Fabrication of a reusable magnetic multi-walled carbon nanotube–TiO2 nanocomposite by electrostatic adsorption: enhanced photodegradation of malachite green

Ghazale Daneshvar Tarighab, Farzaneh Shemirani*a and Nezam Seif Maz'haric
aDepartment of Analytical Chemistry, University College of Science, University of Tehran, P.O. Box 14155-6455, Tehran, Iran. E-mail: ghdaneshvartarigh@yahoo.com; Shemiran@khayam.ut.ac.ir; Fax: +98 21 66495291; Tel: +98 21 66495291
bNuclear Science and Technology Research Institute, End of North Karegar Ave., P. O. Box 1439951113, Tehran, Iran
cMaster of Science, University of Amirkabir, Tehran, Iran

Received 1st December 2014 , Accepted 9th April 2015

First published on 9th April 2015


Abstract

A simple, quick and efficient method for the fabrication of a magnetic multi-walled carbon nanotube–TiO2 (MMWCNT–TiO2) nanocomposite through electrostatic attraction was proposed as a novel method. First of all, TiO2 nanoparticles were synthesized by an atmospheric pressure chemical vapor synthesis (APCVS) method and characterized by scanning electron microscopy (SEM), X-ray diffraction (XRD) and UV-Vis diffuse reflectance spectroscopy (DRS). Next, the magnetic multi-walled carbon nanotube (MMWCNT) nanocomposite was prepared. Finally, TiO2 nanoparticles were coated onto the surface of the MMWCNT through electrostatic attraction in ethanol solution. The morphology and structure of the final composite were investigated by SEM, XRD, Energy dispersive X-ray spectrometry (EDS) and Fourier transform infrared spectrophotometry (FT-IR). Both TiO2 and MMWCNT–TiO2 were characterized by BET surface area/pore volume analysis. The photocatalytic function of the TiO2 and MMWCNT–TiO2 composite was validated for malachite green (MG) degradation under irradiation with ultra violet (UV) light. MMWCNT–TiO2, adsorbing MG, was easily separated from the aqueous solutions with the help of an external magnet; so, no filtration or centrifugation was necessary. The effects of pH, irradiation time, catalyst concentration, MG concentration, etc. on the photocatalytic activity were studied. The optimal conditions were an initial MG concentration of 20 mg L−1 at pH 5.0 with a catalyst concentration of 0.2 g L−1 under UV irradiation for 240 min with good recyclisation of the MMWCNT–TiO2 catalyst.


Introduction

Several chemical, physical and biological processes are presently utilized for wastewater treatment such as flocculation, ultrafiltration, adsorption, ozonation and chlorination.1,2 These procedures are not effective because they come out in solid waste, thus creating other environmental problems requiring further treatment. Therefore, it is necessary to find an effective method of wastewater treatment in order to remove hazardous dyes and organics from industry effluents.3 One of the methods of wastewater treatment containing dyes is their photocatalytic degradation in solutions illuminated with ultra violet (UV) irradiation, which contains a suitable photocatalyst. Among all the catalysts, titanium dioxide (TiO2) is found to be the most efficient catalyst for the photocatalytic degradation of a pollutant due to the faster electron transfer of molecular oxygen.4 TiO2 is considered as the most promising materials due to its relatively high activity, non-toxicity, chemical stability and low cost. However, the practical applications are still retarded by several limitations:5,6 (1) wide-band gap, limiting the photoactivation only in the ultraviolet light region (3.0 and 3.2 eV for the rutile and anatase phases, respectively); (2) high carrier-recombination probability; (3) poor adsorbing affinity toward organic molecules and (4) is the separation of micro- or nanosized photocatalysts from wastewater, which is an often a difficult and energy-intensive process. To overcome this shortcoming of TiO2, many researchers have tried to improve the photocatalytic activity of TiO2 under visible light. There has been considerable advancement in the production of novel functional materials by coupling TiO2 with other organic or inorganic materials.7 One of the most commonly employed methods for improving the photocatalytic activity is the combination of multi-wall carbon nanotubes (MWCNTs) with TiO2.8–11

MWCNTs have been extensively studied because they have many useful properties such as good electrical conductivity, nanosize absolute black, excellent mechanical properties, large surface area and high adsorption capacity.12 In addition, MWCNTs have a large electricity-storage capacity and, therefore, it may accept photon-excited electrons in mixtures or composites with titania, thus retarding or hindering recombination.7 Hence, the combination of MWCNTs with TiO2 can reduce charge recombination, enhance reactivity and photocatalytic ability of TiO2.13,14 On the other hand, a dispersion of TiO2 on the MWCNTs surface could create many active sites for the photocatalytic degradation. It was suggested that the photo generated charge carriers could transfer from the TiO2 to the CNTs and increase the photocatalytic activity of TiO2 because the excited electron in a conduction band of TiO2 might migrate into the CNTs and the recombination of electron–hole pairs decreased.15,16 To overcome problems of separation, photocatalysts with easily controllable magnetic properties that exhibit valuable advantages in environmental and biomedical applications have been developed.17 A number of binary magnetic photocatalysts consisting of TiO2-coated Fe3O4,18,19 TiO2–Fe3O4 hollow spheres,20 mesoporous mixed γ-Fe2O3–TiO2,21 and γ-Fe2O3–TiO2 Janus hollow bowls22 have been developed to realize photocatalyst recovery by taking advantage of the magnetic properties of γ-Fe2O3 and Fe3O4. Thus, TiO2 (the semiconductor photocatalyst) degrades organic contaminants; Fe3O4 provides magnetic properties for separation and recovery; and MWCNT provides an electron pathway to suppress charge recombination and enhance photocatalytic activity. These functions of TiO2–Fe3O4–MWCNT result in enhanced photocatalytic activity and decreased photocatalyst loss.

For the TiO2 preparation, various methods such as microemulsion,23 solvothermal,24 sol–gel,25 precipitation,26 combustion synthesis,27 electrochemical synthesis,28 chemical vapor synthesis (CVS),29 inert gas condensation (IGC),30 chemical vapor deposition (CVD),31 physical vapor deposition (PVD)32 etc. have been proposed. During the last decades, chemical vapor synthesis (CVS) has become more popular for developing high quality, ultra fine, unagglomerated, high purity and air-free nanocrystalline powders. The major limitation of the CVS process is an expensive vacuum apparatus.29 The atmospheric pressure chemical vapor synthesis (APCVS) route is a new, economic and affordable method for synthesizing high purity, well-structured and uniform nanoparticles and is developed by Rahiminezhad et al.33 In this procedure, some low-cost gasses and precursor will be brought out to a vertical or horizontal quartz tube at different temperatures, and chemical reactions take place at the atmospheric pressure. After synthesis process, nanoparticles will be collected on the surface of a cold trap by passing the exiting gas through a cold trap. As noted above, one of the significant advantages of this procedure is that APCVS operates at atmospheric pressure and thus expensive vacuum apparatus will not be required.

In the past decades, CNT–TiO2 hybrids have been largely fabricated by the sol–gel method,34 electro-spinning,13,35 electrophoretic deposition,36 and chemical vapor deposition (CVD).37 Besides the sol–gel method, TiO2 nanoparticles could be attached onto shortened CNTs by electrostatic attraction. To our knowledge, only one paper was reported about attaching CNTs–TiO2 by electrostatic attraction in water.38 This physical synthesis has some advantages in comparison with those syntheses of chemical methods. In sol–gel method,34 various reagents, too much time aging and high temperature needed. For electro-spinning method,13,35 specific instrument, high voltage, suitable viscosity solution, high temperature calcinations, many reagents and too much time processes needed. Electrophoretic deposition36 needs electric field and too much time for synthesis. At last, chemical vapor deposition (CVD),37 needs high temperature calcinations and it takes too much time. Our proposed method for fabrication of MMWCNT–TiO2 composite offers a simple, fast, efficient, high purity, low time and cost, no consumption of toxic solvents, no specific instrument, ability of synthesis in each laboratory and it was done at room temperature. For investigation of photocatalytic activity of this nanocomposite, malachite green was chosen as a sample model.

Malachite Green (MG), a basic dye has been widely applied for dyeing of leather, silk and wool and in distilleries.38 Its application extents in the aquaculture, commercial fish hatchery and animal farming as an antifungal therapeutic agent, while for human, it is used as antiseptic and fungicidal. Nonetheless, its oral consumption is carcinogenic. The available toxicological information, reveals that in the tissues of fish and mice MG easily reduces to persistable leuco-Malachite Green,39,40 which behaves as a tumor promoter. Thus, the detection of MG in fishes, animal, milk and other foodstuff designed for human consumption are of great alarm for the human health. Studies also confirm that the products formed after the degradations of MG are also not safe and have been carcinogenic potential.41 Therefore, it becomes necessary to transfer such a toxic dye from wastewater before it discharged into the aquatic environment.

Our present investigation has been aimed at three different aspects: first, synthesis of TiO2 nanoparticles by APCVS method, second, attaching of TiO2 to MMWCNT composite by electrostatic attraction in ethanol and thirdly evaluating its performance towards the photocatalytic degradation of MG. A comparative photocatalytic account of MG degradation was made by MMWCNT–TiO2 and TiO2. The preparation, morphology and structure, light absorption, adsorptivity to MG, and photocatalytic activity of the MMWCNT–TiO2 was investigated in detail.

Experimental

Chemicals

MWCNTs of 95% purity with outer diameter: 10–20 nm and inner diameter: 5–10 nm, length 10–30 μm, numbers of walls 3–15 were purchased from Plasma Chem. GmbH (Germany). MMWCNT was prepared as reported by our previously worked.42 Other reagents, including malachite green oxalate (MG oxalate), ammonium ferric sulfate, ammonium ferrous sulfate, titanium tetrachloride, hydrochloric acid, ammonia, were purchased from Merck. All reagents were of analytical grade and used without further purification. All aqueous solutions were prepared using the ultra-pure Milli-Q purification system.

Synthesis of TiO2 nanoparticle

Synthesis of TiO2 nanoparticles was carried by the APCVS method as reported by Rahiminezhad et al.33 Fig. 1 shows a schematic illustration of the experimental apparatus. A hot-wall, atmospheric pressure, a horizontal quartz reactor with the inner diameter of 80 mm and a length of 800 mm was used. It was run at temperature 800 °C and atmospherically pressure. High pure argon (purity > 99.999%) was employed as a carrier gas and split into three sections. Oxygen (purity > 99.999%) was used as an oxidizing gas. Titanium tetrachloride (TiCl4) (purity > 99.99%) was used as a precursor.
image file: c4ra15593a-f1.tif
Fig. 1 A diagram of APCVS configuration. (1) Oxygen (2) argon (3) needle valve (4) flow meter (5) water (6) TiCl4 (7) pressure indicator (8) furnace (9) quartz tube (10) cooling fan.

Before all runs, the internal surface of the quartz reactor was cleaned by laboratory alcohol (ethanol, 96 vol%). The quartz tube was placed inside the furnace and the furnace was brought to temperature 800 °C. Then, argon gas with a flow rate of 0.5 L min−1 and a pressure of 1 atm was fluxed directly into furnace quartz tubes for 10 min to create pure argon and clean the reactor internal surfaces from any impurities. Afterwards that, oxygen gas with a flow rate of 0.5 L min−1 and 1 atm pressure was introduced into a chamber quartz tube. Water vapor was introduced into the reactor by bubbling argon through a water bubbler without vacuum. The flow rate of water vapor was maintained at 0.5 L min−1. The liquid precursor (TiCl4) was evaporated in a vertical bubbler at oil bath at 800 °C. The precursor was introduced into a hot-wall tubular quartz reactor by bubbling argon through a precursor container without vacuum that was heated by an external resistance furnace. The effective heated region was 20 cm in the middle of quartz reactor. In the reactor, the precursor was oxidized to give TiO2 monomers. The generated monomers by passing through the reactor underwent coalescence, coagulation, agglomeration and sintering, ending up as titania nanoparticles. The produced nanoparticles were collected in the cold trap which consisted of a chamber, inlet and outlet lines and was kept in an ice-water bathroom.

Synthesis of MMWCNT–TiO2

MMWCNT (1.0 g) was dispersed in 20 mL ethanol in an ultrasonic bath for 1 hour. TiO2 nanoparticle (0.2 g) also was dispersed in 10 mL ethanol separately (the ratio of MMWCNT–TiO2 has been optimized before the final synthesis process). After about 1 h, the MMWCNT and TiO2 suspensions were mixed ultrasonically for 30 min. Then, the mixture was stirred using a glassware stirrer (1200 rpm) for eight hours. The MMWCNT–TiO2 composite was isolated from the mixture with the help of a permanent magnet. Separated MMWCNT were washed three times with deionized water followed by ethanol. Finally, MMWCNT was dried at 110 °C for one day. Once the MMWCNT and TiO2 nanoparticle suspensions are mixed, TiO2 nanoparticles instantaneously attach onto CNTs due to electrostatic attractive force between them. The suspension also shows a color change from black (for CNT suspension) and white (for TiO2 suspension) to be dark gray. For attaching TiO2 nanoparticles onto CNT surfaces the point of zero charges (pzc) should be considered. The pHpzc of as-produced CNTs is in the range of pH = 4–6. Numerous research works demonstrate that the pHpzc of acid-modified CNTs and oxidant-modified CNTs are lower than those of as-produced CNTs.43 When pH is <4, the pzc of the original CNT suspension is positive. When pH is >3, the pzc is negative. The TiO2 suspension has pHpzc at 6.5. Under this pH, the TiO2 nanoparticles are positively charged. Above this pH, the TiO2 nanoparticles are negatively charged. Very desirably, the two suspensions have opposite pzc at pH < 6.5. If CNT and TiO2 suspensions with opposite pzc are mixed under a suitable pH, TiO2 nanoparticles and CNTs should attract each other. Attaching TiO2 to MMWCNT was done between pH = 3–7. As a result, pH 5 is the desired condition for TiO2 attachment onto CNTs.

Photocatalytic test

The photocatalytic activity of TiO2–MMWCNT composite was tested in the MG degradation in aqueous solutions under UV light radiation. The experiments were carried out in an open wide glass photochemical reactor charged with a mechanical stirrer. A 250 W high-pressure Hg lamp was placed approximately 20 cm from the reactor, and the wavelength of the emission light was in the range of 200–400 nm (Fig. 2). In a typical experiment; 50 mL of aqueous MG solution with a concentration (C0) of 50 mg L−1 was mixed with the TiO2–MMWCNT composite (0.2 g L−1), and continually stirred in the dark for 50 min in order to establish the adsorption–desorption equilibrium. Prior to turning on the light, the first sample was taken out at the end of the dark adsorption period to confirm the MG concentration in the solution, which was considered as the initial concentration (C0) after dark adsorption. The concentration of the solution was determined, and was taken as the initial concentration (C0) of the MG solution. MMWCNT–TiO2 was easily separated from the aqueous solutions with the help of a strong permanent magnet with 1 Tesla magnetic fields include the dimensions of the magnet employed (10 × 3 × 2 mm3); so, no filtration or centrifugation was necessary. Subsequently, the cleaning solutions were analyzed regularly by UV-vis spectroscopy (Perkin-Elmer Lambda 20) at scheduled irradiation times (λmax: 617 nm). The photocatalytic degradation efficiency (η%) could be calculated as the following equation.
 
image file: c4ra15593a-t1.tif(1)
where C0 and C are the concentration of MG at t = 0 and t, respectively.

image file: c4ra15593a-f2.tif
Fig. 2 A schematic illustration of MMWCNT–TiO2 synthesis and photodegradation of MG.

Characterization

X-ray powder diffraction (XRD) measurements were performed (STADI-MP from a STOE company) with monochromatized Cu Kα radiation (λ = 0.15418 nm). The 2θ scanning angle range was 0–80θ with a step of 0.02°/0.2 s. The Fourier transform infrared spectrophotometer (FT-IR) (4000–400 cm−1) were recorded using KBr pellets (Bruker Vector 22 FT-IR spectrophotometer) with 2 cm−1 resolution. UV-vis diffuse reflectance spectra (DRS) were recorded (UV-3100 spectrophotometer made by Shimadzu Corporation) with spherical diffuse reflectance accessory, using BaSO4 as a reference. The MMWCNTs were characterized by scanning electronic microscope (SEM) (Philips XL30, Eindhoven), with gold coating. Energy dispersive X-ray spectrometry (EDS) was performed by Oxford ED-2000. Transmission electron microscopic (TEM) image was recorded in a (ZEOL ZEM 2010 TEM) instrument with an accelerating voltage of 200 kV. Specific surface areas were measured by using a NovaWin2 analytical system made by Quantachrome Corporation and were calculated by the Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) methods.

Results and discussion

Characterization results

XRD analysis. XRD was used to determine the phase composition of the TiO2 nanoparticles. Fig. 3 shows the XRD patterns of TiO2 nanoparticle. The pattern of the TiO2 sample exhibits several characteristic peaks at 2θ = 25.2°, 37.7°, 48° and 54°, in agreement with anatase phase which was corresponding well with the JCPDS, no. 01-084-1286 data file. Average crystallite size of TiO2 was estimated according to Scherrer's equation, d = λ/β[thin space (1/6-em)]cos[thin space (1/6-em)]θ, where d is the average crystallite size (nm), λ is the wavelength of the Cu Kα applied (λ = 0.154 nm), β is the peak width at half-maximum and k is the constant usually applied as 0.9. As the precursor temperature of 800 °C, the average nanoparticle size was 50 nm.
image file: c4ra15593a-f3.tif
Fig. 3 XRD patterns of TiO2 nanoparticle by APCVS method.

The XRD of the bonding TiO2–MMWCNT composite is shown in Fig. 4 in order to characterize the crystalline structure of the samples. As can be seen, the peaks appeared at 2θ values of 25.32°, 48.07° and 53.95° could be indexed to the anatase TiO2, which was corresponding well with the JCPDS, no. 01-084-1286 data file. It is worth noticing that the characterized peaks of CNT and anatase are both present in the bonding TiO2–MWCNT composite simultaneously, and no other diffraction peaks can be observed. However, the main diffraction peaks of CNT (2θ = 25.4 in comparison with the XRD pattern of CNT reported in our previous work42) are not observed clearly in this composite, this is because the reflection of CNT is overlapped by the reflection of anatase. The XRD results suggest the formation of TiO2–MWCNT composite. In addition, the dominant peaks appeared at 2θ values of 30.17°, 35.45°, 43.25°, 56.78° and 62.73°, all of these diffraction peaks rest with magnetite of Fe3O4 nanoparticles (JCPDS, no. 00-003-0863).


image file: c4ra15593a-f4.tif
Fig. 4 XRD patterns of MMWCNT–TiO2 nanocomposite.
Microscopic analysis. To measure the TiO2 nanoparticle's shape, size and distribution, SEM was used. The morphology, coagulation and agglomeration of the prepared titania nanoparticles were observed by SEM. Fig. 5a shows the SEM images of TiO2 nanoparticles produced at 800 °C. As it can be seen, measurements are consistent with the results of XRD analysis and their distribution was relatively narrow, which indicates the uniformity and homogeneity of the product. The shape of the particles is generally spherical and quite similar to each other, and the size of the particles varies in the range of 49–70 nm. Fig. 5b shows the TEM image of TiO2 nanoparticles. This image shows that titania nanoparticles are perfectly round.
image file: c4ra15593a-f5.tif
Fig. 5 (a) SEM (b) TEM images of TiO2 nanoparticle.

The morphology of MMWCNT–TiO2 was investigated by SEM and TEM. As shown in Fig. 6a and b, the SEM and TEM images revealed the presence of two types of particles (TiO2 and Fe3O4) with similar sizes, transparencies, and distributions across the CNT support. In the MMWCNT–TiO2 photocatalyst, Fe3O4 NPs were magnetically aggregated in certain domains, while TiO2 NPs with were spread homogeneously on the CNT surface. The Fe3O4 crystallites were ultrafine, connecting tightly to one another to form spheres. The EDS (Fig. 6c) of the illuminating electron beams on the obtained MMWCNT–TiO2 composites revealed the existence of Fe, C, Ti and O elements, further confirming the successful modification of MMWCNTs with TiO2. The quantitative analysis gives weight percentages of Fe (21.71%), C (44.58%), Ti (14.18%) and O (19.53%). Due to the results of the EDS, the optimized concentration of MMWCNT and TiO2 obtained 8.33 g L−1 and 1.66 g L−1 respectively.


image file: c4ra15593a-f6.tif
Fig. 6 (a) SEM image (b) TEM image and (c) EDS spectrum of MMWCNT–TiO2.
Spectroscopic analysis. FT-IR spectrum of TiO2–MMWCNT was shown in Fig. 7. A peak near 583 cm−1 which belonged to the characterized absorption peaks of Ti–O and Fe–O. The transmittance band at 717 cm−1 is considered to be the signature marker of TiO2 and is symbolized to be the stretching vibration of Ti–O–Ti. The sharp peak at 1637 cm−1 was a resultant of a bending vibration of a coordinated hydroxyl group as well as Ti. The minor band at 1042 cm−1 in TiO2–MMWCNT can be ascribed to the stretching vibration of C[double bond, length as m-dash]O. The vibration bands at 2851 and 2930 cm−1 were assigned to the stretching vibration of –CH. Moreover; the absorption peak at 1391 cm−1 was ascribed to the O–H bond vibration of the surface-adsorbed species of particles (Ti–OH). Additionally, the broadband near 3421 cm−1 is assigned not only to the presence of hydroxyl groups of TiO2 but also to a strong interaction through hydrogen bonding between the hydroxyl groups on the surface of titanium dioxide.
image file: c4ra15593a-f7.tif
Fig. 7 FT-IR spectrum of MMWCNT–TiO2.

DRS technique is a useful technique to characterize the optical absorption properties of nanoparticles. Fig. 8 shows the action spectra of TiO2, MMWCNT, and MMWCNT–TiO2. For TiO2, an absorption edge rising steeply toward the UV below 387 nm can be attributed to band-gap excitation of anatase (∼3.2 eV). TiO2, which consists of the anatase phase only, exhibits no visible-light absorption and a threshold wavelength around 400 nm. According to the Planck's law and some further calculation, the band gap can be estimated by using Eg = hc/λg = 1239.85/λg (eV) where h is Planck's constant (4.13566733 × 10−15 eV s); c is the speed of light (2.99792458 × 1017 nm s−1), Eg was the energy gap and λg was absorption threshold. The action regions of MMWCNT and MMWCNT–TiO2 extended significantly into the visible-light region, suggesting that MMWCNT–TiO2 can be activated by visible-light. From the calculation, the band gap value for TiO2, MMWCNT and MMWCNT–TiO2 was obtained 3.2 eV, 2.3 eV and 1.9 eV, respectively. These observations illustrate that the presence of MMWCNT–TiO2 not only enhances the visible-light photocatalytic performance in organic dye degradation over TiO2 or MMWCNT, but also exhibits excellent photostability in the visible-light range.


image file: c4ra15593a-f8.tif
Fig. 8 DRS spectrum of TiO2, MMWCNT and MMWCNT–TiO2.

Photodegradation of MG

Prior to the photocatalytic activity tests, measurements of BET surface area, an important factor with respect to photocatalytic behavior, were obtained for TiO2, MMWCNT and TiO2–MMWCNT.44 The calculated BET surface areas of MMWCNT and TiO2–MMWCNT were 180.0 and 173.1 m2 g−1, respectively, which were notably higher than that of TiO2 (65.66 m2 g−1). The pore volume analysis showed that the total volume of MMWCNT was 0.708 cm3 g−1, and the total volume of TiO2–MMWCNT was even decreased (0.625 cm3 g−1) (Table 1). The average pore sizes calculated by the BJH method in TiO2, MMWCNT and TiO2–MMWCNT in the order 2.74, 2.73, and 2.74 nm, respectively, indicating that the mesoporous structures of the MMWCNT and TiO2–MMWCNT composites were retained. As a result, MMWCNT and TiO2–MMWCNT materials with large specific surface areas and ideal pore size distributions have shown to possess desirable photocatalytic properties for the removal of organic pollutants from wastewater.
Table 1 Physicochemical properties of TiO2, MMWCNT and MMWCNT–TiO2
Samples SBETa (m2 g−1) Vtotb (cm3 g−1) Ave. pore diameterc (nm)
a The calculated BET surface area.b The total pore volume analysis.c Average pore diameter.
TiO2 65.66 0.083 2.74
MMWCNT 180.0 0.708 2.73
MMWCNT–TiO2 173.1 0.625 2.74


Fig. 9 compares the photocatalytic degradation of MG in the presence of the neat TiO2 powder and TiO2–MMWCNT under irradiation of UV light. It is obvious that TiO2–MMWCNT present a high photocatalytic activity compared to the neat TiO2 powder. This behavior may be attributed to the positive effect of MWCNTs: (1) acting as a dispersing agent, the MWCNTs prevent TiO2 from agglomeration and (2) acting as an adsorbent, the adsorption efficiency of TiO2–MMWCNTs is better than that of neat TiO2. It has been confirmed that MWCNTs in TiO2–MMWCNTs are useful to absorb the MG and transfer the compound to the surface of TiO2 (3) acting as a photosensitizer, there is a synergetic effect between MWCNTs and TiO2. The photo-induced electrons in MWCNTs may trigger the formation of radicals in TiO2 (superoxide radical ion and/or hydroxyl radical), which are responsible for the degradation of the organic compound.34


image file: c4ra15593a-f9.tif
Fig. 9 Comparison of photocatalytic behavior of TiO2 and MMWCNT–TiO2 composite on MG degradation.

The mechanism of photocatalytic can be summarized as follows: TiO2 nanoparticles on photo-irradiation by UV light can be excited leading to the generation of electrons and holes. These excited electrons can then migrate into the conduction band of MWCNT since its work function is higher than TiO2 nanoparticles.45 This migration is thermodynamically favorable since the conduction band and valence band of TiO2 are above than that of MWCNT. Further, MWCNT acts as a good electron acceptor for the electrons excited by UV light in the conduction band of TiO2. Thus, the whole lifetime of the photo-generated electrons and holes is prolonged in the electron transfer process retarding the recombination and inducing higher quantum efficiency. It is hypothesized that the photo-generated electrons in MWCNT might react with the dissolved oxygen molecules, thereby producing oxygen peroxide radicals O2˙. Holes generated in TiO2 nanoparticles may react with the OH which is obtained from water molecules to form hydroxyl radicals OH˙. MG can then be photocatalytically degraded by both oxygen peroxide radicals O2˙and hydroxyl radicals OH˙.

 
TiO2 + → TiO2 (h+, e) (2)
 
TiO2 (e) + MWCNTs → MWCNTs (e) + TiO2 (h+, e) (3)
 
TiO2 (e) + O2 → TiO2 + O˙ (4)
 
MWCNTs (e) + O2 → MWCNTs + O˙ (5)
 
TiO2 (h+) + H2O → TiO2 + OH˙ + H+ (6)
 
(TiO2)e + MG → (TiO2)e + MG+˙ (7)
 
MG+˙ + O2 → “degraded products” (8)
 
MG+˙ + HO˙ → “degraded products” (9)

The effect of reaction variables such as recycling test, pH of the solution, initial concentration of MG, catalyst concentration, irradiation time, etc. on the degradation efficiency was studied and the results are delineated below.

To investigate whether TiO2–MMWCNT composites can be recycled and reused for MG degradation, the materials were regenerated by washing with ethanol three times. Noticeably, the catalytic efficiency of the composite photocatalyst was still higher than 90% after be used for five cycles under 240 min UV irradiation. From the experimental results, since the photocatalyst was collected by a magnet, we can infer that the TiO2 in the composite catalyst did not release or dissolve into solution (Fig. 10a).


image file: c4ra15593a-f10.tif
Fig. 10 Investigation of recycling test (a), the influence of pH (b) photocatalyst amount (c) MG initial concentration (d) and irradiation time (e) on degradation efficiency of MG using TiO2 and MMWCNT–TiO2 photocatalysts (for (a)–(d) irradiation time = 240 min).

The effect of pH on the degradation of MG was investigated by keeping all other experimental conditions constant and varying the initial pH of the MG solution from 3 to 7. Diluted hydrochloride acid solution or sodium hydroxide solution was used to adjust the pH value when necessary. The experimental results reveal that the degree of degradation increases with pH value up to 5, beyond which the photodegradation efficiency starts to decrease, indicating an optimum pH of approximately 5.0 for best performance. This is due to the amount of hydroxyl radical formation on TiO2 is influenced by the solution pH. The adsorption of MG is quite well at pH near the pzc of photocatalyst (Fig. 10b).

Experiments were performed to study the variations in the rate of degradation at different catalyst concentration ranging from 5 to 15 mg. It is observed that the degradation efficiency increases sharply with the catalyst concentration up to 10 mg. This is due to an increased number of available adsorption and catalytic sites on the surface of the MMWCNT–TiO2 composite catalyst. A further increase in catalyst concentration, however, may cause light scattering and screening effect and thus reduce the specific activity of the catalyst. The result indicates that the optimum catalyst concentration for degradation of MG is 10 mg (Fig. 10c).

After optimizing the photocatalyst dosage, the effect of initial dye concentration ranging from 10 to 70 mg L−1 on the photodegradation of MG was investigated. It has been observed that the rate of photodegradation increased with increasing in dye up to 20 mg L−1. This may be due to the fact that as the dye concentration was increased, more dye molecules were available for consecutive degradation. The rate of photodegradation was found to decrease with a further increase in dye concentration, i.e. above 20 mg L−1. The reason for this decrease is attributed to the shielding effect of dye at high concentration that retards the penetration of light to the dye molecules deposited over the catalyst surface (Fig. 10d).

The effect of irradiation time on the photocatalytic degradation of MG from its aqueous solution was investigated from 0 to 240 min, at 20 mg L−1 MG concentration, 10 mg catalyst concentration and pH = 5.0. The obtained results are shown in Fig. 10e. The TiO2–MMWCNT composite can achieve above 90% MG removal for 240 min while neat TiO2 achieved only 65% MG removal for the same irradiation time. The addition of MMWCNT can enhance photoactivity of TiO2 remarkably. For the neat TiO2 or TiO2–MMWCNT composite catalysts, the photodegradation efficiency increases with time, up to 240 minute. This indicates that photocatalytic degradation of MG with catalysts for 240 min is the optimum irradiation time.

Conclusion

We have shown, for the first time, the successful fabrication of a recoverable and effective TiO2–MMWCNT photocatalyst by electrostatic attraction from MMWCNT dispersion onto TiO2 suspension in ethanol. For this aim, firstly, the anatase TiO2 nanoparticle was synthesized by APCVS method and qualified. Secondly, MMWCNT was prepared and mixed with TiO2 in ethanol to form MMWCNT–TiO2 composite and characterized. Finally, the UV lights photocatalytic activity of TiO2–MMWCNT and TiO2 as the photocatalysts for degradation of MG were investigated. All the resulting of MMWCNT–TiO2 exhibited significantly enhanced photocatalytic activities compared with the pure TiO2 nanoparticle. The photocatalytic stability of MMWCNT–TiO2 remained almost unchanged after recycling five cycles. In addition, this composite could be easily recovered from water using an external magnet. Because of these attractive, stable and recyclable features, the as-prepared composites may have been promising applications in the photodegradation of organic compounds in aqueous solution under the irradiation of UV light.

Acknowledgements

The financial support from the University of Tehran is gratefully acknowledged.

References

  1. S. Gül and O. Ozcan-Yıldırım, Chem. Eng. J., 2009, 155, 684–690 CrossRef PubMed.
  2. F. Sayilkan, M. Asiltürk, P. Tatar, N. Kiraz, E. Arpaç and H. Sayilkan, J. Hazard. Mater., 2007, 148, 735–744 CrossRef CAS PubMed.
  3. H. Wang, J. Li, X. Quan, Y. Wu, G. Li and F. Wang, J. Hazard. Mater., 2007, 141, 336–343 CrossRef PubMed.
  4. A. Fijishima, T. N. Rao, K. Try and D. A. Tryk, J. Photochem. Photobiol., C, 2001, 1, 1–21 CrossRef.
  5. R. Asaki, T. Morikawa, T. Ohwaki, K. Aoki and Y. Taga, Science, 2001, 293, 269–271 CrossRef PubMed.
  6. X. B. Chen and S. S. Mao, Chem. Rev., 2007, 107, 2891–2959 CrossRef PubMed.
  7. T. H. T. Vu, T. T. T. Nguyen, P. H. T. Nguyen, M. H. Do, H. T. Au and T. B. Nguyen, et al., Mater. Res. Bull., 2012, 47, 308–314 CrossRef PubMed.
  8. A. de Morais, L. M. D. Loiola, J. E. Benedetti, A. S. Gonçalves, C. A. O. Avellaneda and J. H. Clerici, et al., J. Photochem. Photobiol., A, 2013, 251, 78–84 CrossRef PubMed.
  9. G. Oza, S. Pandey, A. Gupta, S. Shinde, A. Mewada and P. Jagadale, et al., Mater. Sci. Eng., C, 2013, 33, 4392–4400 CrossRef PubMed.
  10. X. H. Xia, Z.-J. Jia, Y. Yu, Y. Liang, Z. Wang and L.-L. Ma, Carbon, 2007, 45, 717–721 CrossRef PubMed.
  11. M. Chen, F. Zhang and W. Oh, Carbon, 2009, 47, 2943 CrossRef PubMed.
  12. B.-T. Zhang, X. Zheng, H.-F. Li and J.-M. Lin, Anal. Chim. Acta, 2013, 784, 1–17 CrossRef PubMed.
  13. S. Aryal, C. K. Kim, K.-W. Kim, M. S. Khil and H. Y. Kim, Mater. Sci. Eng., C, 2008, 28, 75–79 CrossRef PubMed.
  14. A. Jitianu, T. Cacciaguerra, R. Benoit, S. Delpeux, F. Béguin and S. Bonnamy, Carbon, 2004, 42, 1147–1151 CrossRef PubMed.
  15. Y. Yu, J. C. Yu, J. G. Yu, Y. C. Kwok, Y. K. Che and J. C. Zhao, et al., Appl. Catal., A, 2005, 61, 186–196 CrossRef PubMed.
  16. H. Wang, X. Quan, H. Yu and S. Chen, Carbon, 2008, 46, 1126–1132 CrossRef PubMed.
  17. V. Polshettiwar, R. Luque, A. Fihri, H. Zhu, M. Bouhrara and J. M. Basset, Chem. Rev., 2011, 111, 3036–3075 CrossRef CAS PubMed.
  18. S. Anandan, G. J. Lee, S. H. Hsieh, M. Ashokkumar and J. J. Wu, Ind. Eng. Chem. Res., 2011, 50, 7874–7881 CrossRef CAS.
  19. C. F. Chang and C. Y. Man, Ind. Eng. Chem. Res., 2011, 50, 11620–11627 CrossRef CAS.
  20. S. Xuan, W. Jiang, X. Gong, Y. Hu and Z. Chen, J. Phys. Chem. C, 2009, 113, 553–558 CAS.
  21. B. Palanisamy, C. M. Babu, B. Sundaravel, S. Anandan and V. Murugesan, J. Hazard. Mater., 2013, 252, 233–242 CrossRef PubMed.
  22. F. Mou, L. Xu, H. Ma, J. Guan, D. R. Chen and S. Wang, Nanoscale, 2012, 4, 4650–4657 RSC.
  23. F. A. Deorsola and D. Vallauri, Powder Technol., 2009, 190, 304–309 CrossRef CAS PubMed.
  24. C. Wang, Z. X. Deng, G. Zhang, S. Fan and Y. Li, Powder Technol., 2002, 125, 39–44 CrossRef CAS.
  25. L. Mao, Q. Li, H. Dang and Z. Zhang, Mater. Res. Bull., 2005, 40, 201–208 CrossRef CAS PubMed.
  26. S. Mayadevi, S. S. Kulkarni, A. J. Patil, M. H. Shinde, H. S. Potdar and S. B. Deshpande, et al., J. Mater. Sci., 2000, 35, 3943–3949 CrossRef CAS.
  27. Y. Kitamura, N. Okinaka, T. Shibayama, O. P. Mahaney, D. Kusano and B. Ohtani, et al., Powder Technol., 2007, 176, 93–98 CrossRef CAS PubMed.
  28. H. Ishihara, J. P. Bock, R. Sharma, F. Hardcastle, G. K. Kannarpady and M. K. Mazumder, et al., Chem. Phys. Lett., 2010, 489, 81–85 CrossRef CAS PubMed.
  29. I. Ahmad and S. S. Bhattacharya, J. Phys. D: Appl. Phys., 2008, 41, 155313–155319 CrossRef.
  30. A. Simchi, R. Ahmadi, S. M. Seyed Reyhani and A. Mahdavi, Mater. Des., 2007, 28, 850–856 CrossRef CAS PubMed.
  31. C. S. Kim, K. Nakaso, B. Xia, K. Okuyama and M. Shimada, Aerosol Sci. Technol., 2005, 39, 104–112 CrossRef CAS.
  32. C. J. Tavares, S. M. Marques, T. Viseu, V. Teixeira, J. O. Carneiro and E. Alves, et al., J. Appl. Phys., 2009, 106, 113535–113542 CrossRef PubMed.
  33. M. Rahiminezhad-soltani, K. Saberyan, F. Shahri and A. Simchi, Powder Technol., 2011, 209, 15–24 CrossRef PubMed.
  34. H. Zhou, C. Zhang, X. Wang, H. Li and Z. Du, Synth. Met., 2011, 161, 2199–2205 CrossRef CAS PubMed.
  35. J. Cho, S. Schaab, J. A. Roether and A. R. Boccaccini, J. Nanopart. Res., 2007, 10, 99–105 CrossRef.
  36. M. Corrias, B. Caussat, A. Ayral, J. Durand, Y. Kihn and P. Kalck, et al., Chem. Eng. Sci., 2003, 58, 4475–4482 CrossRef CAS.
  37. H. Dong and K. Lu, Int. J. Appl. Ceram. Technol., 2009, 6, 216–222 CrossRef PubMed.
  38. J. Zhang, Y. Li, C. Zhang and Y. Jing, J. Hazard. Mater., 2008, 150, 774–782 CrossRef CAS PubMed.
  39. A. Mittal, J. Hazard. Mater., 2006, 113, 196–202 CrossRef PubMed.
  40. S. Nethaji, A. Sivasamy, G. Thennarasu and S. Saravanan, J. Hazard. Mater., 2010, 181, 271–280 CrossRef CAS PubMed.
  41. V. K. Gupta, I. Ali and V. K. Saini, Ind. Eng. Chem. Res., 2004, 43, 1740–1747 CrossRef CAS.
  42. G. Daneshvar and F. Shemirani, Talanta, 2013, 115, 744–750 CrossRef PubMed.
  43. X. Ren, C. Chen, M. Nagatsu and X. Wang, Chem. Eng. J., 2011, 170, 395–410 CrossRef CAS PubMed.
  44. S. Brunauer, P. H. Emmett and E. Teller, J. Am. Chem. Soc., 1938, 60, 309–319 CrossRef CAS.
  45. S. T. Martin, M. R. Hoffmann, W. Choi and D. W. Bahnemannt, Chem. Rev., 1995, 95, 69–96 CrossRef.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.