Blog List

Tuesday 25 July 2017

Efficiency of Polymeric Membrane Graphene Oxide-TiO2 for Removal of Azo Dye

Published Date
Journal of Chemistry
Volume 2017 (2017), Article ID 6217987, 13 pages
https://doi.org/10.1155/2017/6217987
Author
1Department of Environmental Engineering, Faculty of Environment and Energy, Islamic Azad University, Science and Research Branch, Tehran, Iran
2Department of Environmental Health Engineering, School of Public Health, Iran University of Medical Sciences, P.O. Box 15875-4199, Tehran, Iran
3Chemistry Department, Islamic Azad University, Central Tehran Branch, Tehran, Iran
4Nanotechnology Research Institute, Chemical Engineering Department, Babol University of Technology, Babol, Iran
Correspondence should be addressed to Roshanak Rezaei Kalantary
Received 22 April 2016; Revised 11 August 2016; Accepted 14 September 2016; Published 27 February 2017
Academic Editor: Zuhui Zhang
Copyright © 2017 Elahe Dadvar et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

Abstract

Achieving the desired standard of drinking water quality has been one of the concerns across water treatment plants in the developing countries. Processes such as grid chamber, coagulation, sedimentation, clarification, filtration, and disinfection are typically used in water purification plants. Among these methods, unit filtration which employs polymers is one of the new technologies. There have been many studies about the use of semiconductive TiO2 with graphene oxide (GO) on the base of different polymeric membranes for the removal of azo dyes, especially methylene blue (MB). Polymeric GO-TiO2 membranes have high photocatalytic, antifouling property and permeate the flux removal of organic pollutants. The aim of this study was to investigate the characteristics of different polymeric membranes such as anionic perfluorinated polymer (Nafion), cellulose acetate, polycarbonate (PC), polysulfone fluoride (PSF), and polyvinylidene fluoride (PVDF). The result of this study showed that the GO-TiO2 membrane can be used in the field of water treatment and will be used for the removal of polycyclic aromatic hydrocarbons (PAHs) from wastewater.

1. Introduction

With the development of heavy industries in recent decades, obtaining healthy drinking water has become a global concern [1]. Organic substances, industrial dyes, microorganisms, and heavy metals (heavy metals include cobalt, cadmium, mercury, chromium, and lead) are among the water pollutants [27]. In recent decades, there has been increasing interest in the use of clean water [810]. With high efficiency and low energy consumption, filtration is one of the most appropriate technologies for decreasing pollution [1115]. The formation of a good membrane is an important and practical step toward increasing the efficiency of water treatment; this film could be made from materials such as polymer, fiber, ceramic, and carbon nanotubes [1619].
In recent decades, ceramic and polymeric membranes have been increasingly used due to their strong mechanical properties, chemical stability, high efficiency in minimizing pollutants, high photocatalytic power, and odor reduction ability [2021]. The photocatalytic process of separating organic pollutants from the membrane, under visible light and UV radiation, can occur without any energy or chemical consumption. This process is especially important for the removal of different types of environmental pollutants [22]. The photocatalytic process of isolating organic pollutants from membrane, under visible light and UV radiation, can demonstrate higher efficiency [23]. Titanium dioxide is one of the important semiconductors, having properties such as low toxicity, low cost, and highly efficient removal of pollutants [212425].
TiO2 has the ability to degrade organic materials (organic dye, oil, and toxic pollutants) into H2O and CO2 [2425]. As a semiconductor, nanoparticles of anatase titanium dioxide have higher photocatalytic power than rutile and brookite crystals and exhibit absorption at 390 nm when exposed to UV radiation [26]. The present study includes a review of the semiconductive properties of TiO2 semiconductor and its photocatalytic power. Different methods, such as doping and surface chemical modification, were used to increase the photocatalytic power of TiO2 [27]. It is worth mentioning that, in different industries, such as the textile industry, TiO2graphene membranes are increasingly used to improve the efficiency of water treatment.
In recent decades, there have been many reports on the mechanical and chemical properties of graphene (G) and graphene oxide (GO) [28]. In studies concerning water treatment and purification, nanostructured carbon has been used in the form of carbon nanotube and graphene. These structures have high absorption capacity and the ability to absorb organic materials in aqueous solutions [29]. After studies on the absorption of organic aromatic pollutants by graphene, carbon nanotube, and granular activated carbon, Onu et al. discovered that graphene possessed higher efficiency in the absorption of natural organic materials.
The presence of natural organic matter (NOM) reduced the absorption of synthetic organic compound (SOC) in carbon nanotube, granular activated carbon, and, finally, graphene [30]. After comparing the absorption of methylene blue and atrazine pollutants by super fine powdered activated carbon (SPAC), carbon nanotubes (CNTs), and graphene, Ellerie et al. found that SPAC had the highest efficiency, while multiwalled carbon nanotube (MWCNT) was the least efficient in absorbing pollutants [31].
Compared to membranes without absorbents, the direct use of absorbents in membrane structure improves the efficiency of absorbed pollutants. GO consists of graphene sheets having functional groups (R-O-H, epoxy, COOH, and C=O). Therefore, the GO-TiO2 nanocomposite, under UV radiation and visible light, has the capability of photocatalytic degradation, high absorption capacity, and the capability of being more exposed to water pollutants as a result of its huge surface area of 2360 m2/gr [32]. Liu et al. studied the nanorod of  on the surface of graphene oxide sheets interface water and toluene. They produced a high efficiency nanorod (GO-TiO2 NRCs) composite. The TiO2 nanorod was stabilized with oleic acid, low temperature, and hydrolysis approach. The photocatalytic activity (GO-TiO2 NRCs) desired in the degradation of methylene blue (MB) under UV radiation was higher than GO-25 and the other original TiO2 nanorod states [33].
Rao et al. reported efficient removal of 4-chlorophene, 2,4-dichlorophene, and 2,4,6-trichlorophene by ZrO2graphene composite and, herein, removal efficiency of 4-chlorophene was by far better than the others, and the absorbability decreased with the increase in pH [34].
According to the report of Zhang et al., TiO2 nanowire membrane was able to remove TOC and HA pollution with a removal efficiency of 93.6 and 100%, respectively. This membrane was able to degrade organic materials and produce CO2 [35].
Mele et al. carried out a characterization of polycrystalline bare TiO2 for the degradation of 4-nitrophenol (4-NP) in aqueous suspension using TRMC, EPR, and XPS. The results of this study showed that propyline impregnated on the surface of TiO2 improved the photocatalytic properties compared to bare TiO2 [35].
Afzal et al. synthesized anatase TiO2/TCPP (meso-tetra(4-carboxyphenyl)porphyrin) coated with cotton. TiO2/TCPP coated with cotton for the purpose of making comparison with bare TiO2 was found to be better. There was complete degradation of methylene blue in 110 min and removal of coffee and red wine in 16 h by visible light [36].
There are many studies regarding the  semiconductor and its photocatalytic features in removing pollutants, especially industrial dyes and heavy metals; however, to date, there has been no comprehensive study regarding the photocatalytic features, flow permeability, and antifouling property. Fouling resistance in GO-TiO2 membrane, with regard to various polymer bases, has been used in this article.
1.1. Titanium Dioxide
Titanium(IV) oxide is the oxide of titanium, with a molecular weight of 79.87 g/mol and chemical formula of TiO2. When used as a pigment, it is called white titanium or pigment white. Titanium dioxide is used in industries as a mineral pigment. It occurs in nature in three forms: brookite, rutile, and anatase. White titanium dioxide is 200 to 300 nm in diameter and has a particular geometric form and structure. Table 1 shows the properties of different kinds of titanium. The first production of  was manufactured in Norway, USA, and Germany in 1918 [3942].
Table 1: The structural properties of  crystals [4345].
1.2. Photocatalysts
For several years in industrialized countries, photocatalysts have been used for the removal of pollutants which are not removed by bioprocesses. Most photocatalysts are semiconductor solid oxides which, under radiation, are activated with sufficient energy [47]. Chlorophyll in plants acts in a similar manner to photocatalysts. When compared with photosynthesis in which chlorophyll absorbs sunlight and produces oxygen and glucose by water and carbon dioxide, in the process of photocatalysis, organic materials are converted into water and carbon dioxide in the presence of light, water, and catalyst (Figure 1) [48].
Figure 1: Comparison between the actions of a catalyst and chlorophyll.
1.3. TiO2 Photocatalytic Degradation Mechanism
One of the important properties of  solid and inorganic nanomaterials is the photocatalytic activity. In many cases, this feature is used for antibacterial surfaces. Also, the photocatalytic activity of  has been applied in a wide range of metal oxides and their sulfides [4950] including ZnO [51], WO3 [52], WS2 [53], Fe2O3 [54], V2O5 [55], CeO2 [56], CdS [57], and ZnS [58]. The band gap energy values for some common semiconductor materials are presented in Table 2 [46].
Table 2: Band gap energy for some common semiconductor materials [46].
TiO2 was found to be the best semiconductor because of its chemical stability and nontoxicity. Also, it is cost-effective and has an excellent photocatalytic activity in the presence of UV irradiation [5962].
One of the commercial uses of TiO2 is P25, which is often used as a photocatalytic material. It consists of 70 to 80% anatase and 20 to 30% rutile, with specific surface area of 50 m2/g. There exists an energy gap between the valence band and the conduction band in semiconductors (Figure 2). Contrary to nonconductors, it is small; therefore, after radiation of light to the photocatalysts of semiconductors, photons with energy equal to or higher than the gap energy are absorbed, and electrons can be excited from the valence band to the conduction band.
Figure 2: Energy gap in semiconductive, nonconductive, and conductive material.
After receiving solar energy (photons), semiconductors gain more energy than usual, and, as a result, electrons travel from the valence band to the conduction band and create holes in the valence band  (Figure 3).
Figure 3: The mechanism of photocatalysis on TiO2 nanomaterials [37].
TiO2 produces this energy by receiving sunlight or UV radiation and acts as an important photocatalyst [63]. Moreover, the use of TiO2 nanomaterials as a photocatalyst for the removal of pollutants has attracted much attention due to its physical and chemical properties, high efficiency, low cost, and low toxicity.(1) is suitable for producing  hydroxyl radicals on the surface of  and also allows excitation of electrons from the valence band to the conduction band [51] and excitation of conduction band electrons to reduce the oxygen molecule.(2)(3)(4) (in alkaline solution)(5) (in neutral solution)Anatase and rutile are the two main types of  with energies of 3.2 and 3.1 eV, respectively, showing that the photocatalytic activity of anatase is much greater than that of rutile [6465].
The photocatalytic power of  is dependent not only on the energy band but also on the surface characteristics. The presence of a high surface area in each mass increases the photocatalytic power. The degradation of azo dye, especially methylene blue, by  film, depends on the surface of the photocatalyst. This film could be either in anatase crystal form, having different thicknesses and surfaces, or in low pressure and chemical deposition, from which vapor is produced [66].
 nanomaterials can be used for the photocatalytic cleaning of different types of organic compounds, such as chlorinated hydrocarbons (like organic dye, insecticides, surfactants, , and phenol), decreasing heavy metals such as , and , and are also capable of the high efficiency removal of bacteria and viruses [6569].
The pH of the solution has a great influence on the degradation of dye and TiO2 surface charge. The surface of TiO2 is positively charged in acidic pH and negatively charged in alkaline condition. The anionic dyes have strong adsorption in acidic pH while the cationic dyes have weak adsorption in alkaline pH [7071].

2. Background of Azo Dye

Dyes of textile industries and other kinds of dyes are dangerous to humans and the environment. During coloring, 1 to 20% of dyes are expelled from the wastewater of textile industries [7275].
The discharge of wastewater from textile industries into the environment is regarded as a major source of pollution and main cause of eutrophication in rivers and lakes. Also, the oxidation, hydrolysis, and chemical activities in the discharged wastewater cause potential dangers [7678].
The application of old processes (such as absorption, activated carbon, ultrafiltration, reverse osmosis, coagulation, and ion exchange) for the removal of dyes shows high efficiency. The transmission of organic materials from water to other phases, regeneration of absorbents and wastewater pretreatment, is quite expensive [7982].
Due to high levels of aromatic compounds in dyes and their stability, the biological treatment takes a longer time and the crystals produced are not suitable for degradation [8386]. Chlorination and ozonation are practical for the removal of some kinds of dyes. Both methods are expensive since they consume chemical materials, consume power, have low efficiency and limitations in the removal of carbon, and are not suitable for degradation [8790].
In recent decades, advanced oxidation processes (AOPs) have become widespread. This process is based on the production of super active species like hydroxyl (), which is capable of oxidizing a wide range of materials, but its action is nonselective. Fenton and photo-Fenton, photocatalytic activities [9195], UV/H2O2 [9697], and modified photocatalytic activity of  [98101] are some of the advanced oxidation processes (AOPs). In AOPs, the use of  as a photocatalyst has been recognized as a suitable technology [102107].
Finally, the use of  as a photocatalyst under visible light has attracted much attention because it is economical. The use of a photocatalyst in the presence of sunlight can help in reducing dyes, especially methylene blue and methyl orange, and also their mineralization [108110]. Many studies have considered the use of the desired photocatalyst for cleaning of textile industrial wastewater [111114].
2.1. Azo Dye
The main problem of pollution in textile industries is water pollution. Methylene blue and methyl orange are among the commonly used dye compounds in the industry, and they are often known as azo dyes because they contain the functional group R-N=N-R, in which N=N is known as azo. The structures of methyl orange and methylene blue compounds are shown in Figure 4. Azo dyes are divided into three groups: mono, di, and tri, and this classification is dependent on the number of azo groups.  and  can be aryl and alkyl groups, and as we know it is quite difficult to break down benzene ring; hence, the simplest benzene ring is in the aryl group.
Figure 4: Structure of (a) methyl orange and (b) methylene blue.
Studies have shown that the presence of UV radiation helps  in removing organic dyes [5051].

3. Strategies for Improved Power of TiO2

3.1. Doping
Doping titanium dioxide with metal elements such as Cr, Co, Cu, V, Mo, V, Ag, Au, Pt, Nb, and Ru and nonmetal elements such as C, N, S, P, I, F, and B increases the charge transfer reaction and thermal stability of the photocatalyst [27]. For example, X. Z. Li and F. B. Li reported the photodegradation of methylene blue in aqueous solutions under visible light over AU3+ with modified TiO2 power [115].
3.2. Surface Chemical Modification
Surface chemical modification is essential for increasing the efficiency of photocatalysts, thus preventing the recombination charge (electron and hole (h+)) and their further separation, which is performed through two methods: sensitization and coupling. Sensitization, where different groups have used narrow band gap semiconductors to enhance optical absorption properties of TiO2 nanoparticles in the visible light region, can be used to sensitize TiO2 materials. Coupling different semiconductors with different energy systems provides another way to improve charge-carrier separation.
3.2.1. Synthesis of Types of GO-TiO2 Membrane
The use of graphene, especially in the water treatment industry, has increased significantly in recent years due to its perfect mechanical and electrical features as well as high surface area. In the water treatment industry, graphene has been so useful in the removal of organic pollution, bacteria, heavy metals, and dyes, especially azo dyes [116], MB dyes and MO dyes [117118]. These dyes are produced, as mentioned before, in textile factories and have to be closely monitored to have standard concentration before being released into the environment [116]. In recent years, the synthesis of GO-TiO2 membrane for the reduction of dye pollutants, especially azo dyes, has been studied in different cases in the water treatment industry to improve TiO2.
The GO- membrane can be placed on any texture or on any membrane surface of the water filter. The photocatalytic process is an introduction to membrane properties, such as hydrophilicity, water penetration, and degradation of pollutants, and, as a result, there is a decrease in the odor produced from degradation of organic materials [117].
Studies were performed to achieve two major objectives: () studying the absorption ability of the membrane without considering the photocatalytic properties [118] and () studying the levels of degradation of organic dyes under UV radiation and considering the photocatalytic properties of the membrane [34117119].
The GO-sheet can be placed on the following:(1)Anionic perfluorinated polymer (Nafion) [120](2)Cellulose acetate [118](3)Polycarbonate (PC) [117](4)Polysulfone fluoride (PSF) [116](5)Polyvinylidene fluoride (PVDF) [120]The polymeric membranes possess excellent chemical resistance, thermal stability, and good membrane formation [121122].
The photocatalytic effects of the produced membrane on the removal of organic dyes, especially for removed RB [117], MB [116], and MO dyes [117], which are considered as aromatic compounds, have been studied.
A comparison of hybrid membranes and standard nanomembranes showed that hybrid membranes act 5 times faster than standard nanomembranes in removing pollution [123]. Also, the energy consumption of hybrid membranes is half that of standard nanomembranes. Due to their mechanical and chemical features, these filters have stable photocatalytic activity in the presence of UV radiation [124125]. It is worth mentioning that GO is synthesized from natural graphite through Hommer’ methods, similar to previous studies [126]. The TiO2 microsphere was synthesized by the reported method, with some modifications [127129].
The LbL (layer-by-layer) method has the ability to form a hybrid membrane. This experiment used a gold coated sensor to determine the mass of TiO2 and GO during the LbL procedure. The sensor was coated with a base polymer membrane. The sensor was put in a desiccator for drying, after which it was put in DI water. A fixed concentration on the surface of the sensor was at 0 ng/cm2 after 15 min, and the DI water was replaced with TiO2 solution. TiO2 nanoparticles were adsorbed on the base polymer and, after 2 h, the mass of the nanoparticle slowed down and became stabilized. After that, TiO2 solution was replaced with DI water for the removal of packed nanoparticles. Next, the TiO2 coated sensor was soaked in the GO solution [130].
According to Athanasekou et al., the reduced graphene oxide-TiO2 (GO-TiO2) was synthesized with GO suspension. GO was prepared according to previous studies, similar to the Hommer method, using LPD (liquid phase deposition) method at room temperature. Among all the membranes, GO-TiO2 was the best membrane for removing MO under UV, and GO-T10 was the best membrane for removing MB. The hybrid membrane when compared with nanofiltration was better for the removal of pollutants and energy consumption [131].
3.2.2. Reaction Schemes for Surface Modification GO-TiO2
By forming Ti-O band between Ti4+ or functional group membrane and hydrogen band,  nanomaterials are placed on the surface of the base membrane between groups and the functional group [132]. In the next phase, GO is placed on  layers via Ti-O band or the band between hydrogen and Ti4+ and carboxyl groups of GO, and, eventually, GO reduced and bonded to  by ethanol under UV radiation [133134]. One of the advantages of the GO-TiO2 membrane is flexibility and great force in passing flow; the resistance to GO-TiO2 is obvious after drying GO-TiO2 and the absence of cracks on the membrane. TiO2 alone is not highly efficient in removing organic materials and pollutants [135136].

4. GO-TiO2 Membrane

Membranes are typically made of inorganic (ceramic) materials and a polymeric membrane. Membrane separation is basically on three principles: adsorption, sieving, and electrostatics as shown in Figure 5 [38]. The pore size determines the different membrane types: microfiltration (MF), ultrafiltration (UF), nanofiltration (NF), and reverse osmosis (RO). Among the membrane processes, UF with polymeric membrane has the efficiency of removing contaminants such as organic pollutants, heavy metal, suspended solid, and different dyes. The polymeric membrane graphene oxide-TiO2 is significantly efficient in the removal of organic pollutants and dye from solution. It has specific characteristics which completely explain some special traits.
Figure 5: Schematic representing the basic principles involved in membrane separation [38].
4.1. Analysis of the GO-TiO2 Membrane
(i)Membrane analysis through transmission electron microscopy (TEM) and EDX shows the location of  anatase crystal particles on the surface of GO [137138].(ii)The amount of TiO2 on the surface of GO sheets is determined by thermogravimetric analysis (TGA) [139].(iii)The amounts of Ti and GO have been studied in the following researches through CS and XPS analysis [140143].(iv)The absorption capacity of the GO-membrane is determined when different dye solutions are used in water [144145].
4.1.1. Porosity of Membrane
Atomic force microscopy (AFM) can record the roughness and surface morphology of polymeric membranes. The overall porosity of a membrane can be calculated using the gravimetric method as presented in [146]where  and  are the wet and dry weights of the membrane,  is the effective area of the membrane (m2),  is the membrane thickness (m), and  is the water of density 0.998 g/cm3. Using pure water flux, the porosity of the membrane and the mean pore radius of the membrane () can be calculated using the Guerout–Elford–Ferry equation:where  is the water viscosity (8.9 × 10−4 pas), ΔP is the operation pressure (3 MPa),  is the permeated pure water amount (m3/s), L is the membrane thickness (m), and A is the surface area (m2) [146147].
Atomic force microscopy, field emission scanning electron microscopy, and the Guerout–Elford–Ferry equation applied for pure water permeation rate were used for determination of average pore size. The pore size calculated from the method of Guerout–Elford–Ferry was the highest, and, based on the make-ready sample (FESEM), the pore size calculated was the least in size [148].
4.2. GO-TiO2 Photocatalytic Membrane for Dye Removal
Photocatalytic properties of  nanoparticles under UV radiation for removal of dye occur according to the following reactions [149]. In the photocatalytic oxidation, TiO2 has to be irradiated and excited in near-UV energy to induce charge separation. On the other hand, dyes rather than TiO2 are excited by visible light followed by electron injection onto the TiO2 conduction band, which leads to photosensitized oxidation. Subsequently, this is followed by electron injection from the excited dye molecule onto the conduction band of the TiO2 particles, whereas the dye is converted to the cationic dye radicals  that undergo degradation to yield products:(6) +  excitation(7) + () electron injection(8) + ( recombination(9) degradation(10)O2 + ( (electron scavenging)Anatase crystals which form mesosphere can have high mass transfer flux and photocatalytic activity [150152]. The mesosphere structure of GO-TiO2 has a huge surface area which forms numerous channels for passing water flow and photocatalysts for photocatalytic degradation [153154]. It is absolutely obvious that a membrane with high cross flow, removal efficiency, and low cost is suitable for filtration. GO sheets increase the filtration efficiency [154].
According to studies performed in different researches, GO-TiO2 on a base membrane and under UV radiation shows great photocatalytic power in degrading organic materials. Membrane pretreatment, using ethanol and UV radiation, helps in increasing the degradation of MB [117].
The photocatalytic properties of porphyrin TiO2 and graphene TiO2 under visible light and UV radiation have been studied. Porphyrin graphene TiO2 shows 35% higher efficiency than pure TiO2 [155].
GO- is used for the degradation of RhB and  dyes under UV radiation within 20 to 30 min. In the absence of UV radiation, GO-TiO2 membrane shows low efficiency, lower than 15%, in the degradation of RhB and  dyes. The presence of UV radiation increases removal efficiency by 50% [118].
The presence of photocatalysts on a membrane prevents the accumulation of organic materials and macromolecules and therefore allows higher flow to pass through the membrane. Absorption of MO in darkness has minimal effects on the Nafion-GO SULF membrane, but under UV radiation the initial concentration is reduced by 90% [28118]. Pores size depends on the amount of GO- which is placed on the base membrane. GO-10 membrane (membrane having 10 nm pores) shows higher removal efficiency of dyes than other membranes [156].
The application of the method of liquid phase deposition (LPD) for synthesis of GO-TiO2 composites and their stabilization on the integrated membranes is suitable for the removal of industrial dyes. Membranes with 1 m length and 30 monoliths consume less energy than nanomembranes [156].
4.3. Permeation Flux through GO-TiO2 Membrane
Filtration tests on GO-TiO2 membrane with 5%-5 mL/min flow rate, 0.6 nm constant membrane thickness, and 30 mg/L concentration cobalt ion show that the lesser the flow which is passed through, the more the cobalt removed. As the initial concentration of cobalt ion increases from 1 to 9 mg/L, the removal efficiency decreases from 89 to 21%. The high absorption capacity of nanosheets of graphene oxide ammonium for the removal of cobalt could be related to the ammonium functional groups. These absorbents are able to remove 30 mg/L concentration of cobalt ions in 5 min with 90% efficiency [119].
The presence of sulfonic functional group, along with graphene on Nafion membrane, allows more water to pass through it [2838118157]. The movement of proton on Nafion membrane can help in separating ion compounds from the water molecule. These effects have been studied by adding the sulfonic functional group to the surface membrane of Nafion [38109158].
The thickness of GO-membrane could be changed easily by changing the GO-TiO2 mass which is placed on the base. Membrane thickness affects the flow passing through the membrane, meaning that a decrease in membrane thickness allows higher flow to pass through the membrane and decreases the removal efficiency [118].
4.4. Fouling Resistant Membrane
The total fouling resistance of the polymeric membrane () is calculated from  and  is a reversible fouling ratio which describes the fouling caused by concentration polarization, and  is an irreversible fouling ratio which describes the fouling caused by adsorption or deposition of protein molecules on the membrane surface. is the sum or  and , and  is the permeation water flux (kg/m2 h):where  is the weight of collected permeation flux,  is the membrane effective area,  is the permeation time,  is the water flux cleaned membrane, and  is the flux of the BSA solution [144]. Hence,Under special pressure, after the pure water test, the BSA solutions are immediately replaced in the filtration cell for 90 min [159].
4.5. Antifouling Property of GO-TiO2 Membrane
Since the cost of cleaning normal membranes is too high and, over time, they produce odors because of the accumulation of organic materials, the application of GO- membrane in the water treatment industry seems to have become more widespread. Humic acid as a standard pollutant formed from natural organic materials and carcinogen disinfectants has been studied [160].
In the first phases of filtration, the accumulation of humic acid on the surface of membrane and closing of the pores are really obvious after 30 min [161]. After 2 h, with 7 L/ flow rate, a layer of odorous HA cake becomes noticeable [162]. UV radiation reduces the produced odor significantly [118]. The reason is the degradation of humic acid and production of  and  under UV radiation; in this state, there is no reduction in the flow which passed through the membrane.

5. Future Work

GO-TiO2 membrane, because of its significant applications such as high photocatalytic and antifouling property, flux permeation, and removal of organic pollutants, can be potentially used in the degradation of polycyclic aromatic hydrocarbon (PAH) as a low cost membrane in the future.

6. Conclusion

This study was carried out on the characteristics of different polymeric membranes of GO-TiO2. According to previous studies, this membrane has lots of advantages like the capability of passing flow. Graphene oxide has a number of covalent attached oxygen containing groups such as epoxy, carbonyl, and carboxyl. The existence of these groups makes membrane possess good hydrophilicity. Another good property of this membrane is the existence of photocatalytic TiO2 semiconductor, which can help in the degradation of the cake layer of the organic material and the reduction of odor. The object is very economical to reduce cost of washing membrane compared to the usual form. Fouling of this membrane because of GO-TiO2 photocatalytic properties is also reduced from other forms of membranes. This result shows that the GO-TiO2 polymer membrane could make considerable improvement in the field of water treatment and could have significant effects on the filtration efficiency.

Competing Interests

The authors declare that they have no competing interests.

References

  1. M. Elimelech and W. A. Phillip, “The future of seawater desalination: energy, technology, and the environment,” Science, vol. 333, no. 6043, pp. 712–717, 2011. View at Publisher · View at Google Scholar ·View at Scopus
  2. M. A. Shannon, P. W. Bohn, M. Elimelech, J. G. Georgiadis, B. J. MarÄ©as, and A. M. Mayes, “Science and technology for water purification in the coming decades,” Nature, vol. 452, no. 7185, pp. 301–310, 2008.View at Publisher · View at Google Scholar · View at Scopus
  3. S. Kaur, R. Gopal, W. J. Ng, S. Ramakrishna, and T. Matsuura, “Next-generation fibrous media for water treatment,” MRS Bulletin, vol. 33, no. 1, pp. 21–26, 2008. View at Publisher · View at Google Scholar ·View at Scopus
  4. Z. Wu and D. Zhao, “Ordered mesoporous materials as adsorbents,” Chemical Communications, vol. 47, no. 12, pp. 3332–3338, 2011. View at Publisher · View at Google Scholar · View at Scopus
  5. S. E. Bailey, T. J. Olin, R. M. Bricka, and D. D. Adrian, “A review of potentially low-cost sorbents for heavy metals,” Water Research, vol. 33, no. 11, pp. 2469–2479, 1999. View at Publisher · View at Google Scholar · View at Scopus
  6. A. Netzer and D. E. Hughes, “Adsorption of copper, lead and cobalt by activated carbon,” Water Research, vol. 18, no. 8, pp. 927–933, 1984. View at Publisher · View at Google Scholar · View at Scopus
  7. C. Gómez-Lahoz, F. Garcia-Herruzo, J. M. Rodriguez-Maroto, and J. J. Rodriguez, “Cobalt(II) removal from water by chemical reduction with sodium borohydride,” Water Research, vol. 27, no. 6, pp. 985–992, 1993. View at Publisher · View at Google Scholar
  8. M. M. Pendergast and E. M. V. Hoek, “A review of water treatment membrane nanotechnologies,” Energy and Environmental Science, vol. 4, no. 6, pp. 1946–1971, 2011. View at Publisher · View at Google Scholar· View at Scopus
  9. H.-W. Liang, X. Cao, W.-J. Zhang et al., “Robust and highly efficient free-standing carbonaceous nanofiber membranes for water purification,” Advanced Functional Materials, vol. 21, no. 20, pp. 3851–3858, 2011. View at Publisher · View at Google Scholar · View at Scopus
  10. Y. Liu, Z. Wu, X. Chen, Z. Shao, H. Wang, and D. Zhao, “A hierarchical adsorption material by incorporating mesoporous carbon into macroporous chitosan membranes,” Journal of Materials Chemistry, vol. 22, no. 24, pp. 11908–11911, 2012. View at Publisher · View at Google Scholar · View at Scopus
  11. J. Cadotte, R. Forester, M. Kim, R. Petersen, and T. Stocker, “Nanofiltration membranes broaden the use of membrane separation technology,” Desalination, vol. 70, no. 1–3, pp. 77–88, 1988. View at Publisher ·View at Google Scholar · View at Scopus
  12. E. A. Jackson and M. A. Hillmyer, “Nanoporous membranes derived from block copolymers: from drug delivery to water filtration,” ACS Nano, vol. 4, no. 7, pp. 3548–3553, 2010. View at Publisher · View at Google Scholar · View at Scopus
  13. H. P. Dijkstra, G. P. M. van Klink, and G. van Koten, “The use of ultra- and nanofiltration techniques in homogeneous catalyst recycling,” Accounts of Chemical Research, vol. 35, no. 9, pp. 798–810, 2002. View at Publisher · View at Google Scholar · View at Scopus
  14. C. Stoquart, P. Servais, P. R. Bérubé, and B. Barbeau, “Hybrid membrane processes using activated carbon treatment for drinking water: a review,” Journal of Membrane Science, vol. 411-412, pp. 1–12, 2012. View at Publisher · View at Google Scholar · View at Scopus
  15. P. Wang, J. Ma, F. Shi, Y. Ma, Z. Wang, and X. Zhao, “Behaviors and effects of differing dimensional nanomaterials in water filtration membranes through the classical phase inversion process: a review,” Industrial and Engineering Chemistry Research, vol. 52, no. 31, pp. 10355–10363, 2013. View at Publisher· View at Google Scholar · View at Scopus
  16. B. W. Stanton, J. J. Harris, M. D. Miller, and M. L. Bruening, “Ultrathin, multilayered polyelectrolyte films as nanofiltration membranes,” Langmuir, vol. 19, no. 17, pp. 7038–7042, 2003. View at Publisher · View at Google Scholar · View at Scopus
  17. S. El-Safty, A. Shahat, M. R. Awual, and M. Mekawy, “Large three-dimensional mesocage pores tailoring silica nanotubes as membrane filters: nanofiltration and permeation flux of proteins,” Journal of Materials Chemistry, vol. 21, no. 15, pp. 5593–5603, 2011. View at Publisher · View at Google Scholar · View at Scopus
  18. B. Zhao, L. Zhang, X.-Y. Wang et al., “Research progress in nanofiltration membrane based on carbon nanotubes,” New Carbon Materials, vol. 26, no. 5, pp. 321–327, 2011. View at Google Scholar · View at Scopus
  19. J. Sekulić, J. E. Ten Elshof, and D. H. A. Blank, “A microporous titania membrane for nanofiltration and pervaporation,” Advanced Materials, vol. 16, no. 17, pp. 1546–1550, 2004. View at Publisher · View at Google Scholar · View at Scopus
  20. B. Van Der Bruggen, C. Vandecasteele, T. Van Gestel, W. Doyen, and R. Leysen, “A review of pressure-driven membrane processes in wastewater treatment and drinking water production,” Environmental Progress, vol. 22, no. 1, pp. 46–56, 2003. View at Publisher · View at Google Scholar · View at Scopus
  21. X. Zhang, T. Zhang, J. Ng, and D. D. Sun, “High-performance multifunctional TiO2 nanowire ultrafiltration membrane with a hierarchical layer structure for water treatment,” Advanced Functional Materials, vol. 19, no. 23, pp. 3731–3736, 2009. View at Publisher · View at Google Scholar · View at Scopus
  22. M. N. Chong, B. Jin, C. W. K. Chow, and C. Saint, “Recent developments in photocatalytic water treatment technology: a review,” Water Research, vol. 44, no. 10, pp. 2997–3027, 2010. View at Publisher ·View at Google Scholar · View at Scopus
  23. J. Grzechulska-Damszel, M. Tomaszewska, and A. W. Morawski, “Integration of photocatalysis with membrane processes for purification of water contaminated with organic dyes,” Desalination, vol. 241, no. 1–3, pp. 118–126, 2009. View at Publisher · View at Google Scholar · View at Scopus
  24. G. L. Liu, C. Han, M. Pelaez et al., “Enhanced visible light photo catalytic activity of C-N-codoped TiO2films for the degradation of microcystin-LR,” Journal of Molecular Catalysis A: Chemical, vol. 372, pp. 58–65, 2013. View at Publisher · View at Google Scholar
  25. G. L. Liu, C. Han, M. Pelaez et al., “Synthesis, characterization and photocatalytic evaluation of visible light activated C-doped TiO2 nanoparticles,” Nanotechnology, vol. 23, no. 29, Article ID 294003, 2012.View at Publisher · View at Google Scholar · View at Scopus
  26. X. Zhang, A. J. Du, P. Lee, D. D. Sun, and J. O. Leckie, “TiO2 nanowire membrane for concurrent filtration and photocatalytic oxidation of humic acid in water,” Journal of Membrane Science, vol. 313, no. 1-2, pp. 44–51, 2008. View at Publisher · View at Google Scholar · View at Scopus
  27. R. M. Mohamed, D. L. McKinney, and W. M. Sigmund, “Enhanced nanocatalysts,” Materials Science and Engineering: R: Reports, vol. 73, no. 1, pp. 1–13, 2012. View at Publisher · View at Google Scholar · View at Scopus
  28. T.-F. Yeh, J. Cihlář, C.-Y. Chang, C. Cheng, and H. Teng, “Roles of graphene oxide in photocatalytic water splitting,” Materials Today, vol. 16, no. 3, pp. 78–84, 2013. View at Publisher · View at Google Scholar ·View at Scopus
  29. G. P. Rao, C. Lu, and F. Su, “Sorption of divalent metal ions from aqueous solution by carbon nanotubes: a review,” Separation and Purification Technology, vol. 58, no. 1, pp. 224–231, 2007. View at Publisher ·View at Google Scholar · View at Scopus
  30. A. G. Onu, W. Qil, Z. Yan, and K. Tan, “Adsorption of aromatic organic contaminants by graphene nanosheets: comparison with carbon nanotubes and activated carbon,” Water Research, vol. 47, no. 4, pp. 1648–1654, 2013. View at Google Scholar
  31. J. R. Ellerie, O. G. Apul, T. Karanfil, and D. A. Ladner, “Comparing graphene, carbon nanotubes, and superfine powdered activated carbon as adsorptive coating materials for microfiltration membranes,” Journal of Hazardous Materials, vol. 261, pp. 91–98, 2013. View at Publisher · View at Google Scholar ·View at Scopus
  32. A. W. H. Mau, C. B. Huang, N. Kakuta et al., “Hydrogen photoproduction by Nafion/cadmium sulfide/platinum films in water/sulfide ion solutions,” Journal of the American Chemical Society, vol. 106, no. 65, pp. 37–42, 1984. View at Publisher · View at Google Scholar
  33. J. Liu, H. Bai, Y. Wang, Z. Liu, X. Zhang, and D. D. Sun, “Self-assembling TiO2 nanorods on large graphene oxide sheets at a two-phase interface and their anti-recombination in photocatalytic applications,” Advanced Functional Materials, vol. 20, no. 23, pp. 4175–4181, 2010. View at Publisher ·View at Google Scholar · View at Scopus
  34. R. A. K. Rao, S. Singh, B. R. Singh, W. Khan, and A. H. Naqvi, “Synthesis and characterization of surface modified graphene-zirconium oxide nanocomposite and its possible use for the removal of chlorophenol from aqueous solution,” Journal of Environmental Chemical Engineering, vol. 2, no. 1, pp. 199–210, 2014.View at Publisher · View at Google Scholar · View at Scopus
  35. G. Mele, R. D. Sole, G. Vasapollo et al., “TRMC, XPS, and EPR characterizations of polycrystalline TiO2porphyrin impregnated powders and their catalytic activity for 4-nitrophenol photodegradation in aqueous suspension,” The Journal of Physical Chemistry B, vol. 109, no. 25, pp. 12347–12352, 2005. View at Publisher · View at Google Scholar · View at Scopus
  36. S. Afzal, W. A. Daoud, and S. J. Langford, “Self-cleaning cotton by porphyrin-sensitized visible-light photocatalysis,” Journal of Materials Chemistry, vol. 22, no. 9, pp. 4083–4088, 2012. View at Publisher ·View at Google Scholar · View at Scopus
  37. A. R. Khataee and M. B. Kasiri, “Photocatalytic degradation of organic dyes in the presence of nanostructured titanium dioxide: influence of the chemical structure of dyes,” Journal of Molecular Catalysis A: Chemical, vol. 328, no. 1-2, pp. 8–26, 2010. View at Publisher · View at Google Scholar · View at Scopus
  38. M. Padaki, R. Surya Murali, M. S. Abdullah et al., “Membrane technology enhancement in oil-water separation. A review,” Desalination, vol. 357, pp. 197–207, 2015. View at Publisher · View at Google Scholar · View at Scopus
  39. Kirk-Othmer, Encyclopadia of Chemical Technology, vol. 19, Wiley-Inter Science Publication, 4th edition, 1996.
  40. H. Takeda and O. Ishitani, “Development of efficient photocatalytic systems for CO2 reduction using mononuclear and multinuclear metal complexes based on mechanistic studies,” Coordination Chemistry Reviews, vol. 254, no. 3-4, pp. 346–354, 2010. View at Publisher · View at Google Scholar · View at Scopus
  41. R. Zallen and M. P. Moret, “The optical absorption edge of brookite TiO2,” Solid State Communications, vol. 137, no. 3, pp. 154–157, 2006. View at Publisher · View at Google Scholar · View at Scopus
  42. X. Chen, “Titanium dioxide nanomaterials and their energy applications,” Chinese Journal of Catalysis, vol. 30, no. 8, pp. 839–851, 2009. View at Publisher · View at Google Scholar · View at Scopus
  43. D. T. Chomer and K. Herrington, “The structures of anatase and rutile,” Journal of the American Chemical Society, vol. 77, no. 18, pp. 4708–4709, 1955. View at Publisher · View at Google Scholar · View at Scopus
  44. V. W. H. Baur, “Atomabstände und bindungswinkel im brookit, TiO2,” Acta Crystallographica, vol. 14, no. 3, pp. 214–216, 1961. View at Publisher · View at Google Scholar
  45. S.-D. Mo and W. Y. Ching, “Electronic and optical properties of three phases of titanium dioxide: rutile, anatase, and brookite,” Physical Review B, vol. 51, no. 19, pp. 13023–13032, 1995. View at Publisher ·View at Google Scholar · View at Scopus
  46. R. Richards, Surface and Nanomolecular Catalysis, CRC Press/Taylor & Francis Group, 2006.
  47. M. Zhang, G. Sheng, J. Fu, T. An, X. Wang, and X. Hu, “Novel preparation of nanosized ZnO–SnO2 with high photocatalytic activity by homogeneous co-precipitation method,” Materials Letters, vol. 59, no. 28, pp. 3641–3644, 2005. View at Publisher · View at Google Scholar · View at Scopus
  48. M. Soltaninezhad and A. Aminifar, “Study nanostructures of semiconductor zinc oxide (ZnO) as a photocatalyst for the degradation of organic pollutants,” International Journal of Nano Dimension, vol. 2, no. 2, pp. 137–145, 2011. View at Google Scholar
  49. A. Mills and S. Le Hunte, “An overview of semiconductor photocatalysis,” Journal of Photochemistry and Photobiology A: Chemistry, vol. 108, no. 1, pp. 1–35, 1997. View at Publisher · View at Google Scholar ·View at Scopus
  50. N. Daneshvar, D. Salari, and A. R. Khataee, “Photocatalytic degradation of azo dye acid red 14 in water: investigation of the effect of operational parameters,” Journal of Photochemistry and Photobiology A: Chemistry, vol. 157, no. 1, pp. 111–116, 2003. View at Publisher · View at Google Scholar · View at Scopus
  51. N. Daneshvar, D. Salari, and A. R. Khataee, “Photocatalytic degradation of azo dye acid red 14 in water on ZnO as an alternative catalyst to TiO2,” Journal of Photochemistry and Photobiology A: Chemistry, vol. 162, no. 2-3, pp. 317–322, 2004. View at Publisher · View at Google Scholar · View at Scopus
  52. X. F. Cheng, W. H. Leng, D. P. Liu, J. Q. Zhang, and C. N. Cao, “Enhanced photoelectrocatalytic performance of Zn-doped WO3 photocatalysts for nitrite ions degradation under visible light,” Chemosphere, vol. 68, no. 10, pp. 1976–1984, 2007. View at Publisher · View at Google Scholar · View at Scopus
  53. D. Jing and L. Guo, “WS2 sensitized mesoporous TiO2 for efficient photocatalytic hydrogen production from water under visible light irradiation,” Catalysis Communications, vol. 8, no. 5, pp. 795–799, 2007.View at Publisher · View at Google Scholar · View at Scopus
  54. J. Bandara, U. Klehm, and J. Kiwi, “Raschig rings-Fe2O3 composite photocatalyst activate in the degradation of 4-chlorophenol and Orange II under daylight irradiation,” Applied Catalysis B: Environmental, vol. 76, no. 1-2, pp. 73–81, 2007. View at Publisher · View at Google Scholar · View at Scopus
  55. K. Teramura, T. Tanaka, M. Kani, T. Hosokawa, and T. Funabiki, “Selective photo-oxidation of neat cyclohexane in the liquid phase over V2O5/Al2O3,” Journal of Molecular Catalysis A: Chemical, vol. 208, no. 1-2, pp. 299–305, 2004. View at Publisher · View at Google Scholar · View at Scopus
  56. Y. Zhai, S. Zhang, and H. Pang, “Preparation, characterization and photocatalytic activity of CeO2nanocrystalline using ammonium bicarbonate as precipitant,” Materials Letters, vol. 61, no. 8-9, pp. 1863–1866, 2007. View at Publisher · View at Google Scholar · View at Scopus
  57. K. G. Kanade, J.-O. Baeg, U. P. Mulik, D. P. Amalnerkar, and B. B. Kale, “Nano-CdS by polymer-inorganic solid-state reaction: visible light pristine photocatalyst for hydrogen generation,” Materials Research Bulletin, vol. 41, no. 12, pp. 2219–2225, 2006. View at Publisher · View at Google Scholar · View at Scopus
  58. C. L. Torres-Martínez, R. Kho, O. I. Mian, and R. K. Mehra, “Efficient photocatalytic degradation of environmental pollutants with mass-produced ZnS nanocrystals,” Journal of Colloid and Interface Science, vol. 240, no. 2, pp. 525–532, 2001. View at Publisher · View at Google Scholar · View at Scopus
  59. M. R. Hoffmann, S. T. Martin, W. Y. Choi, and D. W. Bahnemann, “Environmental applications of semiconductor photocatalysis,” Chemical Reviews, vol. 95, no. 1, pp. 69–96, 1995. View at Publisher ·View at Google Scholar · View at Scopus
  60. A. L. Linsebigler, G. Q. Lu, and J. T. Yates, “Photocatalysis on TiO2 surfaces: principles, mechanisms, and selected results,” Chemical Reviews, vol. 95, no. 3, pp. 735–758, 1995. View at Publisher · View at Google Scholar · View at Scopus
  61. O. Zahraa, S. Maire, F. Evenou et al., “Treatment of wastewater dyeing agent by photocatalytic process in solar reactor,” International Journal of Photoenergy, vol. 2006, Article ID 46961, 9 pages, 2006. View at Publisher · View at Google Scholar · View at Scopus
  62. A. Fujishima, T. N. Rao, and D. A. Tryk, “Titanium dioxide photocatalysis,” Journal of Photochemistry and Photobiology C: Photochemistry Reviews, vol. 1, no. 1, pp. 1–21, 2000. View at Publisher · View at Google Scholar · View at Scopus
  63. N. Daneshvar, D. Salari, A. Niaei, M. H. Rasoulifard, and A. R. Khataee, “Immobilization of TiO2nanopowder on glass beads for the photocatalytic decolorization of an azo dye C.I. Direct Red 23,” Journal of Environmental Science and Health, Part A: Toxic, vol. 40, no. 8, pp. 1605–1617, 2005. View at Publisher · View at Google Scholar · View at Scopus
  64. V. Loddo, G. Marcì, C. Martín, L. Palmisano, V. Rives, and A. Sclafani, “Preparation and characterisation of TiO2 (anatase) supported on TiO2 (rutile) catalysts employed for 4-nitrophenol photodegradation in aqueous medium and comparison with TiO2 (anatase) supported on Al2O3,” Applied Catalysis B: Environmental, vol. 20, no. 1, pp. 29–45, 1999. View at Publisher · View at Google Scholar · View at Scopus
  65. S. Bakardjieva, J. Å ubrt, V. Å tengl, M. J. Dianez, and M. J. Sayagues, “Photoactivity of anatase-rutile TiO2nanocrystalline mixtures obtained by heat treatment of homogeneously precipitated anatase,” Applied Catalysis B: Environmental, vol. 58, no. 3-4, pp. 193–202, 2005. View at Publisher · View at Google Scholar · View at Scopus
  66. S.-C. Jung, S.-J. Kim, N. Imaishi, and Y.-I. Cho, “Effect of TiO2 thin film thickness and specific surface area by low-pressure metal–organic chemical vapor deposition on photocatalytic activities,” Applied Catalysis B: Environmental, vol. 55, no. 4, pp. 253–257, 2005. View at Publisher · View at Google Scholar ·View at Scopus
  67. X. Wang, J. C. Yu, P. Liu, X. Wang, W. Su, and X. Fu, “Probing of photocatalytic surface sites on SO42−/TiO2 solid acids by in situ FT-IR spectroscopy and pyridine adsorption,” Journal of Photochemistry and Photobiology A: Chemistry, vol. 179, no. 3, pp. 339–347, 2006. View at Publisher ·View at Google Scholar · View at Scopus
  68. M. Zheng, M. Gu, Y. Jin, and G. Jin, “Preparation, structure and properties of TiO2–PVP hybrid films,” Materials Science and Engineering: B, vol. 77, no. 1, pp. 55–59, 2000. View at Publisher · View at Google Scholar · View at Scopus
  69. K. Tennakone and K. G. U. Wijayantha, “Heavy-metal extraction from aqueous medium with an immobilized TiO2 photocatalyst and a solid sacrificial agent,” Journal of Photochemistry and Photobiology A: Chemistry, vol. 113, no. 1, pp. 89–92, 1998. View at Publisher · View at Google Scholar · View at Scopus
  70. M. N. Pons, A. Alinsafi, F. Evenou et al., “Treatment of textile industry wastewater by supported photocatalysis,” Dyes and Pigments, vol. 74, no. 2, pp. 439–445, 2007. View at Publisher · View at Google Scholar · View at Scopus
  71. N. Keller, G. Rebmann, E. Barraud, O. Zahraa, and V. Keller, “Macroscopic carbon nanofibers for use as photocatalyst support,” Catalysis Today, vol. 101, no. 3-4, pp. 323–329, 2005. View at Publisher · View at Google Scholar · View at Scopus
  72. H. Zollinger, Ed., Color Chemistry: Synthesis, Properties and Applications of Organic Dyes and Pigments, VCH, 2nd edition, 1991.
  73. E. J. Weber and V. C. Stickney, “Hydrolysis kinetics of reactive blue 19-vinyl sulfone,” Water Research, vol. 27, no. 1, pp. 63–67, 1993. View at Publisher · View at Google Scholar · View at Scopus
  74. C. Ràfols and D. Barceló, “Determination of mono- and disulphonated azo dyes by liquid chromatography–atmospheric pressure ionization mass spectrometry,” Journal of Chromatography A, vol. 777, no. 1, pp. 177–192, 1997. View at Publisher · View at Google Scholar · View at Scopus
  75. A. Houas, H. Lachheb, M. Ksibi, E. Elaloui, C. Guillard, and J.-M. Herrmann, “Photocatalytic degradation pathway of methylene blue in water,” Applied Catalysis B: Environmental, vol. 31, no. 2, pp. 145–157, 2001. View at Publisher · View at Google Scholar · View at Scopus
  76. U. Pagga and D. Brown, “The degradation of dyestuffs: part II behaviour of dyestuffs in aerobic biodegradation tests,” Chemosphere, vol. 15, no. 4, pp. 479–491, 1986. View at Publisher · View at Google Scholar · View at Scopus
  77. A. B. Prevot, C. Baiocchi, M. C. Brussino et al., “Photocatalytic degradation of acid blue 80 in aqueous solutions containing TiO2 suspensions,” Environmental Science and Technology, vol. 35, no. 5, pp. 971–976, 2001. View at Publisher · View at Google Scholar · View at Scopus
  78. M. Saquib and M. Muneer, “TiO2-mediated photocatalytic degradation of a triphenylmethane dye (gentian violet), in aqueous suspensions,” Dyes and Pigments, vol. 56, no. 1, pp. 37–49, 2003. View at Publisher · View at Google Scholar · View at Scopus
  79. W. Z. Tang and H. An, “UV/TiO2 photocatalytic oxidation of commercial dyes in aqueous solutions,” Chemosphere, vol. 31, no. 9, pp. 4157–4170, 1995. View at Publisher · View at Google Scholar · View at Scopus
  80. V. Meshko, L. Markovska, M. Mincheva, and A. E. Rodrigues, “Adsorption of basic dyes on granular acivated carbon and natural zeolite,” Water Research, vol. 35, no. 14, pp. 3357–3366, 2001. View at Publisher · View at Google Scholar · View at Scopus
  81. W. S. Kuo and P. H. Ho, “Solar photocatalytic decolorization of methylene blue in water,” Chemosphere, vol. 45, no. 1, pp. 77–83, 2001. View at Publisher · View at Google Scholar · View at Scopus
  82. C. Galindo, P. Jacques, and A. Kalt, “Photooxidation of the phenylazonaphthol AO20 on TIO2: kinetic and mechanistic investigations,” Chemosphere, vol. 45, no. 6-7, pp. 997–1005, 2001. View at Publisher ·View at Google Scholar · View at Scopus
  83. S. S. Patll and V. M. Shinde, “Biodegradation studies of aniline and nitrobenzene in aniline plant wastewater by gas chromatography,” Environmental Science and Technology, vol. 22, no. 10, pp. 1160–1165, 1988. View at Publisher · View at Google Scholar · View at Scopus
  84. A. T. Moore, A. Vira, and S. Fogel, “Biodegradation of trans-1,2-dichloroethylene by methane-utilizing bacteria in an aquifer simulator,” Environmental Science and Technology, vol. 23, no. 4, pp. 403–406, 1989.View at Publisher · View at Google Scholar · View at Scopus
  85. V. M. Correia, T. Stephenson, and S. J. Judd, “Characterization of textile wastewaters—a review,” Environmental Technology, vol. 15, no. 10, pp. 917–929, 1994. View at Publisher · View at Google Scholar
  86. I. Arslan and I. A. Balcioglu, “Degradation of commercial reactive dyestuffs by heterogenous and homogenous advanced oxidation processes: a comparative study,” Dyes and Pigments, vol. 43, no. 2, pp. 95–108, 1999. View at Publisher · View at Google Scholar · View at Scopus
  87. S. H. Lin and C. M. Lin, “Treatment of textile waste effluents by Ozonation and chemical coagulation,” Water Research, vol. 27, no. 12, pp. 1743–1748, 1993. View at Publisher · View at Google Scholar · View at Scopus
  88. S. H. Lin and W. Y. Liu, “Continuous treatment of textile water by ozonation and coagulation,” Journal of Environmental Engineering, vol. 120, no. 2, pp. 437–446, 1994. View at Publisher · View at Google Scholar· View at Scopus
  89. F. Strickland and S. Perkins, “Decolorization of continuous dyeing wastewater by ozonation,” Textile Chemist & Colorist, vol. 27, no. 5, pp. 11–15, 1995. View at Google Scholar
  90. P. Aranyosi, Z. Csepregi, I. Rusznák, L. Töke, and A. Víg, “The light stability of azo dyes and azo dyeings. III. The effect of artificial perspiration on the light stability of reactive and non-reactive derivatives of two selected azo chromophores in aqueous solution,” Dyes and Pigments, vol. 37, no. 1, pp. 33–45, 1998. View at Publisher · View at Google Scholar · View at Scopus
  91. W. G. Kuo, “Decolorizing dye wastewater with Fenton's reagent,” Water Research, vol. 26, no. 7, pp. 881–886, 1992. View at Publisher · View at Google Scholar
  92. E. Balanosky, J. Fernadez, J. Kiwi, and A. Lopez, “Degradation of membrane concentrates of the textile industry by fenton like reactions in iron-free solutions at biocompatible pH values (pH7-8),” Water Science & Technology, vol. 40, no. 4-5, pp. 417–424, 1999. View at Publisher · View at Google Scholar
  93. W. Feng, D. Nansheng, and Z. Yuegang, “Discoloration of dye solutions induced by solar photolysis of ferrioxalate in aqueous solutions,” Chemosphere, vol. 39, no. 12, pp. 2079–2085, 1999. View at Publisher ·View at Google Scholar · View at Scopus
  94. C. Morrison, J. Bandara, and J. Kiwi, “Sun light induced decolouration/degradation of non-biodegradation Orange dye by advanced Oxidation technologies in homogenous and heterogeneous media,” Journal of Advanced Oxidation Technologies, vol. 1, no. 2, pp. 160–169, 1996. View at Google Scholar
  95. S.-F. Kang, C.-H. Liao, and S.-T. Po, “Decolorization of textile wastewater by photo-fenton oxidation technology,” Chemosphere, vol. 41, no. 8, pp. 1287–1294, 2000. View at Publisher · View at Google Scholar · View at Scopus
  96. I. Arslan, I. A. Balcioglu, and T. Tuhkanen, “Advanced oxidation of synthetic dyehouse effluent by O3, H2O2/O3 and H2O2/UV processes,” Environmental Technology, vol. 20, no. 9, pp. 921–931, 1999. View at Publisher · View at Google Scholar · View at Scopus
  97. N. H. Ince and D. T. Gönenç, “Treatability of a textile azo dye by UV/H2O2,” Environmental Technology, vol. 18, no. 2, pp. 179–185, 1997. View at Publisher · View at Google Scholar · View at Scopus
  98. K. Vinodgopal and P. V. Kamat, “Combine electrochemistry with photocatalysis,” CHEMTECH, vol. 26, no. 4, pp. 18–22, 1996. View at Google Scholar · View at Scopus
  99. C. Lizama, M. C. Yeber, J. Freer, J. Baeza, and H. D. Mansilla, “Reactive dyes decolouration by TiO2photo-assisted catalysis,” Water Science and Technology, vol. 44, no. 5, pp. 197–203, 2001. View at Google Scholar · View at Scopus
  100. F. Zhang, J. Zhao, T. Shen, H. Hidaka, E. Pelizzetti, and N. Serpone, “TiO2-assisted photodegradation of dye pollutants II. Adsorption and degradation kinetics of eosin in TiO2 dispersions under visible light irradiation,” Applied Catalysis B: Environmental, vol. 15, no. 1-2, pp. 147–156, 1998. View at Publisher ·View at Google Scholar · View at Scopus
  101. P. Qu, J. Zhao, T. Shen, and H. Hidaka, “TiO2-assisted photodegradation of dyes: a study of two competitive primary processes in the degradation of RB in an aqueous TiO2 colloidal solution,” Journal of Molecular Catalysis A: Chemical, vol. 129, no. 2-3, pp. 257–268, 1998. View at Publisher · View at Google Scholar · View at Scopus
  102. G. Liu, T. Wu, J. Zhao, H. Hidaka, and N. Serpone, “Photoassisted degradation of dye pollutants. 8. Irreversible degradation of alizarin red under visible light radiation in air-equilibrated aqueous TiO2dispersions,” Environmental Science and Technology, vol. 33, no. 12, pp. 2081–2087, 1999. View at Publisher · View at Google Scholar · View at Scopus
  103. J. Zhao, T. Wu, K. Wu, K. Oikawa, H. Hidaka, and N. Serpone, “Photoassisted degradation of dye pollutants. 3. Degradation of the cationic dye rhodamine B in aqueous anionic surfactant/TiO2dispersions under visible light irradiation: evidence for the need of substrate adsorption on TiO2particles,” Environmental Science and Technology, vol. 32, no. 16, pp. 2394–2400, 1998. View at Publisher ·View at Google Scholar · View at Scopus
  104. M. R. Hoffmann, S. T. Martin, W. Choi, and D. W. Bahnemann, “Environmental applications of semiconductor photocatalysis,” Chemical Reviews, vol. 95, no. 1, pp. 69–96, 1995. View at Publisher ·View at Google Scholar · View at Scopus
  105. I. K. Konstantinou and T. A. Albanis, “Photocatalytic transformation of pesticides in aqueous titanium dioxide suspensions using artificial and solar light: intermediates and degradation pathways,” Applied Catalysis B: Environmental, vol. 42, no. 4, pp. 319–335, 2003. View at Publisher · View at Google Scholar ·View at Scopus
  106. A. L. Linsebigler, G. Lu, and J. T. Yates, “Photocatalysis on TiO2 surfaces: principles, mechanisms, and selected results,” Chemical Reviews, vol. 95, no. 3, pp. 735–758, 1995. View at Publisher · View at Google Scholar · View at Scopus
  107. P. Reeves, R. Ohlhausen, D. Sloan et al., “Photocatalytic destruction of organic dyes in aqueous TiO2suspensions using concentrated simulated and natural solar energy,” Solar Energy, vol. 48, no. 6, pp. 413–420, 1992. View at Publisher · View at Google Scholar · View at Scopus
  108. X. Z. Li and M. Zhang, “Decolorization and biodegradability of dyeing wastewater treated by a TiO2-sensitized photo-oxidation process,” Water Science and Technology, vol. 34, no. 9, pp. 49–55, 1996. View at Publisher · View at Google Scholar · View at Scopus
  109. Y. Wang, “Solar photocatalytic degradation of eight commercial dyes in TiO2 suspension,” Water Research, vol. 34, no. 3, pp. 990–994, 2000. View at Publisher · View at Google Scholar · View at Scopus
  110. S. Sakthivel, B. Neppolian, M. V. Shankar, B. Arabindoo, M. Palanichamy, and V. Murugesan, “Solar photocatalytic degradation of azo dye: comparison of photocatalytic efficiency of ZnO and TiO2,” Solar Energy Materials and Solar Cells, vol. 77, no. 1, pp. 65–82, 2003. View at Publisher · View at Google Scholar · View at Scopus
  111. V. Augugliaro, C. Baiocchi, A. B. Prevot et al., “Azo-dyes photocatalytic degradation in aqueous suspension of TiO2 under solar irradiation,” Chemosphere, vol. 49, no. 10, pp. 1223–1230, 2002. View at Publisher · View at Google Scholar · View at Scopus
  112. M. Stylidi, D. I. Kondarides, and X. E. Verykios, “Pathways of solar light-induced photocatalytic degradation of azo dyes in aqueous TiO2 suspensions,” Applied Catalysis B: Environmental, vol. 40, no. 4, pp. 271–286, 2003. View at Publisher · View at Google Scholar · View at Scopus
  113. S. Gomes de Moraes, R. Sanches Freire, and N. Durán, “Degradation and toxicity reduction of textile effluent by combined photocatalytic and ozonation processes,” Chemosphere, vol. 40, no. 4, pp. 369–373, 2000. View at Publisher · View at Google Scholar · View at Scopus
  114. I. A. Balcioglu and I. Arslan, “Treatment of textile waste water by heterogenous photocatalytic oxidation processes,” Environmental Technology, vol. 18, no. 10, pp. 1053–1059, 1997. View at Publisher · View at Google Scholar
  115. X. Z. Li and F. B. Li, “Study of Au/Au3+-TiO2 photocatalysts toward visible photooxidation for water and wastewater treatment,” Environmental Science and Technology, vol. 35, no. 11, pp. 2381–2387, 2001. View at Publisher · View at Google Scholar · View at Scopus
  116. H. Zollinger, Properties of Organic Dyes and Pigments in Color Chemistry, VCH Publishers, New York, NY, USA, 1978.
  117. B. Kraeutler and A. J. Bard, “Heterogeneous photocatalytic decomposition of saturated carboxylic acids on titanium dioxide powder. Decarboxylative route to alkanes,” Journal of the American Chemical Society, vol. 100, no. 19, pp. 5985–5992, 1978. View at Publisher · View at Google Scholar
  118. W. Dunn, K. Wilboun, F. Fan, and A. Brad, “Heterogeneous photocatalytic oxidation of hydrocarbons on platinized TiO2 powders,” The Journal of Physical Chemistry A, vol. 84, no. 32, pp. 7–10, 1980. View at Google Scholar
  119. Y. Gao, M. Hu, and B. Mi, “Membrane surface modification with TiO2–graphene oxide for enhanced photocatalytic performance,” Journal of Membrane Science, vol. 455, pp. 349–356, 2014. View at Publisher· View at Google Scholar · View at Scopus
  120. C. Xu, A. Cui, Y. Xu, and X. Fu, “Graphene oxide-TiO2 composite filtration membranes and their potential application for water purification,” Carbon, vol. 62, pp. 465–471, 2013. View at Publisher · View at Google Scholar · View at Scopus
  121. A. L. Ahmad, M. A. Majid, and B. S. Ooi, “Functionalized PSf/SiO2 nanocomposite membrane for oil-in-water emulsion separation,” Desalination, vol. 268, no. 1–3, pp. 266–269, 2011. View at Publisher · View at Google Scholar · View at Scopus
  122. Z. Wang, H. Yu, J. Xia et al., “Novel GO-blended PVDF ultrafiltration membranes,” Desalination, vol. 299, pp. 50–54, 2012. View at Publisher · View at Google Scholar · View at Scopus
  123. C. P. Athanasekou, G. E. Romanos, F. K. Katsaros, K. Kordatos, V. Likodimos, and P. Falaras, “Very efficient composite titania membranes in hybrid ultrafiltration/photocatalysis water treatment processes,” Journal of Membrane Science, vol. 392-393, pp. 192–203, 2012. View at Publisher · View at Google Scholar · View at Scopus
  124. Y. S. Lin, “Microporous and dense inorganic membranes: current status and prospective,” Separation and Purification Technology, vol. 25, no. 1–3, pp. 39–55, 2001. View at Publisher · View at Google Scholar
  125. L. G. A. van de Water and T. Maschmeyer, “Mesoporous membranes—a brief overview of recent developments,” Topics in Catalysis, vol. 29, no. 1-2, pp. 67–77, 2004. View at Publisher · View at Google Scholar · View at Scopus
  126. W. S. Hummers and R. E. Offeman, “Preparation of graphitic oxide,” Journal of the Preparation of Graphitic Oxide, American Chemical Society, vol. 80, no. 6, p. 1339, 1958. View at Publisher · View at Google Scholar
  127. P. Gao, J. Liu, S. Lee, T. Zhang, and D. D. Sun, “High quality graphene oxide-CdS-Pt nanocomposites for efficient photocatalytic hydrogen evolution,” Journal of Materials Chemistry, vol. 22, no. 5, pp. 2292–2298, 2012. View at Publisher · View at Google Scholar · View at Scopus
  128. J. Liu, H. Jeong, K. Lee et al., “Reduction of functionalized graphite oxides by trioctylphosphine in non-polar organic solvents,” Carbon, vol. 48, no. 8, pp. 2282–2289, 2010. View at Publisher · View at Google Scholar · View at Scopus
  129. D. C. Marcano, D. V. Kosynkin, J. M. Berlin et al., “Improved synthesis of graphene oxide,” ACS Nano, vol. 4, no. 8, pp. 4806–4814, 2010. View at Publisher · View at Google Scholar · View at Scopus
  130. Y. Gao, M. Hu, and B. Mi, “Membrane surface modification with TiO2-graphene oxide for enhanced photocatalytic performance,” Journal of Membrane Science, vol. 455, pp. 349–356, 2014. View at Publisher· View at Google Scholar · View at Scopus
  131. C. P. Athanasekou, N. G. Moustakas, S. Morales-Torres et al., “Ceramic photocatalytic membranes for water filtration under UV and visible light,” Applied Catalysis B: Environmental, vol. 178, pp. 12–19, 2015.View at Publisher · View at Google Scholar
  132. M.-L. Luo, J.-Q. Zhao, W. Tang, and C.-S. Pu, “Hydrophilic modification of poly (ether sulfone) ultrafiltration membrane surface by self-assembly of TiO2 nanoparticles,” Applied Surface Science, vol. 249, no. 1–4, pp. 76–84, 2005. View at Publisher · View at Google Scholar · View at Scopus
  133. H.-B. Yao, L.-H. Wu, C.-H. Cui, H.-Y. Fang, and S.-H. Yu, “Direct fabrication of photoconductive patterns on LBL assembled graphene oxide/PDDA/titania hybrid films by photothermal and photocatalytic reduction,” Journal of Materials Chemistry, vol. 20, no. 25, pp. 5190–5195, 2010. View at Publisher · View at Google Scholar · View at Scopus
  134. G. Williams, B. Seger, and P. V. Kamt, “TiO2-graphene nanocomposites. UV-assisted photocatalytic reduction of graphene oxide,” ACS Nano, vol. 2, no. 7, pp. 1487–1491, 2008. View at Publisher · View at Google Scholar · View at Scopus
  135. S. Morales-Torres, L. M. Pastrana-Martínez, J. L. Figueiredo, J. L. Faria, and A. M. T. Silva, “Design of graphene-based TiO2 photocatalysts—a review,” Environmental Science and Pollution Research, vol. 19, no. 9, pp. 3676–3687, 2012. View at Publisher · View at Google Scholar · View at Scopus
  136. X. B. Ke, H. Y. Zhu, X. P. Gao, J. W. Liu, and Z. F. Zheng, “High-performance ceramic membranes with a separation layer of metal oxide nanofibers,” Advanced Materials, vol. 19, no. 6, pp. 785–790, 2007. View at Publisher · View at Google Scholar · View at Scopus
  137. X. Chen and S. S. Mao, “Titanium dioxide nanomaterials: synthesis, properties, modifications, and applications,” Chemical Reviews, vol. 107, no. 7, pp. 2891–2959, 2007. View at Publisher · View at Google Scholar · View at Scopus
  138. Y. Xu, H. Bai, G. Lu, C. Li, and G. Shi, “Flexible graphene films via the filtration of water-soluble noncovalent functionalized graphene sheets,” Journal of the American Chemical Society, vol. 130, no. 18, pp. 5856–5857, 2008. View at Publisher · View at Google Scholar · View at Scopus
  139. R. Bissessur, P. K. Y. Liu, W. White, and S. F. Scully, “Encapsulation of polyanilines into graphite oxide,” Langmuir, vol. 22, no. 4, pp. 1729–1734, 2006. View at Publisher · View at Google Scholar · View at Scopus
  140. C. Xu, Z. Chen, and X. Fu, “Graphene oxide-mediated formation of freestanding, thickness controllable metal oxide films,” Journal of Materials Chemistry, vol. 21, no. 34, pp. 12889–12893, 2011. View at Publisher · View at Google Scholar · View at Scopus
  141. H.-K. Jeong, P. L. Yun, R. J. W. E. Lahaye et al., “Evidence of graphitic AB stacking order of graphite oxides,” Journal of the American Chemical Society, vol. 130, no. 4, pp. 1362–1366, 2008. View at Publisher· View at Google Scholar · View at Scopus
  142. O. Akhavan, “Photocatalytic reduction of graphene oxides hybridized by ZnO nanoparticles in ethanol,” Carbon, vol. 49, no. 1, pp. 11–18, 2011. View at Publisher · View at Google Scholar · View at Scopus
  143. O. Akhavan and E. Ghaderi, “Photocatalytic reduction of graphene oxide nanosheets on TiO2 thin film for photoinactivation of bacteria in solar light irradiation,” Journal of Physical Chemistry C, vol. 113, no. 47, pp. 20214–20220, 2009. View at Publisher · View at Google Scholar · View at Scopus
  144. J. Zhao, W. Ren, and H.-M. Cheng, “Graphene sponge for efficient and repeatable adsorption and desorption of water contaminations,” Journal of Materials Chemistry, vol. 22, no. 38, pp. 20197–20202, 2012. View at Publisher · View at Google Scholar · View at Scopus
  145. C. Chen, W. Cai, M. Long et al., “Synthesis of visible-light responsive graphene oxide/TiO2 composites with p/n heterojunction,” ACS Nano, vol. 4, no. 11, pp. 6425–6432, 2010. View at Publisher · View at Google Scholar · View at Scopus
  146. J.-F. Li, Z.-L. Xu, H. Yang, L.-Y. Yu, and M. Liu, “Effect of TiO2 nanoparticles on the surface morphology and performance of microporous PES membrane,” Applied Surface Science, vol. 255, no. 9, pp. 4725–4732, 2009. View at Publisher · View at Google Scholar · View at Scopus
  147. N. A. A. Hamid, A. F. Ismail, T. Matsuura et al., “Morphological and separation performance study of polysulfone/titanium dioxide (PSF/TiO2) ultrafiltration membranes for humic acid removal,” Desalination, vol. 273, no. 1, pp. 85–92, 2011. View at Publisher · View at Google Scholar · View at Scopus
  148. E. Yuliwati, A. F. Ismail, T. Matsuura, M. A. Kassim, and M. S. Abdullah, “Characterization of surface-modified porous PVDF hollow fibers for refinery wastewater treatment using microscopic observation,” Desalination, vol. 283, pp. 206–213, 2011. View at Publisher · View at Google Scholar · View at Scopus
  149. H. Park and W. Choi, “Photocatalytic reactivities of nafion-coated TiO2 for the degradation of charged organic compounds under UV or visible light,” Journal of Physical Chemistry B, vol. 109, no. 23, pp. 11667–11674, 2005. View at Publisher · View at Google Scholar · View at Scopus
  150. J. S. Chen, C. Chen, J. Liu, R. Xu, S. Z. Qiao, and X. W. Lou, “Ellipsoidal hollow nanostructures assembled from anatase TiO2 nanosheets as a magnetically separable photocatalyst,” Chemical Communications, vol. 47, no. 9, pp. 2631–2633, 2011. View at Publisher · View at Google Scholar · View at Scopus
  151. H. B. Wu, H. H. Hng, and X. W. D. Lou, “Direct synthesis of anatase TiO2 nanowires with enhanced photocatalytic activity,” Advanced Materials, vol. 24, no. 19, pp. 2567–2571, 2012. View at Publisher ·View at Google Scholar · View at Scopus
  152. P. Gao, J. Liu, T. Zhang, D. D. Sun, and W. Ng, “Hierarchical TiO2/CdS ‘spindle-like’ composite with high photodegradation and antibacterial capability under visible light irradiation,” Journal of Hazardous Materials, vol. 229-230, pp. 209–216, 2012. View at Publisher · View at Google Scholar · View at Scopus
  153. J.-Y. Liao, B.-X. Lei, D.-B. Kuang, and C.-Y. Su, “Tri-functional hierarchical TiO2 spheres consisting of anatase nanorods and nanoparticles for high efficiency dye-sensitized solar cells,” Energy and Environmental Science, vol. 4, no. 10, pp. 4079–4085, 2011. View at Publisher · View at Google Scholar ·View at Scopus
  154. T. Zhang, J. Liu, and D. D. Sun, “A novel strategy to fabricate inorganic nanofibrous membranes for water treatment: use of functionalized graphene oxide as a cross linker,” RSC Advances, vol. 2, no. 12, pp. 5134–5137, 2012. View at Publisher · View at Google Scholar · View at Scopus
  155. C. Ruan, L. Zhang, Y. Qin et al., “Synthesis of porphyrin sensitized TiO2/graphene and its photocatalytic property under visible light,” Materials Letters, vol. 141, pp. 362–365, 2015. View at Publisher · View at Google Scholar · View at Scopus
  156. C. P. Athanasekou, S. Morales-Torres, V. Likodimos et al., “Prototype composite membranes of partially reduced graphene oxide/TiO2 for photocatalytic ultrafiltration water treatment under visible light,” Applied Catalysis B: Environmental, vol. 158-159, pp. 361–372, 2014. View at Publisher · View at Google Scholar
  157. C. Heitner-Wirguin, “Recent advances in perfluorinated ionomer membranes: structure, properties and applications,” Journal of Membrane Science, vol. 120, no. 1, pp. 1–33, 1996. View at Publisher · View at Google Scholar · View at Scopus
  158. S. J. Sondheimer, N. J. Bunce, M. E. Lemke, and C. A. Fyfe, “Acidity and catalytic activity of Nafion-H,” Macromolecules, vol. 19, no. 2, pp. 339–343, 1986. View at Publisher · View at Google Scholar · View at Scopus
  159. N. Zhang, Y. Zhang, and Y.-J. Xu, “Recent progress on graphene-based photocatalysts: current status and future perspectives,” Nanoscale, vol. 4, no. 19, pp. 5792–5813, 2012. View at Publisher · View at Google Scholar · View at Scopus
  160. J. Wiszniowski, D. Robert, J. Surmacz-Gorska, K. Miksch, and J.-V. Weber, “Photocatalytic decomposition of humic acids on TiO2 part I: discussion of adsorption and mechanism,” Journal of Photochemistry and Photobiology A: Chemistry, vol. 152, no. 1–3, pp. 267–273, 2002. View at Publisher ·View at Google Scholar · View at Scopus
  161. W. Yuan and A. L. Zydney, “Humic acid fouling during microfiltration,” Journal of Membrane Science, vol. 157, no. 1, pp. 1–12, 1999. View at Publisher · View at Google Scholar · View at Scopus
  162. A. W. Zularisam, A. F. Ismail, and R. Salim, “Behaviours of natural organic matter in membrane filtration for surface water treatment-a review,” Desalination, vol. 194, no. 1–3, pp. 211–231, 2006. View at Publisher · View at Google Scholar · View at Scopus
For further details log on website :
https://www.hindawi.com/journals/jchem/2017/6217987/

Advantages and Disadvantages of Fasting for Runners

Author BY   ANDREA CESPEDES  Food is fuel, especially for serious runners who need a lot of energy. It may seem counterintuiti...