A simple UV-ozone surface treatment to enhance photocatalytic performance of TiO2 loaded polymer nanofiber membranes

S. Dilpazira, M. Usmana, S. Rasulb and S. N. Arshad*a
aDepartment of Chemistry, Syed Babar Ali School of Science & Engineering, Lahore University of Management Sciences, Lahore 54792, Pakistan. E-mail: salman.arshad@lums.edu.pk; Tel: +92-42-35608478
bKing Abdullah University of Science and Technology (KAUST), Laboratory for Carbon Nanostructures, Thuwal, 23955-6900, Saudi Arabia

Received 31st October 2015 , Accepted 12th January 2016

First published on 29th January 2016


Abstract

Homogeneously dispersed titanium dioxide loaded polyacrylonitrile nanofiber membranes with increased active mass loading, Ti3+ surface defects and hydrophilicity were fabricated by combining electrospinning and UV-ozone surface treatment. The photocatalytic activity improved by a factor of ∼2 and the kinetics of photodegradation switched from pseudo-first order to pseudo-second order with increasing TiO2 content with a maximum rate constant of 20.7 h−1.


Introduction

TiO2 and ZnO are very promising photocatalysts for utilizing solar energy for degradation of organic dyes and bacteria from wastewater.1–3 These photocatalysts absorb the right energy photons, generate excitons and cause the oxidation of the biomaterial that accumulates on its surface.4 The charge carrier generation, transfer and recombination affect the overall photocatalytic activity and need to be understood individually for optimum performance.5

Nanoparticles of these photocatalysts can provide a large number of active sites per unit area but their natural tendency to agglomerate restricts their efficiency. Immobilization of these nanoparticles on a support with homogeneous dispersion is an effective method to increase their photocatalytic performance.6 One such support is nanofiber membranes by electrospinning which have several attributes suitable for water filtration and purification such as interconnected open pore sizes in nm to μm range, high permeability, and large surface to volume ratio allowing greater adsorption of the pollutants.7,8 However, bio-fouling of these membranes by bacteria and viruses is a major limiting factor which determines the efficiency and life of the filter.9 Recently, the focus has been on synthesizing self-cleaning filters by immobilizing photocatalyst nanoparticles on selective polymer10–13 and carbon14–16 nanofibers. The electrospun nanofiber mesh will carry out the filtering part, isolating bacteria and most viruses, whereas the photocatalyst nanoparticles will oxidize them to CO2 and H2O. The need for a safe, reliable and low cost method for efficient removal of organic pollutants can potentially be addressed by such multifunctional nanofiber membranes.

Ding et al.10 made polyimide/ZnO fibers by electrospinning using a direct ion-exchange process which exhibited pseudo-first order photodegradation of methylene blue (MB) with rate constants in 0.6–1.7 h−1 range. Ramasundaram et al.13 used a combination of electrospinning of polyvinylidene fluoride (PVDF) and electrospraying of TiO2/DMF solution to fabricate PVDF/TiO2 hybrid nanofibers in which TiO2 nanoparticles were non-uniformly agglomerated on the surface. Other approaches include composite TiO2 nanofibers17 but their fragile nature is not suitable for pressure driven applications. Recently, a promising air jet spinning technique has been used to fabricate TiO2/polymer composite membranes18 with mean fiber diameters and pore sizes of 505–901 nm and 1.58–5.12 μm, respectively, but the process needs optimization to achieve uniform TiO2 dispersion. Moreover, surface defects such as Ti3+ on TiO2 and oxygen vacancies play a vital role.19 A surface plasma treatment on TiO2 resulted in enhanced efficiency in dye sensitized solar cells20,21 due to efficient charge transport. Besides these TiO2 based composites, various 1D spatially ordered architectures22 and graphene based composite photocatalysts23 are being intensively investigated for solar every conversion applications.

Enhanced photocatalytic performance of these membranes require optimum morphology, distribution, surface chemical states of TiO2 and improved hydrophilicity. Here, we report a simple and low cost UV-ozone surface treatment to enhance these factors and obtain improved photocatalytic performance of TiO2/polymer composite nanofiber membranes. These offer great promise as anti-fouling filters for water remediation.

Experimental

Polyacrylonitrile (PAN, Sigma-Aldrich, Mw = 150[thin space (1/6-em)]000) was stirred for 12 h in N,N-dimethyl formamide (DMF) at room temperature to form a 9 wt% solution. Separately, TiO2 agglomerates (P25, Sigma-Aldrich, particle size 21 nm) were ultrasonicated for 4 hours in DMF to homogeneously disperse them. The two solutions were then blended using ultrasonication for additional 4 hours. Electrospinning at 15 kV with 15 cm needle-to-collector distance yielded nanofiber membranes. The composition of the blend solutions were adjusted to obtain 9–47 wt% TiO2/PAN nanofibers as confirmed by thermal gravimetric analysis (TGA). In this paper, the samples will be identified by their wt% TiO2. The nanofiber membranes were then exposed to UV-ozone for 15 minutes. Nanofiber sizes, surface morphology and TiO2 dispersion were characterized by scanning electron microscopy (SEM) and transmission electron microscopy (TEM) equipped with energy-dispersive X-ray spectroscopy (EDS). The oxidation states of Ti, before and after UV-ozone treatment, were characterized by studying the L2,3 energy-loss near edge structure (ELNES) by electron energy loss spectroscopy (EELS). The membrane hydrophilicity was characterized by water contact angle measurement. Photocatalytic performance of the membranes were evaluated by photodegradation of 10 ppm MB solution by 10 mg of composite nanofiber membranes under UV irradiation and monitored by UV-Vis spectrophotometry. A calibration curve of 0–20 ppm MB solution by serial dilutions was used as a reference. Adsorption of MB on nanofiber surfaces were taken into account prior to all photodegradation experiments by establishing an adsorption–desorption equilibrium. 17 and 47 wt% samples, without UV-ozone treatment, were also tested for comparison.

Results and discussion

PAN was selected because it can be easily electrospun with very smooth surfaces and minimum bead formation. Moreover, it's a widely used precursor for carbon nanofibers (CNFs) thus expanding the scope of this study to making TiO2/CNFs composites with enhanced photocatalytic performance.23 In fact, TiO2–carbon composites have shown similar results to TiO2–graphene composites on adsorption of pollutants, light absorption intensity and charge carrier lifetimes.23 Fig. 1a is a representative SEM image showing the interpenetrating random network of 25 wt% sample with interstices on the order of few μm which can trap most of the microorganisms. Fig. 1b shows the diameter distribution of all the samples which is characteristically log-normal. The mean diameter increases from 225 to 335 nm with increasing TiO2 due to increased blend solution viscosity and hindrance in nanofiber thinning during later stages of electrospinning. Fig. 1c and d shows SEM images of the 47 wt% nanofibers, before and after a 15 min UV-ozone treatment, respectively, which clearly indicates the enhanced exposure of TiO2 in the UV-ozone treated sample. UV-ozone combines decomposition of organic materials by UV radiations and strong oxidation by atomic oxygen during formation/decomposition of ozone. Here, we used it to selectively etch PAN and expose TiO2 on the nanofiber surface. The TiO2 active mass loading can be adjusted by varying the TiO2 loading in the blend solution as shown in Fig. 1e and f for 40 and 25 wt% samples, respectively. A PAN/TiO2 blend solution by stirring only, without ultrasonication, results in agglomerated TiO2 non-uniformly dispersed in PAN nanofibers as shown in Fig. S1 of the ESI. Thus, ultrasonication of the blend solution gives excellent dispersion of TiO2 in PAN and a post UV-ozone treatment enhances the exposure of TiO2 as depicted in a high resolution SEM image in the graphical abstract.
image file: c5ra22903k-f1.tif
Fig. 1 (a) A representative SEM image of the 25 wt% TiO2/PAN nanofiber membrane (inset: larger magnification), (b) nanofiber diameter distribution of all the samples which is characteristically log-normal, (c) a 47 wt% nanofiber before UV-ozone treatment and (d) 47 wt% after UV-ozone treatment, (e) 40 wt%, (f) 25 wt% nanofibers after UV-ozone treatment.

The amount of TiO2 loading in PAN nanofibers was estimated by TGA as the residual weight after complete burning of PAN. Fig. 2 shows the simultaneous DSC and TGA scans of all the samples. The first weight loss occurs at ∼300 °C which coincides with the exothermal DSC peak of PAN stabilization reaction.24 The samples remain relatively stable between 300 and 450 °C. After that, the thermal degradation of PAN occurs as shown by a second exothermic peak. This exothermic event shifts to lower temperatures (by as much as ∼70 °C for 47 wt% sample) as the amount of TiO2 increases due to the effect of large thermal mass of TiO2 which aids in the early decomposition.


image file: c5ra22903k-f2.tif
Fig. 2 DSC (top) and TGA (bottom) scans of TiO2/PAN nanofibers. The weight percent of TiO2 in the samples is calculated from TGA as the residual weight at 800 °C after complete thermal degradation of PAN in air.

To verify the morphology and distribution of nanosized TiO2 on PAN nanofibers, TEM analysis was carried out. Fig. 3a and b shows the bright field TEM images and EDS elemental maps of Ti (green) and O (red) for a 40 wt% nanofiber before and after UV-ozone treatment, respectively. The TiO2 nanoparticles are finely distributed in PAN nanofibers and are ∼20 nm in size. Moreover, the active particles of TiO2 are not only located at the surface but are very well embedded in nanofiber bulk as well.


image file: c5ra22903k-f3.tif
Fig. 3 Bright field TEM image, Ti and O elemental EDS map and water contact angle of 40 wt% nanofiber sample (a) before and (b) after UV-ozone treatment. (c) EELS edge spectra showing Ti L-edges (L3 and L2) and O K-edge before and after UV-ozone treatment (inset: zoomed-in of L3 and L2 edges).

Several factors affect the photocatalytic activity such as surface chemical states, electron–hole recombination, charge separation and transfer kinetics. Moreover, the dye adsorption is affected by the hydrophilicity of the nanofiber membranes. These factors are now discussed in context of the UV-ozone treatment. It's reported that radio frequency (RF)20 and oxygen21 plasma treatment on TiO2 increases hydrophilicity and Ti3+ surface sites resulting in reduced electron–hole recombination rate. Here, the effect of UV-ozone treatment on the chemical states of TiO2 nanoparticles is characterized by EELS by studying the L2,3 energy-loss near edge structure (ELNES) in TiO2 and comparing the L3/L2 intensity ratio with the expected value of 0.8.25 L2,3 near edge fine structure correspond to transitions to unoccupied 3d states and is an indicator of the oxidation state (higher L3/L2 ratio corresponds to higher concentration of Ti3+).26 The ratio was estimated by using the integrated area above the step level and under each intensity peak (Fig. 3c). The step level is extrapolated from the tail ends of each peak. This ratio increased from 0.76 (untreated) to 0.82 (UV-ozone treated) indicating an increase in the Ti3+ oxidation state.27 Thus, UV-ozone treatment results in a surface reaction of TiO2 (Ti4+) to Ti2O3 (Ti3+). This generates oxygen vacancies and excess electrons in Ti at surface which aids the charge transport between dye and photocatalyst and reduces electron–hole recombination rate.27 Kuznetsov et al. reported a ∼14% conversion of Ti4+ to Ti3+ under UV irradiation.28 Similar results have been reported for dye sensitized solar cells.20,21 Moreover, electrochemical impedance spectroscopy showed that charge transfer resistance is reduced in UV-ozone treated TiO2.21 To determine the change in the hydrophilic nature of we have measured the water contact angle of the nanofiber membrane. The water contact angle of the nanofiber membrane (insets in Fig. 3a and b) decreased from 29.6° to 18.6° after UV-ozone treatment resulting in increased hydrophilicity which aids in the adsorption of dye.

Finally, the photocatalytic performance of the TiO2/PAN composite nanofiber membranes were evaluated by photodegradation of MB under UV irradiation and monitored by determining the decrease in absorbance by UV-Vis spectrophotometry. However, the decrease in absorbance is also due to adsorption of MB molecules on PAN and TiO2.29 In this study, the mixture of TiO2/PAN nanofibers in MB solution were shaken for 2 hours in dark until an adsorption–desorption equilibrium was attained as confirmed by a steady state absorbance value of the mixture. Xu et al.29 reported 30 min to achieve adsorption–desorption equilibrium for TiO2 nanoparticles in MB solution. Fig. S2 of the ESI shows the UV-Vis spectrums of MB solutions taken at different times of photodegradation under UV irradiation. There is negligible change in the absorbance of the control sample. The photocatalytic efficiency of nanofiber membranes is quantitatively shown in Fig. 4a as C/Co versus time plot where Co is the MB solution concentration after the adsorption–desorption equilibrium has been achieved. The initial rate of photodegradation increases with TiO2 loading and UV-ozone treatment further enhances it due to the increased exposure of TiO2, increase Ti3+ surface states, and increase hydrophilicity. The total degradation after 60 min of exposure is 92, 93, 95, 97 and 98% for 9, 17, 25, 40, and 47 wt% TiO2/PAN samples, respectively. Fig. 4a also shows the performance of as-prepared 17 and 47 wt% samples (without UV-ozone treatment) which takes about twice as much time to photodegrade.


image file: c5ra22903k-f4.tif
Fig. 4 (a) Photocatalytic performance of TiO2/PAN nanofiber membranes plotted as C/Co versus time, (b) ln(C) versus t plot for 9, 17, and 25 wt% and (inset) 1/C versus t plot for 40 and 47 wt% membranes, and (c) photocatalytic performance of 45 wt% membrane over 6 cycles.

It is relevant to remark that the UV-ozone treatment didn't cause any significant change in the mechanical strength of the nanofiber membranes as they all were handled easily without any damage. In fact, TiO2 will act as reinforcement in the flexible PAN matrix to make a strong and tough multifunctional composite membrane which can withstand the swelling pressures generated during water uptake.30 Fig. 4b, which plots ln(C) versus t, shows that the photodegradation follows a pseudo-first order kinetics for 9, 17, and 25 wt% samples with rate constants of 3.9, 4.2, 4.7 h−1, respectively. Interestingly, the kinetics switches to pseudo-second order for 40 and 47 wt% samples with rate constants of 18.8 and 20.7 h−1, respectively, as shown in the plot of 1/C versus t in the inset of Fig. 4b. Xu et al.29 reported a detailed study on photodegradation of MB by TiO2 powder and observed that the concentration of TiO2 and MB has a complex impact on the overall kinetics with both desorption and degradation exhibiting pseudo-first order kinetics. He observed a desorption dominated slight increase (less than 10%) in the absorbance of 20 ppm MB solution within first 5 min of UV irradiation attributed to a shift in adsorption equilibrium due to altered electric charges on MB functional group and TiO2 surface. Thus, our results from a much longer time scale are expected to be negligibly affected by this initial desorption phenomenon. Strong adhesion of TiO2 to nanofiber membrane is critical for long term and multiple cycles use. The 40 wt% membrane showed a consistent performance over 6 cycles of photodegradation as shown in Fig. 4c. This suggests a strong physical interlocking of TiO2 on the nanofiber surface by the encapsulating PAN without using any interfacial chemistry based methods to enhance the adhesion. No leaching of TiO2 in the MB solution was observed. This also implies Ti3+ concentration doesn't change over multiple cycles of use.

Conclusions

The simple process of electrospinning from an ultrasonicated PAN/TiO2 blend solution followed by controlled UV-ozone treatment results in flexible and strong composite nanofiber membranes with excellent photocatalytic activity and reliability over multiple cycles. This enhanced photocatalytic activity is attributed to the enhanced active mass exposure, increased Ti3+ surface defects and increased hydrophilicity. At higher TiO2 concentration the photodegradation kinetics switches to pseudo-second order.

References

  1. X. Wang, J. C. Yu, C. Ho, Y. Hou and X. Fu, Langmuir, 2005, 21, 2552–2559 CrossRef CAS PubMed .
  2. H. Lachheb, E. Puzenat, A. Houas, M. Ksibi, E. Elaloui, C. Guillard and J. Herrmann, Appl. Catal., B, 2002, 39, 75–90 CrossRef CAS .
  3. J. Yu and X. Yu, Environ. Sci. Technol., 2008, 42, 4902–4907 CrossRef CAS PubMed .
  4. K. Nakata, T. Ochiai, T. Murakami and A. Fujishima, Electrochim. Acta, 2012, 84, 103–111 CrossRef CAS .
  5. J. Schneider, M. Matsuoka, M. Takeuchi, J. Zhang, Y. Horiuchi, M. Anpo and D. W. Bahnemann, Chem. Rev., 2014, 114, 9919–9986 CrossRef CAS PubMed .
  6. S. Leong, A. Razmjou, K. Wang, K. Hapgood, X. Zhang and H. Wang, J. Membr. Sci., 2014, 472, 167–184 CrossRef CAS .
  7. V. Thavasi, G. Singh and S. Ramakrishna, Energy Environ. Sci., 2008, 1, 205–221 CAS .
  8. A. Sato, R. Wang, H. Ma, B. S. Hsiao and B. Chu, J. Electron Microsc., 2011, 60, 201–209 CrossRef CAS PubMed .
  9. J. Prince, S. Bhuvana, V. Anbharasi, N. Ayyanar, K. Boodhoo and G. Singh, Sci. Rep., 2014, 4, 6555 CrossRef CAS PubMed .
  10. Q. Ding, Y.-E. Miao and T. Liu, ACS Appl. Mater. Interfaces, 2013, 5, 5617–5622 CAS .
  11. T. He, Z. Zhou, W. Xu, F. Ren, H. Ma and J. Wang, Polymer, 2009, 50, 3031–3036 CrossRef CAS .
  12. N. Daels, M. Radoicic, M. Radetic, S. W. H. Van Hulle and K. De Clerck, Sep. Purif. Technol., 2014, 133, 282–290 CrossRef CAS .
  13. S. Ramasundaram, A. Son, M. G. Seid, S. Shim, S. H. Lee, Y. C. Chung, C. Lee, J. Lee and S. W. Hong, J. Hazard. Mater., 2015, 285, 267–276 CrossRef CAS PubMed .
  14. S.-I. Lee, S.-M. Jo, H.-I. Joh, M.-H. Lee and S. Lee, Chem. Commun., 2015, 51, 2718–2720 RSC .
  15. C. H. Kim, B. H. Kim and K. S. Yang, Carbon, 2012, 50, 2472–2481 CrossRef CAS .
  16. B. Pant, N. A. M. Barakat, H. R. Pant, M. Park, P. S. Saud, J.-W. Kim and H.-Y. Kim, J. Colloid Interface Sci., 2014, 434, 159–166 CrossRef CAS PubMed .
  17. M. Zhang, L. Li and X. Zhang, RSC Adv., 2015, 5, 29693–29697 RSC .
  18. A. Abdal-hay, A. S. Hamdy Makhlouf and K. A. Khalil, ACS Appl. Mater. Interfaces, 2015, 7, 13329–13341 CAS .
  19. B. Ohtani, Catalysts, 2013, 3, 942–953 CrossRef CAS .
  20. K.-H. Park and M. Dhayal, Electrochem. Commun., 2009, 11, 75–79 CrossRef CAS .
  21. R. Liu, W.-D. Yang and L.-S. Qiang, J. Power Sources, 2012, 199, 418–425 CrossRef CAS .
  22. S. Liu, Z.-R. Tang, Y. Sun, J. C. Colmenares and Y.-J. Xu, Chem. Soc. Rev., 2015, 44, 5053–5075 RSC .
  23. Y. Zhang, Z.-R. Tang, X. Fu and Y.-J. Xu, ACS Nano, 2010, 4, 7303–7314 CrossRef CAS PubMed .
  24. Z. Wangxi, L. Jie and W. Gang, Carbon, 2003, 41, 2805–2812 CrossRef .
  25. R. D. Leapman, L. A. Grunes and P. L. Fejes, Phys. Rev. B: Condens. Matter Mater. Phys., 1982, 26, 614–635 CrossRef CAS .
  26. E. Stoyanov, F. Langenhorst and G. Steinle-Neumann, Am. Mineral., 2007, 92, 577–586 CrossRef CAS .
  27. J. Jun, M. Dhayal, J. H. Shin, J. C. Kim and N. Getoff, Radiat. Phys. Chem., 2006, 75, 583–589 CrossRef CAS .
  28. A. I. Kuznetsov, O. Kameneva, A. Alexandrov, N. Bityurin, P. Marteau, K. Chhor, C. Sanchez and A. Kanaev, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2005, 71, 1–7 CrossRef PubMed .
  29. C. Xu, G. P. Rangaiah and X. S. Zhao, Ind. Eng. Chem. Res., 2014, 53, 14641–14649 CrossRef CAS .
  30. M. Satterfield, P. Majsztrik, H. Ota, J. Benziger and A. Bocarsly, J. Polym. Sci., Part B: Polym. Phys., 2006, 44, 2327–2345 CrossRef CAS .

Footnote

Electronic supplementary information (ESI) available: SEM images of the nanofibers without ultrasonication of the TiO2/PAN blend solution. UV-Vis spectrums of the MB solution after various times of photodegradation. See DOI: 10.1039/c5ra22903k

This journal is © The Royal Society of Chemistry 2016