Room Temperature Synthesis of Bismuth Oxyiodide with Different Morphologies for the Photocatalytic Degradation of Norfloxacin

Authors

  • Saifullahi Shehu Imam Department of Pure and Industrial Chemistry, Bayero University, P.M.B 3011, Kano, Nigeria
  • Zakariyya Uba Zango Department of Chemistry, Al-Qalam University Katsina, P.M.B 2341, Katsina, Nigeria
  • Haruna Abdullahi Department of Applied Chemistry, Federal University Dutsinma, P.M.B 5001, Katsina, Nigeria

Keywords:

Photocatalytic degradation, room-temperature synthesis, BiOI spheres, BiOI Plates, norfloxacin.

Abstract

In this study, both BiOI spheres and BiOI plates were synthesized successfully at room temperature and applied for the photocatalytic degradation of norfloxacin using indoor fluorescent light illumination. The samples were characterized by X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), UV-Vis absorption spectrum, scanning electron microscope (SEM) and Brunauer-Emmett-Teller (BET). Although both BiOI spheres and BiOI plates displayed approximately similar absorption band edge, higher photocatalytic degradation was noticed in the case of BiOI spheres as compared with BiOI plates. This was attributed to the unique features of BiOI spheres, such as enlarged specific surface area and enhanced adsorption capacity. The study demonstrated that morphology has a key role in improving the degradation efficiency of a photocatalyst.

References

[1]. Simoens, S., et al., Consumption patterns and in vitro resistance of Streptococcus pneumoniae to fluoroquinolones. Antimicrobial agents and chemotherapy, 2011. 55(6): p. 3051-3053.
[2]. Hirsch, R., et al., Occurrence of antibiotics in the aquatic environment. Science of the Total Environment, 1999. 225(1): p. 109-118.
[3]. Babi?, S., M. Periša, and I. Škori?, Photolytic degradation of norfloxacin, enrofloxacin and ciprofloxacin in various aqueous media. Chemosphere, 2013. 91(11): p. 1635-1642.
[4]. An, T., et al., Mechanistic considerations for the advanced oxidation treatment of fluoroquinolone pharmaceutical compounds using TiO2 heterogeneous catalysis. The Journal of Physical Chemistry A, 2010. 114(7): p. 2569-2575.
[5]. Rigos, G., et al., Potential drug (oxytetracycline and oxolinic acid) pollution from Mediterranean sparid fish farms. Aquatic Toxicology, 2004. 69(3): p. 281-288.
[6]. Prutthiwanasan, B., C. Phechkrajang, and L. Suntornsuk, Fluorescent labelling of ciprofloxacin and norfloxacin and its application for residues analysis in surface water. Talanta, 2016. 159: p. 74-79.
[7]. Zorita, S., L. Mårtensson, and L. Mathiasson, Occurrence and removal of pharmaceuticals in a municipal sewage treatment system in the south of Sweden. Science of the total environment, 2009. 407(8): p. 2760-2770.
[8]. Ferreira, V.R., et al., Fluoroquinolones biosorption onto microbial biomass: activated sludge and aerobic granular sludge. International Biodeterioration & Biodegradation, 2016. 110: p. 53-60.
[9]. Ebrahimpour, B., Y. Yamini, and M. Moradi, Application of ionic surfactant as a carrier and emulsifier agent for the microextraction of fluoroquinolones. Journal of pharmaceutical and biomedical analysis, 2012. 66: p. 264-270.
[10]. Zhu, L., et al., Electrochemical oxidation of fluoroquinolone antibiotics: Mechanism, residual antibacterial activity and toxicity change. Water research, 2016. 102: p. 52-62.
[11]. Martín, M.B., et al., Degradation of alachlor and pyrimethanil by combined photo-Fenton and biological oxidation. Journal of hazardous materials, 2008. 155(1): p. 342-349.
[12]. Murthy, H.R. and H. Manonmani, Aerobic degradation of technical hexachlorocyclohexane by a defined microbial consortium. Journal of hazardous materials, 2007. 149(1): p. 18-25.
[13]. Ahmad, A., L. Tan, and S.A. Shukor, Dimethoate and atrazine retention from aqueous solution by nanofiltration membranes. Journal of hazardous materials, 2008. 151(1): p. 71-77.
[14]. Maldonado, M., et al., Partial degradation of five pesticides and an industrial pollutant by ozonation in a pilot-plant scale reactor. Journal of hazardous materials, 2006. 138(2): p. 363-369.
[15]. Saritha, P., et al., Comparison of various advanced oxidation processes for the degradation of 4-chloro-2 nitrophenol. Journal of Hazardous Materials, 2007. 149(3): p. 609-614.
[16]. Kato, S., et al., Photocatalytic degradation of gaseous sulfur compounds by silver-deposited titanium dioxide. Applied Catalysis B: Environmental, 2005. 57(2): p. 109-115.
[17]. Xu, Y., Comparative studies of the Fe 3+/2+–UV, H 2 O 2–UV, TiO 2–UV/vis systems for the decolorization of a textile dye X-3B in water. Chemosphere, 2001. 43(8): p. 1103-1107.
[18]. Tokunaga, S., H. Kato, and A. Kudo, Selective preparation of monoclinic and tetragonal BiVO4 with scheelite structure and their photocatalytic properties. Chemistry of Materials, 2001. 13(12): p. 4624-4628.
[19]. Wu, T., et al., Bi 2 S 3 nanostructures: a new photocatalyst. Nano Research, 2010. 3(5): p. 379-386.
[20]. Saison, T., et al., New insights into Bi2WO6 properties as a visible-light photocatalyst. The Journal of Physical Chemistry C, 2013. 117(44): p. 22656-22666.
[21]. Zhang, L., et al., Sonochemical synthesis of nanocrystallite Bi 2 O 3 as a visible-light-driven photocatalyst. Applied Catalysis A: General, 2006. 308: p. 105-110.
[22]. Yao, W.F., et al., Synthesis and photocatalytic property of bismuth titanate Bi 4 Ti 3 O 12. Materials Letters, 2003. 57(13): p. 1899-1902.
[23]. Shimodaira, Y., et al., Photophysical properties and photocatalytic activities of bismuth molybdates under visible light irradiation. The Journal of Physical Chemistry B, 2006. 110(36): p. 17790-17797.
[24]. Cao, X.-F., et al., Persimmon-like (BiO) 2 CO 3 microstructures: hydrothermal preparation, photocatalytic properties and their conversion into Bi 2 S 3. CrystEngComm, 2011. 13(6): p. 1939-1945.
[25]. Chang, X., et al., BiOX (X= Cl, Br, I) photocatalysts prepared using NaBiO 3 as the Bi source: characterization and catalytic performance. Catalysis Communications, 2010. 11(5): p. 460-464.
[26]. Shi, X., et al., Solvothermal synthesis of BiOI hierarchical spheres with homogeneous sizes and their high photocatalytic performance. Materials Letters, 2012. 68: p. 296-299.
[27]. Cheng, H., B. Huang, and Y. Dai, Engineering BiOX (X= Cl, Br, I) nanostructures for highly efficient photocatalytic applications. Nanoscale, 2014. 6(4): p. 2009-2026.
[28]. Cao, S., P. Zhou, and J. Yu, Recent advances in visible light Bi-based photocatalysts. Chinese Journal of Catalysis, 2014. 35(7): p. 989-1007.
[29]. He, R., et al., 3D BiOI–GO composite with enhanced photocatalytic performance for phenol degradation under visible-light. Ceramics International, 2015. 41(3): p. 3511-3517.
[30]. Han, S., et al., Fabrication of a ?-Bi 2 O 3/BiOI heterojunction and its efficient photocatalysis for organic dye removal. Chinese Journal of Catalysis, 2015. 36(12): p. 2119-2126.
[31]. Li, H., et al., Photocatalytic properties of BiOI synthesized by a simple hydrothermal process. Materials Letters, 2013. 107: p. 262-264.
[32]. Ye, L., et al., BiOI thin film via chemical vapor transport: photocatalytic activity, durability, selectivity and mechanism. Applied Catalysis B: Environmental, 2013. 130: p. 1-7.
[33]. Cao, Y., et al., Rapid one-step room-temperature solid-state synthesis and formation mechanism of ZnO nanorods as H 2 S-sensing materials. Solid-State Electronics, 2013. 82: p. 67-71.
[34]. Xie, J., et al., Room-temperature solid-state synthesis of BiOCl hierarchical microspheres with nanoplates. Catalysis Communications, 2015. 69: p. 34-38.
[35]. He, R., et al., Room-temperature synthesis of BiOI with tailorable (001) facets and enhanced photocatalytic activity. Journal of colloid and interface science, 2016. 478: p. 201-208.
[36]. Gao, F., et al., Chemically bonded graphene/BiOCl nanocomposites as high-performance photocatalysts. Physical Chemistry Chemical Physics, 2012. 14(30): p. 10572-10578.
[37]. Hu, Y.H., A highly efficient photocatalyst—hydrogenated black TiO2 for the photocatalytic splitting of water. Angewandte Chemie International Edition, 2012. 51(50): p. 12410-12412.
[38]. Song, J.-M., et al., Hierarchical structured bismuth oxychlorides: self-assembly from nanoplates to nanoflowers via a solvothermal route and their photocatalytic properties. CrystEngComm, 2010. 12(11): p. 3875-3881.
[39]. Jamil, T.S., E.S. Mansor, and R.A. Nasr, Degradation of Lindane using two nanosized BiOXs and their heterojunction under visible light. Desalination and Water Treatment, 2016. 57(31): p. 14750-14761.
[40]. Xia, J., et al., Improved photocatalytic activity of few-layer Bi 4 O 5 I 2 nanosheets induced by efficient charge separation and lower valence position. Journal of Alloys and Compounds, 2017. 695: p. 922-930.
[41]. Wang, J.-j., et al., Effect of bismuth tungstate with different hierarchical architectures on photocatalytic degradation of norfloxacin under visible light. Transactions of Nonferrous Metals Society of China, 2017. 27(8): p. 1794-1803.
[42]. Li, J., et al., Fabrication of FeVO 4/Fe 2 TiO 5 composite catalyst and photocatalytic removal of norfloxacin. Chemical Engineering Journal, 2016. 298: p. 300-308.
[43]. Sun, J., et al., Photocatalytic degradation of Orange G on nitrogen-doped TiO 2 catalysts under visible light and sunlight irradiation. Journal of hazardous materials, 2008. 155(1): p. 312-319.
[44]. Bahnemann, W., M. Muneer, and M. Haque, Titanium dioxide-mediated photocatalysed degradation of few selected organic pollutants in aqueous suspensions. Catalysis Today, 2007. 124(3): p. 133-148.

Downloads

Published

2018-03-11

How to Cite

Shehu Imam, S., Uba Zango, Z., & Abdullahi, H. (2018). Room Temperature Synthesis of Bismuth Oxyiodide with Different Morphologies for the Photocatalytic Degradation of Norfloxacin. American Scientific Research Journal for Engineering, Technology, and Sciences, 41(1), 26–39. Retrieved from https://asrjetsjournal.org/index.php/American_Scientific_Journal/article/view/3944

Issue

Section

Articles