use sunlight; in addition, TiO2 has attracted much attention as a photocatalyst due to its superior chemical and electrical properties [22,23,24]. However, the industrial application of TiO2 has been limited because of its poor photoelectric conversion efficiency, which is due to its wide band gap (~3.2 eV) and the rapid recombination rate of electron-hole pairs [25,26,27,28,29]. Synthesis methods, such as hydrothermal, microwave, anodization, chemical vapor deposition, and electrospinning, have been reported for the fabrication of various TiO2 nanostructures [30,31,32,33,34,35,36,37]; in addition, among them, electrospinning is a simple and economic method to obtain uniform and continuous one-dimensional nanofibers. Electrospinning is a process to obtain nanofibers with a nanoscale diameter through jet stretching of a viscous polymer solution by an electric field and rapid solvent volatilization. At the same time, the improvement of nanofibers, such as element doping and morphology modification, can be achieved simply by controlling the composition of the precursor solution [38,39,40,41,42,43,44,45]. In addition, to improve photocatalytic activity, studies on the synthesis of TiO2 nanofibers containing noble metals and transition metal dopants, such as Ag, Cu, C, Fe, and Co, have been prominently reported [46,47,48,49,50].
In this study, we modified mesoporous nanofibers using polyvinylpyrrolidone with different molecular weights and diisopropyl azodicarboxylate (DIPA) for a precursor solution. Previous studies using titanium butoxide, polyvinyl pyrrolidone (PVP), and polyvinyl alcohol have been reported for the mesoporous electrospun nanofibers fabrication; however, studies using a titanium tetraisopropoxide (TTIP) and PVP mixture of different molecular weights are difficult to find. The surface and pore properties of the nanofibers were improved via control of the calcination temperature ramping speed. To compare the modifed TiO2 nanofibers (TNF) and ordinary nanofibers, a methylene blue (MB) photodegradation test was carried out using a UV lamp, and the absorbance of the MB solution was measured using UV–Vis. The test results confirmed that the modified nanofibers achieved improved photocatalytic efficiency over ordinary nanofibers [51,52].

2. Materials and Methods

2.1. Materials

PVP (K-90, polyvinyl pyrrolidone, M.W. 1,300,000), (K-30, polyvinyl pyrrolidone, M.W. 58,000), and DIPA (diisopropyl azodicarboxylate, 94%) were purchased from Alfa Aesar Korea Co., Ltd. (Incheon, Republic of Korea). TTIP (titanium tetraisopropoxide, ≥98.0%) and ACAC (acetyl acetone, ≥99.0%) were purchased from Junsei Co., Ltd. (Tokyo, Japan). EtOH (ethyl alcohol, ≥99.5% EP) was purchased from Daejung (Siheung, Republic of Korea).

2.2. Fabrication of TiO2 Nanofibers

As shown in Table 1, precursor solutions were prepared under different compositions and calcination conditions. First, 10 g of PVP was added to EtOH 50 g and stirred for 24 h using a magnetic stirrer. TNF 1, 2, and 3 were prepared using 5 g of K-90 and 5g of K-30. After that, a second solution was prepared by mixing 15 g of TTIP and 10 g of ACAC in another beaker for 3 h. Then, the second solution was added to the first solution and stirred for 3 h; then, 10 g of DIPA was added to the mixture solution and vigorously stirred until a yellow transparent solution was obtained.
The prepared precursor solutions were electrospun using a nozzle capillary with a diameter of 0.34 mm (23 gauge) at each 1.0 mL/h flow rate. The tip-to-collector distance was kept at 15 cm, and the applied voltage was set to 15 kV. The process was carried out at room temperature and humidity below 50%, and aluminum foil was used as the collector.
A schematic diagram of the electrospinning process is shown in Figure 1. A 3-tip multi-nozzle was used in the process, and 4 h was set as 1 batch. The weight of as-spun nanofibers collected per batch was about 4 g; then, dried and calcined.

2.3. Characterization

As-spun nanofibers were dried in an oven heated to 60 °C for 2 h and then, calcined by heating in a box furnace to 500 °C.
Thermogravimetric analysis (TGA, Q500, TA Instruments, Newcastle, DE, USA) was carried out to decide the calcination temperature of as-spun nanofibers. The surface and average diameter of each sample were measured using a field emission scanning electron microscope (FE-SEM, Inspect F, FEI Korea Co., Ltd., Gyeonggi-do, Republic of Korea). FT-IR spectra of the calcined TiO2 nanofibers were obtained using a Fourier transform infrared spectrometer (FT-IR, iS50, Thermo Fisher Scientific, Waltham, MA, USA), with a KBr pellet in the range 400 and 4000 cm−1. The crystal structure and crystallinity of the TiO2 sample were analyzed using the diffraction pattern obtained using an X-ray diffractometer (XRD, AXS-D8, Bruker Korea Co., Ltd., Gyeonggi-do, Republic of Korea). The specific surface area and pore characteristics were analyzed using a Brunauer-Emmett-Teller surface area analyzer (TriStar II 3020, Micromeritics, Norcross, GA, USA), and the adsorption/desorption of N2 gas was carried out at 77.3 K.

2.4. Photocatalytic MB Degradation

The MB degradation test was carried out in a dark room to compare the photocatalytic efficiency of each sample. An aqueous solution of 5 mg/L was prepared using MB as a pollutant; in addition, 0.2 g of the TiO2 nanofiber sample and 20 mL of deionized water were added to a quartz beaker, and stirred for 30 min. After that, 200 mL of MB aqueous solution was added to the TiO2 dispersed water and stirred for 2 h in a dark room, with temperature control for mixing and stabilizing. After that, the distance between the UV lamp (6 W, 365 nm) and the quartz beaker was fixed at 10 cm, and the beaker was stirred at 240 rpm during the photodegradation reaction. The reaction was carried out for 3 h, and the mixed solution was sampled every 30 min using a syringe. The TiO2 photocatalyst in the sampled solution was filtered and removed using a syringe filter (PVDF filter, 0.2 μm, Whatman, Marlborough, MA, USA); in addition, the filtered solution was stored in a cuvette and refrigerated to block the incident light.

3. Results and Discussion

A TGA analysis was carried out to determine the calcination temperature of TNF, including a blowing agent. The thermal behavior of TNF0 and TNF1 was compared to confirm that the same calcination temperature of ordinary TNF could be applied, and the results are shown in Figure 2. After setting the initial temperature to 30 °C and stabilizing, it was heated to 500 °C for a heating rate of 5 °C/min. Both the balance gas and sample gas flow rates were set to 50.0 mL/min of N2 gas, and each TNF sample weighed 10 mg. The change of the TGA curve was observed in three steps, and the evaporation of the solvent and adsorbed water evaporation was not significant due to the drying process. The first section at ~150 °C was considered as mass reduction via the evaporation of moisture and residual solvents. Starting at 210 °C, the second section can be considered the glass transition and combustion phase. At this time, the decrease rate of TNF 1 was greater than that of TNF 0, which is considered due to the lower burning point of DIPA and low molecular weight PVP. When the section is above 390 °C, the thermal decomposition of the carbonized compound and the residual polymer starts. After 450 °C, the final weights of TNF 0 and TNF 1 were 29.6 wt% and 32.1 wt%, respectively; in addition, it was confirmed that there was no significant difference and the thermal behavior was ended. The calcination temperature was determined to be 500 °C to prevent rutile growth starting at 600 °C and higher, while providing a sufficient driving force for crystallization [53,54].
The surface structure and microstructure of the TiO2 nanofiber (TNF) samples were analyzed using FE-SEM as shown in Figure 3. Figure 3a,b are low and high magnification images of TNF 0, which are ordinary nanofibers fabricated via typical electrospinning. No morphological defects were observed on the nanofiber surface, and a uniform and smooth surface was observed. Figure 3c–h show low and high magnification images of TNF 1, 2, and 3 that used DIPA and a mixture of K-90 and K-30. Each sample was calcined for heating rates of 1 °C/min, 3 °C/min, and 5 °C/min. Compared to the TNF 0 sample, TNF 1, 2, and 3 showed a modified surface; and beads or bubbles were not observed. Depending on the heating rate during the calcination process, byproduct gases of thermal decomposition are emitted at different rates and induce surface deformation. In this experiment, as a chemical blowing agent, DIPA reacted with water to generate a large amount of CO2 and various gases. DIPA contained inside the nanofibers is gasified at high temperatures, and mesopores are formed where the gas was released. Since the speed at which the blowing agent is gasified and released depends on the heating rate, the pore properties can be controlled using this [55,56]. In particular, it was observed that TNF 3 had smaller pores compared to other samples.
More than 300 diameter values were measured using FE-SEM images, and the average diameter graph is shown in Figure 4. The average diameter values of TNF 1, 2, and 3 were verified to be 760 nm, 692 nm, and 507 nm, respectively. The value of TNF 0 was calculated to be 552 nm. Our results verified that the average diameter of calcinated nanofibers also decreased as the heating rate decreased. The results also verified that all of the modified nanofibers compared to TNF 0 had a larger average diameter, and this difference was related to the molecular weight of the polymer and the presence of a blowing agent. At the same time, the unstable nanofiber morphology and fracture of TNF1, 2, and 3 are also explained by the decrease in crystallinity and weakening of mechanical strength due to internal gas emission. Compared to dense and uniform TNF0, TNF1, 2, and 3, which have many defects in microstructure, they have mechanically weak properties [57,58]. The rheological behavior and relaxation of the precursor solution in electrospinning depend on the concentration of the polymer. The viscosity of precursors, including K-90, K-30, and DIPA, controls the surface tension when forming a Taylor cone at the nozzle during electrospinning, resulting in changes in the average diameter. It has been reported that the molecular weight of the polymer or other factors may interact with the diameter of the nanofibers [59,60].
FT-IR measurements were recorded using a tungsten halogen NIR lamp. Figure 5 shows the FT-IR spectroscopy of TiO2 nanofibers at different calcination heating rates. The analysis of chemical bonding and various compositions was performed by FT-IR in 4000 cm−1 and 400 cm−1, as shown in Figure 5. It can be seen that the various vibrational peaks in the observed spectrum are consistent with the reported literature [61,62,63,64]. In relation to the observed spectra of all the samples, the peaks are approximately presented at 3436, 2973, 2345, 1629, and 516 cm−1. The peak at 3432 cm−1 represents the O–H bonding stretching vibration associated with absorbed water. The peak at 2970 cm−1 is attributed to C–H stretching in the polymer (CH2). The peak is also due to the stretching of C=O found in PVP and has a wave number of 1650 cm−1. The vibrational peak at 516 cm−1 was found to be related to the characteristic Ti–O–Ti bonds.
XRD analysis was carried out to verify the crystal structure of the TNF samples and the results are shown in Figure 4. The XRD pattern was obtained from 20° to 80° under conditions of 0.02°/step and 1 step/s, using Co-kα radiation (λ = 1.789010 Å) and a 2theta method.
Figure 6a shows the diffraction pattern of TNF 0, and the (110) and (200) planes of the anatase phase were confirmed at 29.4° and 56.5°, respectively. The peaks at 32.0° and 63.9° were identified as the (100) and (211) planes of rutile, respectively. As a result of Rietveld refinement to determine the ratio of the complex phase, the anatase was 80.8%, and that of rutile was 19.2%. TNF 1, 2, and 3 showed peaks at 29.4°, 44.5°, 56.5°, 63.9°, and 64.0°, which are indexed as (011), (004), (020), (015), and (121) planes, respectively, of a single anatase phase. The color of the as-spun nanofibers was yellow and then, changed to white after calcination. The average crystallite size of each sample was calculated using the Scherrer equation in (1) below.
D = K λ/ β cos θ
where D represents the average crystallite diameter, K = shape factor, λ = wavelength of the wavelength of X-ray used for diffraction (1.789010 Å), β = full width at half maximum (FWHM) of the peak, and θ is the Bragg angle. The average crystallite sizes of TNF 0, TNF 1, TNF 2, and TNF 3 calculated by the Debye-Scherer formula were 31.98 nm, 6.27 nm, 6.67 nm, and 9.27 nm, respectively. According to the calculation, the crystallite size of TNF 1, 2, and 3 decreases as the heating rate increases.
The specific surface area and pore properties were confirmed using a BET method. Adsorption/desorption of N2 gas was carried out at 77.3 K and Figure 7 shows the obtained adsorption/desorption curve for N2. The specific surface area and pore properties calculated using BET theory and the Barrett-Joyner-Halenda (BJH) absorption/desorption method are summarized in Table 2. The obtained curve corresponds to types IV and V, among various types of the International Union of Pure Application Chemistry (IUPAC) classification. The hysteresis originated from the difference in adsorption/desorption partial pressure according to the substance. The shape of the adsorption/desorption curve for relative pressure means that the TiO2 nanofibers had a mesoporous surface with a large specific surface area. The specific surface area of each sample was calculated as 17.2 m2/g, 23.8 m2/g, 27.4 m2/g, and 23.5 m2/g for TNF 0, 1, 2, and 3, respectively. Pore volume means the volume of all the open pores formed on the surface of a material. The BJH desorption average pore width represents the average throat size of pores, and the micropore area represents the pore area per unit mass of the sample. Theoretically, an increase in the specific surface area leads to an increase in the active site and has a proportional relationship with the photocatalytic activity. However, in reality, the photocatalytic activity is rather reduced due to the surface tension of the solution, and the ease of adsorption and desorption of the generated gas, depending on the shape or diameter of the pores. These results show that the specific surface area of nanofibers can be increased by controlling the molecular weight of PVP and adding DIPA. Compared to TNF 0, the mesoporous nanofibers exhibited a specific surface area increased by up to 159%. The average pore widths of TNF 0, 1, 2, and 3 calculated by the BJH method were 13.67 nm, 15.64 nm, 8.7 nm, and 16.2 nm, respectively. TNF 2 was the highest in specific surface area, whereas TNF 3 had better characteristics, such as pore volume, size, and micropore area.
In order to compare the photocatalytic activity, we performed a methylene blue (MB) photodegradation test by TNFs (TiO2 nanofiber samples). Figure 8 shows the absorption spectrum of the MB aqueous solution sampled every 30 min during the photocatalytic reaction. In all of the TNF/MB solution mixtures, the intensity of the absorption peak decreased as time passed, which means that the MB molecules in the aqueous solution were decomposed by TNFs. Among them, the absorbance of TNF 1, 2, and 3 compared to TNF 0 was dramatically decreased, and it can be considered that the mesoporous surface properties positively contribute to the photocatalytic activity.
We obtained normalized values (C/C0) and ln (C0/C) by taking the absorbance at 665 nm to C value, where 665 nm is the lambda max of methylene blue. The C/C0 and ln(C/C0) values as time passed are shown in Figure 9. As seen in Figure 9, TNF 0 showed the lowest photolytic activity and excellent efficiency in the order of TNF 3, 2, and 1. When we considered only the specific surface area, this tendency differed from what we expected; however, it can be explained via pore properties [65]. In this case, since the active sites were relatively small, which can contribute to the photocatalytic reaction due to the surface tension of the aqueous solution and the absorbed byproducts on the surface, the photocatalytic activity decreased. In the case of TNF 1 and 3, the BET data were similar; however, there was a difference in the total pore area and crystallinity. Therefore, the antagonism of these properties in combination can be explained by the fact that the photocatalytic activity was high in the order of TNF 3, 2, 1. After UV irradiation for 3 h, the degradation rates of TNF 0, TNF 1, 2, and 3 were calculated to be 66.7%, 95.5%, 98.4%, and 99.7%, respectively. The aqueous solution sampled using a cuvette is shown in Figure 10. The sample in the cuvette is a solution in which a syringe filter filtered the catalyst, and the color degradation and transmittance can be observed with the naked eye.

4. Conclusions

To improve the photocatalytic activity of TiO2 nanofibers, we prepared modified surfaces using electrospinning and compared the photodegradation efficiency. The precursor solution composition for electrospinning was controlled using the PVP molecular weight and DIPA to obtain the modified nanofibers. An anatase single crystal phase, high specific surface area, and improved photocatalytic activity were achieved in the modified nanofibers. Fabricated nanofibers using DIPA and K-30 showed mesoporous adsorption/desorption curves, and the specific surface area increased from 17.2 m2/g to 27.4 m2/g compared to the ordinary nanofibers.
The photodegradation rate of methylene blue was increased up to 150%, which validates the improvement of the photocatalytic activity of modified nanofibers. Different crystallinity and surface properties were obtained according to the heating rate for the calcination and composition of precursor solutions; in addition, it was necessary to optimize these factors to improve the photocatalytic activity. To contribute to the potential of TiO2 for the field of photocatalyst materials, the mesoporous surface of TiO2 obtained using a simple method such as the one proposed here can be used with other techniques to improve the activity of photocatalysts, such as doping and heterointerface formation.

Author Contributions

Conceptualization, S.-H.Y. and W.-Y.C.; methodology, S.-H.Y. and H.-S.Y.; validation, K.-H.N.; formal analysis, S.-H.Y.; resources, W.-Y.C.; data curation, S.-H.Y. and K.-H.N.; writing—original draft preparation, S.-H.Y.; writing—review and editing, K.-H.N., W.-Y.C. and H.H.; visualization, S.-H.Y. and H.-S.Y.; supervision, W.-Y.C.; project administration, W.-Y.C.; funding acquisition, W.-Y.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work has been supported by the “Basic Science Research Program” through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (MOE) (No. 2022R1F1A1074025) and “Regional Innovation Strategy (RIS)” through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (MOE) (2022RIS-005).

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. Appannagari, R.R. Environmental pollution causes and consequences: A study. North Asian Int. Res. J. Soc. Sci. Humanit. 2017, 3, 151–161. [Google Scholar]
  2. McMichael, A.J.; Campbell-Lendrum, D.; Kovats, S.; Edwards, S.; Wilkinson, P.; Wilson, T.; Nicholls, R.; Hales, S.; Tanser, F.; Le Sueur, D. Global climate change. In Comparative Quantification of Health Risks: Global and Regional Burden of Disease Due to Selected Major Risk Factors; Ezzati, M., Lopez, A., Rodgers, A., Murray, C., Eds.; WHO: Geneva, Switzerland, 2004; pp. 1543–1649. [Google Scholar]
  3. Balat, M. Status of fossil energy resources: A global perspective. Energy Sources Part B Econ. Plan. Policy 2007, 2, 31–47. [Google Scholar] [CrossRef]
  4. Baes, C.; Goeller, H.; Olson, J.; Rotty, R. Carbon Dioxide and Climate: The Uncontrolled Experiment: Possibly severe consequences of growing CO2 release from fossil fuels require a much better understanding of the carbon cycle, climate change, and the resulting impacts on the atmosphere. Am. Sci. 1977, 65, 310–320. [Google Scholar]
  5. Andres, R.J.; Gregg, J.S.; Losey, L.; Marland, G.; Boden, T.A. Monthly, global emissions of carbon dioxide from fossil fuel consumption. Tellus B Chem. Phys. Meteorol. 2011, 63, 309–327. [Google Scholar] [CrossRef] [Green Version]
  6. Kampa, M.; Castanas, E. Human health effects of air pollution. Environ. Pollut. 2008, 151, 362–367. [Google Scholar] [CrossRef]
  7. Mukhopadhyay, K.; Forssell, O. An empirical investigation of air pollution from fossil fuel combustion and its impact on health in India during 1973–1974 to 1996–1997. Ecol. Econ. 2005, 55, 235–250. [Google Scholar] [CrossRef]
  8. Haseena, M.; Malik, M.F.; Javed, A.; Arshad, S.; Asif, N.; Zulfiqar, S.; Hanif, J. Water pollution and human health. Environ. Risk Assess. Remediat. 2017, 1, 2529–8046. [Google Scholar] [CrossRef]
  9. Kishor, R.; Purchase, D.; Saratale, G.D.; Saratale, R.G.; Ferreira, L.F.R.; Bilal, M.; Chandra, R.; Bharagava, R.N. Ecotoxicological and health concerns of persistent coloring pollutants of textile industry wastewater and treatment approaches for environmental safety. J. Environ. Chem. Eng. 2021, 9, 105012. [Google Scholar] [CrossRef]
  10. Rahman, F.A.; Aziz, M.M.A.; Saidur, R.; Bakar, W.A.W.A.; Hainin, M.R.; Putrajaya, R.; Hassan, N.A. Pollution to solution: Capture and sequestration of carbon dioxide (CO2) and its utilization as a renewable energy source for a sustainable future. Renew. Sustain. Energy Rev. 2017, 71, 112–126. [Google Scholar] [CrossRef]
  11. McLarnon, C.R.; Duncan, J.L. Testing of Ammonia Based CO2 Capture with Multi-Pollutant Control Technology. Energy Procedia 2009, 1, 1027–1034. [Google Scholar] [CrossRef] [Green Version]
  12. Boot-Handford, M.E.; Abanades, J.C.; Anthony, E.J.; Blunt, M.J.; Brandani, S.; Mac Dowell, N.; Fernández, J.R.; Ferrari, M.-C.; Gross, R.; Hallett, J.P.; et al. Carbon capture and storage update. Energy Environ. Sci. 2014, 7, 130–189. [Google Scholar] [CrossRef]
  13. Szulczewski, M.L.; MacMinn, C.W.; Herzog, H.J.; Juanes, R. Lifetime of carbon capture and storage as a climate-change mitigation technology. Proc. Natl. Acad. Sci. USA 2012, 109, 5185–5189. [Google Scholar] [CrossRef] [PubMed]
  14. Rafiq, A.; Ikram, M.; Ali, S.; Niaz, F.; Khan, M.; Khan, Q.; Maqbool, M. Photocatalytic degradation of dyes using semiconductor photocatalysts to clean industrial water pollution. J. Ind. Eng. Chem. 2021, 97, 111–128. [Google Scholar] [CrossRef]
  15. Chatterjee, D.; Dasgupta, S. Visible light induced photocatalytic degradation of organic pollutants. J. Photochem. Photobiol. C Photochem. Rev. 2005, 6, 186–205. [Google Scholar] [CrossRef]
  16. Sharma, S.; Dutta, V.; Singh, P.; Raizada, P.; Rahmani-Sani, A.; Hosseini-Bandegharaei, A.; Thakur, V.K. Carbon quantum dot supported semiconductor photocatalysts for efficient degradation of organic pollutants in water: A review. J. Clean. Prod. 2019, 228, 755–769. [Google Scholar] [CrossRef]
  17. Umar, M.; Aziz, H.A. Photocatalytic degradation of organic pollutants in water. Org. Pollut. Monit. Risk Treat. 2013, 8, 196–197. [Google Scholar]
  18. Opoku, F.; Govender, K.K.; van Sittert, C.G.C.E.; Govender, P.P. Recent progress in the development of semiconductor-based photocatalyst materials for applications in photocatalytic water splitting and degradation of pollutants. Adv. Sustain. Syst. 2017, 1, 1700006. [Google Scholar] [CrossRef]
  19. Hariharan, C. Photocatalytic degradation of organic contaminants in water by ZnO nanoparticles: Revisited. Appl. Catal. A Gen. 2006, 304, 55–61. [Google Scholar] [CrossRef]
  20. Moctezuma, E.; Leyva, E.; Aguilar, C.A.; Luna, R.A.; Montalvo, C. Photocatalytic degradation of paracetamol: Intermediates and total reaction mechanism. J. Hazard. Mater. 2012, 243, 130–138. [Google Scholar] [CrossRef]
  21. Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef]
  22. Al-Rasheed, R.A. Water treatment by heterogeneous photocatalysis an overview. In Proceedings of the 4th SWCC Acquired Experience Symposium, Jeddah, Saudi Arabia, 7 May 2005; pp. 1–14. [Google Scholar]
  23. Liu, H.; Wang, C.; Wang, G. Photocatalytic advanced oxidation processes for water treatment: Recent advances and perspective. Chem. Asian J. 2020, 15, 3239–3253. [Google Scholar] [CrossRef] [PubMed]
  24. Hashmi, S.S.; Shah, M.; Muhammad, W.; Ahmad, A.; Ullah, M.A.; Nadeem, M.; Abbasi, B.H. Potentials of phyto-fabricated nanoparticles as ecofriendly agents for photocatalytic degradation of toxic dyes and waste water treatment, risk assessment and probable mechanism. J. Indian Chem. Soc. 2021, 98, 100019. [Google Scholar] [CrossRef]
  25. McIntyre, R.A. Common Nano-Materials and Their Use in Real World Applications. Sci. Prog. 2012, 95, 1–22. [Google Scholar] [CrossRef] [PubMed]
  26. Hamad, S.; Catlow, C.R.A.; Woodley, S.M.; Lago, S.; Mejías, J.A. Structure and Stability of Small TiO2 Nanoparticles. J. Phys. Chem. B 2005, 109, 15741–15748. [Google Scholar] [CrossRef]
  27. Kumar, A.; Jose, R.; Fujihara, K.; Wang, J.; Ramakrishna, S. Structural and Optical Properties of Electrospun TiO2 Nanofibers. Chem. Mater. 2007, 19, 6536–6542. [Google Scholar] [CrossRef]
  28. Sasaki, T.; Nakano, S.; Yamauchi, S.; Watanabe, M. Fabrication of Titanium Dioxide Thin Flakes and Their Porous Aggregate. Chem. Mater. 1997, 9, 602–608. [Google Scholar] [CrossRef]
  29. Greene, L.E.; Law, M.; Yuhas, B.D.; Yang, P. ZnO−TiO2 Core−Shell Nanorod/P3HT Solar Cells. J. Phys. Chem. C 2007, 111, 18451–18456. [Google Scholar] [CrossRef]
  30. Pushpakanth, S.; Srinivasan, B.; Sreedhar, B.; Sastry, T.P. An in situ approach to prepare nanorods of titania–hydroxyapatite (TiO2–HAp) nanocomposite by microwave hydrothermal technique. Mater. Chem. Phys. 2008, 107, 492–498. [Google Scholar] [CrossRef]
  31. Suwarnkar, M.B.; Dhabbe, R.S.; Kadam, A.N.; Garadkar, K.M. Enhanced photocatalytic activity of Ag doped TiO2 nanoparticles synthesized by a microwave assisted method. Ceram. Int. 2014, 40, 5489–5496. [Google Scholar] [CrossRef]
  32. Kavan, L.; Kalbáč, M.; Zukalová, M.; Exnar, I.; Lorenzen, V.; Nesper, R.; Graetzel, M. Lithium Storage in Nanostructured TiO2 Made by Hydrothermal Growth. Chem. Mater. 2004, 16, 477–485. [Google Scholar] [CrossRef]
  33. Li, S.; Zhang, G.; Guo, D.; Yu, L.; Zhang, W. Anodization Fabrication of Highly Ordered TiO2 Nanotubes. J. Phys. Chem. C 2009, 113, 12759–12765. [Google Scholar] [CrossRef]
  34. Mills, A.; Elliott, N.; Parkin, I.P.; O’Neill, S.A.; Clark, R.J. Novel TiO2 CVD films for semiconductor photocatalysis. J. Photochem. Photobiol. A Chem. 2002, 151, 171–179. [Google Scholar] [CrossRef]
  35. Yarin, A.L.; Koombhongse, S.; Reneker, D.H. Taylor cone and jetting from liquid droplets in electrospinning of nanofibers. J. Appl. Phys. 2001, 90, 4836–4846. [Google Scholar] [CrossRef]
  36. Krishnappa, R.V.N. Morphological study of electrospunpolycarbonates as a function of the solventand processing voltage. J. Mater. Sci. 2003, 38, 2357–2365. [Google Scholar] [CrossRef]
  37. Albetran, H.; Dong, Y.; Low, I.M. Characterization and optimization of electrospun TiO2/PVP nanofibers using Taguchi design of experiment method. J. Asian Ceram. Soc. 2015, 3, 292–300. [Google Scholar] [CrossRef] [Green Version]
  38. Motlak, M.; Hamza, A.M.; Hammed, M.G.; Barakat, N.A.M. Cd-doped TiO2 nanofibers as effective working electrode for the dye sensitized solar cells. Mater. Lett. 2019, 246, 206–209. [Google Scholar] [CrossRef]
  39. Song, J.; Guan, R.; Xie, M.; Dong, P.; Yang, X.; Zhang, J. Advances in electrospun TiO2 nanofibers: Design, construction, and applications. Chem. Eng. J. 2022, 431, 134343. [Google Scholar] [CrossRef]
  40. Secundino-Sánchez, O.; Díaz-Reyes, J.; Sánchez-Ramírez, J.F.; Arias-Cerón, J.S.; Galván-Arellano, M.; Vázquez-Cuchillo, O. Controlled synthesis of electrospun TiO2 nanofibers and their photocatalytic application in the decolouration of Remazol Black B azo dye. Catal. Today 2022, 392–393, 13–22. [Google Scholar] [CrossRef]
  41. Hwang, D.; Jo, S.M.; Kim, D.Y.; Armel, V.; MacFarlane, D.R.; Jang, S.-Y. High-Efficiency, Solid-State, Dye-Sensitized Solar Cells Using Hierarchically Structured TiO2 Nanofibers. ACS Appl. Mater. Interfaces 2011, 3, 1521–1527. [Google Scholar] [CrossRef]
  42. Bonincontro, D.; Fraschetti, F.; Squarzoni, C.; Mazzocchetti, L.; Maccaferri, E.; Giorgini, L.; Zucchelli, A.; Gualandi, C.; Focarete, M.L.; Albonetti, S. Pd/Au based catalyst immobilization in polymeric nanofibrous membranes via electrospinning for the selective oxidation of 5-hydroxymethylfurfural. Processes 2020, 8, 45. [Google Scholar] [CrossRef] [Green Version]
  43. Tang, Y.; Zhu, T.; Huang, Z.; Tang, Z.; Feng, L.; Zhang, H.; Li, D.; Xie, Y.; Zhu, C. Preparation of Nanofiber Bundles via Electrospinning Immiscible Polymer Blend for Oil/Water Separation and Air Filtration. Polymers 2022, 14, 4722. [Google Scholar] [CrossRef] [PubMed]
  44. Maccaferri, E.; Mazzocchetti, L.; Benelli, T.; Brugo, T.M.; Zucchelli, A.; Giorgini, L. Self-Assembled NBR/Nomex Nanofibers as Lightweight Rubbery Nonwovens for Hindering Delamination in Epoxy CFRPs. ACS Appl. Mater. Interfaces 2021, 14, 1885–1899. [Google Scholar] [CrossRef] [PubMed]
  45. Liu, D.; Shi, Q.; Jin, S.; Shao, Y.; Huang, J. Self-assembled core-shell structured organic nanofibers fabricated by single-nozzle electrospinning for highly sensitive ammonia sensors. InfoMat 2019, 1, 525–532. [Google Scholar] [CrossRef]
  46. Yang, Y.; Wen, J.; Wei, J.; Xiong, R.; Shi, J.; Pan, C. Polypyrrole-Decorated Ag-TiO2 Nanofibers Exhibiting Enhanced Photocatalytic Activity under Visible-Light Illumination. ACS Appl. Mater. Interfaces 2013, 5, 6201–6207. [Google Scholar] [CrossRef]
  47. Zhang, M.; Shao, C.; Guo, Z.; Zhang, Z.; Mu, J.; Cao, T.; Liu, Y. Hierarchical Nanostructures of Copper(II) Phthalocyanine on Electrospun TiO2 Nanofibers: Controllable Solvothermal-Fabrication and Enhanced Visible Photocatalytic Properties. ACS Appl. Mater. Interfaces 2011, 3, 369–377. [Google Scholar] [CrossRef]
  48. Wang, H.; Lewis, J.P. Effects of dopant states on photoactivity in carbon-doped TiO2. J. Phys. Condens. Matter 2005, 17, L209–L213. [Google Scholar] [CrossRef]
  49. Park, J.-Y.; Lee, J.-H.; Choi, D.-Y.; Hwang, C.-H.; Lee, J.-W. Influence of Fe doping on phase transformation and crystallite growth of electrospun TiO2 nanofibers for photocatalytic reaction. Mater. Lett. 2012, 88, 156–159. [Google Scholar] [CrossRef]
  50. Hamadanian, M.; Reisi-Vanani, A.; Majedi, A. Sol-gel preparation and characterization of Co/TiO2 nanoparticles: Application to the degradation of methyl orange. J. Iran. Chem. Soc. 2010, 7, S52–S58. [Google Scholar] [CrossRef]
  51. Zhao, Y.; Nie, L.; Yang, H.; Song, K.; Hou, H. Tailored fabrication of TiO2/In2O3 hybrid mesoporous nanofibers towards enhanced photocatalytic performance. Colloids Surf. A Physicochem. Eng. Asp. 2021, 629, 127455. [Google Scholar] [CrossRef]
  52. Yu, J.; Yu, H.; Cheng, B.; Zhao, X.; Zhang, Q. Preparation and photocatalytic activity of mesoporous anatase TiO2 nanofibers by a hydrothermal method. J. Photochem. Photobiol. A Chem. 2006, 182, 121–127. [Google Scholar] [CrossRef]
  53. Chun, H.-H.; Jo, W.-K. Polymer material-supported titania nanofibers with different polyvinylpyrrolidone to TiO2 ratios for degradation of vaporous trichloroethylene. J. Ind. Eng. Chem. 2014, 20, 1010–1015. [Google Scholar] [CrossRef]
  54. Nuansing, W.; Ninmuang, S.; Jarernboon, W.; Maensiri, S.; Seraphin, S. Structural characterization and morphology of electrospun TiO2 nanofibers. Mater. Sci. Eng. B 2006, 131, 147–155. [Google Scholar] [CrossRef]
  55. Jia, M.; Guo, S.; Gao, S.; Wang, Q.; Sun, J. Thermal decomposition mechanism of diisopropyl azodicarboxylate and its thermal hazard assessment. Thermochim. Acta 2020, 688, 178601. [Google Scholar] [CrossRef]
  56. Coste, G.; Negrell, C.; Caillol, S. From gas release to foam synthesis, the second breath of blowing agents. Eur. Polym. J. 2020, 140, 110029. [Google Scholar] [CrossRef]
  57. Lim, S.K.; Lee, S.I.; Jang, S.G.; Lee, K.H.; Choi, H.J.; Chin, I.J. Fabrication and physical characterization of biodegradable poly (butylene succinate)/carbon nanofiber nanocomposite foams. J. Macromol. Sci. Part B Phys. 2010, 50, 100–110. [Google Scholar] [CrossRef]
  58. Pan, W.; Ma, Z.; Liu, J.; Liu, Q.; Wang, J. Effect of heating rate on morphology and structure of CoFe2O4 nanofibers. Mater. Lett. 2011, 65, 3269–3271. [Google Scholar] [CrossRef]
  59. Thompson, C.J.; Chase, G.G.; Yarin, A.L.; Reneker, D.H. Effects of parameters on nanofiber diameter determined from electrospinning model. Polymer 2007, 48, 6913–6922. [Google Scholar] [CrossRef]
  60. Dehghan, R.; Barzin, J. Development of a polysulfone membrane with explicit characteristics for separation of low density lipoprotein from blood plasma. Polym. Test. 2020, 85, 106438. [Google Scholar] [CrossRef]
  61. Someswararao, M.V.; Dubey, R.S.; Subbarao, P.S.V.; Singh, S. Electrospinning process parameters dependent investigation of TiO2 nanofibers. Results Phys. 2018, 11, 223–231. [Google Scholar] [CrossRef]
  62. Surovčík, J.; Medvecká, V.; Greguš, J.; Gregor, M.; Roch, T.; Annušová, A.; Ďurina, P.; Vojteková, T. Characterization of TiO2 nanofibers with enhanced photocatalytic properties prepared by plasma assisted calcination. Ceram. Int. 2022, 48, 37322–37332. [Google Scholar] [CrossRef]
  63. Wang, Y.; Yan, L.; He, X.; Li, J.; Wang, D. Controlled fabrication of Ag/TiO2 nanofibers with enhanced stability of photocatalytic activity. J. Mater. Sci. Mater. Electron. 2016, 27, 5190–5196. [Google Scholar] [CrossRef]
  64. Rahma, A.; Munir, M.M.; Prasetyo, A.; Suendo, V.; Rachmawati, H. Intermolecular interactions and the release pattern of electrospun curcumin-polyvinyl (pyrrolidone) fiber. Biol. Pharm. Bull. 2016, 39, 163–173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Yu, J.; Wang, G.; Cheng, B.; Zhou, M. Effects of hydrothermal temperature and time on the photocatalytic activity and microstructures of bimodal mesoporous TiO2 powders. Appl. Catal. B Environ. 2007, 69, 171–180. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of electrospinning process.
Figure 1. Schematic diagram of electrospinning process.
Polymers 15 00134 g001
Figure 2. TGA curve of as-spun TiO2 nanofibers.
Figure 2. TGA curve of as-spun TiO2 nanofibers.
Polymers 15 00134 g002
Figure 3. FE-SEM images of TNFs at low and high magnification: (a,b) TNF 0; (c,d) TNF 1; (e,f) TNF 2; (g,h) TNF 3.
Figure 3. FE-SEM images of TNFs at low and high magnification: (a,b) TNF 0; (c,d) TNF 1; (e,f) TNF 2; (g,h) TNF 3.
Polymers 15 00134 g003
Figure 4. Average diameters of TNF samples with ramping speed.
Figure 4. Average diameters of TNF samples with ramping speed.
Polymers 15 00134 g004
Figure 5. FT-IR spectra curves of electrospun TNF samples.
Figure 5. FT-IR spectra curves of electrospun TNF samples.
Polymers 15 00134 g005
Figure 6. X-ray diffractometer (XRD) spectra pattern of various TiO2 nanofiber samples ((a): TNF 0; (b): TNF 1; (c): TNF 2; (d): TNF 3).
Figure 6. X-ray diffractometer (XRD) spectra pattern of various TiO2 nanofiber samples ((a): TNF 0; (b): TNF 1; (c): TNF 2; (d): TNF 3).
Polymers 15 00134 g006
Figure 7. The isotherm of adsorption/desorption of N2 on the TNF samples ((a): TNF 0; (b): TNF 1; (c): TNF 2; (d): TNF 3).
Figure 7. The isotherm of adsorption/desorption of N2 on the TNF samples ((a): TNF 0; (b): TNF 1; (c): TNF 2; (d): TNF 3).
Polymers 15 00134 g007
Figure 8. UV–Vis absorption spectra of photocatalytic degradation of methylene blue solution by TNFs.
Figure 8. UV–Vis absorption spectra of photocatalytic degradation of methylene blue solution by TNFs.
Polymers 15 00134 g008
Figure 9. Photodegradation efficiency of methylene blue ((a): photodegradation rate of the TNF samples with UV–Vis light irradiation UV irradiation; (b): kinetic linear simulation curves of photodegradation under visible light for TNF samples).
Figure 9. Photodegradation efficiency of methylene blue ((a): photodegradation rate of the TNF samples with UV–Vis light irradiation UV irradiation; (b): kinetic linear simulation curves of photodegradation under visible light for TNF samples).
Polymers 15 00134 g009
Figure 10. Photodegradation results of MB solution as time passed by TNF samples ((a): TNF 0; (b): TNF 1; (c): TNF 2; (d): TNF 3).
Figure 10. Photodegradation results of MB solution as time passed by TNF samples ((a): TNF 0; (b): TNF 1; (c): TNF 2; (d): TNF 3).
Polymers 15 00134 g010
Table 1. Variables sample of precursors’ conditions (K-90, K-30).
Table 1. Variables sample of precursors’ conditions (K-90, K-30).
TNF 0 TNF 1TNF 2TNF 3
K-90: K-30 (g)10 g: 0 g
(11.76wt%)
5 g: 5 g
(10.53wt%)
5 g: 5 g
(10.53wt%)
5 g: 5 g
(10.53wt%)
TTIP (g)15 g
(17.65wt%)
15 g
(15.79wt%)
15 g
(15.79wt%)
15 g
(15.79wt%)
EtOH (g)50 g
(58.82wt%)
50 g
(52.63wt%)
50 g
(52.63wt%)
50 g
(52.63wt%)
ACAC (g)10 g
(11.76wt%)
10 g
(10.53wt%)
10 g
(10.53wt%)
10 g
(10.53wt%)
DIPA (g)0 g
(0wt%)
10 g
(10.53wt%))
10 g
(10.53wt%)
10 g
(10.53wt%)
Heating rate5 °C/min5 °C/min3 °C/min1 °C/min
Table 2. BET test of electrospun TNF samples.
Table 2. BET test of electrospun TNF samples.
Sample
Name
Specific
Surface Area
(m2/g)
Pore Volume
(cm3/g)
BJH Desorption
Average Pore Width
(nm)
Micropore Area
(m2/g)
TNF 017.20.0513.3711.4
TNF 123.80.0915.6414.4
TNF 227.40.068.714.0
TNF 323.50.0916.217.4
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

Yoo, S.-H.; Yoon, H.-S.; Han, H.; Na, K.-H.; Choi, W.-Y. Fabrications of Electrospun Mesoporous TiO2 Nanofibers with Various Amounts of PVP and Photocatalytic Properties on Methylene Blue (MB) Photodegradation. Polymers 2023, 15, 134. https://doi.org/10.3390/polym15010134

AMA Style

Yoo S-H, Yoon H-S, Han H, Na K-H, Choi W-Y. Fabrications of Electrospun Mesoporous TiO2 Nanofibers with Various Amounts of PVP and Photocatalytic Properties on Methylene Blue (MB) Photodegradation. Polymers. 2023; 15(1):134. https://doi.org/10.3390/polym15010134

Chicago/Turabian Style

Yoo, Sun-Ho, Han-Sol Yoon, HyukSu Han, Kyeong-Han Na, and Won-Youl Choi. 2023. "Fabrications of Electrospun Mesoporous TiO2 Nanofibers with Various Amounts of PVP and Photocatalytic Properties on Methylene Blue (MB) Photodegradation" Polymers 15, no. 1: 134. https://doi.org/10.3390/polym15010134

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