Next Article in Journal
Ultrasonic Monitoring of Biocatalysis in Solutions and Complex Dispersions
Next Article in Special Issue
Aerobic Oxidation of Benzyl Alcohol on a Strontium-Based Gold Material: Remarkable Intrinsic Basicity and Reusable Catalyst
Previous Article in Journal
Green Synthesis of Ultraviolet Absorber 2-Ethylhexyl Salicylate: Experimental Design and Artificial Neural Network Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fe-Doped TiO2 Supported on HY Zeolite for Solar Photocatalytic Treatment of Dye Pollutants

1
Laboratoire de Physico-Chimie des Matériaux et Catalyse, Faculté des Sciences Exactes, Université de Bejaia, Bejaia 06000, Algeria
2
CNR-Istituto di Scienze e Tecnologie Molecolari, Via C. Golgi, 19, 20133 Milano, Italy
3
Institut de Chimie et Procédés pour l’Energie, l’Environnement et la Santé (ICPEES), CNRS-UMR7515-University of Strasbourg, Saint-Avold Antenna, Université de Lorraine, 12 rue Victor Demange, 57500 Saint-Avold, France
*
Author to whom correspondence should be addressed.
Catalysts 2017, 7(11), 344; https://doi.org/10.3390/catal7110344
Submission received: 24 October 2017 / Revised: 9 November 2017 / Accepted: 10 November 2017 / Published: 14 November 2017

Abstract

:
Zeolite-supported TiO2 and Fe-doped TiO2/zeolite photocatalysts were synthesized, aiming at improving the adsorption properties of the microporous support towards polluting dyes and the photocatalytic performance of TiO2 in the treatment of the adsorbed organics. The TiO2/HY zeolite catalyst exhibited the best performance in the photocatalytic degradation of methylene blue, MB, over 10 wt% TiO2/HY under UV light at 254 nm. The addition of Fe species in the synthesis mixture led to Fe-TiO2/HY catalyst. The combination of adequate zeolite, good titanium dioxide dispersion, and Fe doping led to a remarkable performance in the degradation of the model dye. Over a 10 wt% Fe-doped TiO2/HY catalyst a total removal of MB (>98%) was achieved after 60 min under very mild conditions and simple visible light irradiation.

Graphical Abstract

1. Introduction

Dyes are extensively used in the manufacture of textiles, leather, printing, food, plastic, and cosmetics to give color to commercial products [1]. The presence of dyes in industrial waste water, even at low concentration, is highly visible and undesirable. They present mutagenic and carcinogenic properties and can cause significant damage to both aquatic life and human health, such as dysfunction of the kidneys, reproductive system, liver, brain, and central nervous system [2]. Methylene blue is one of the most widely used dyes. Although methylene blue is not highly hazardous in moderate doses (lethal dose LD50 oral rat = 1180 mg kg−1), it can cause some harmful effects [3]. Acute exposure to methylene blue causes increased heart rate, vomiting, diarrhea, nausea, shock, Heinz body formation, and tissue necrosis in humans [4,5].
With respect to the above critical negative effects, different strategies have been developed aiming at an efficient removal of this kind of pollutant from water and the human environment. In recent years, new processes based on the use of heterogeneous photocatalysts have attracted ever-increasing attention and titanium dioxide, TiO2, proved to be one of the most appropriate semiconductor materials to be employed as a photocatalyst. However, two main fundamental disadvantages limit the application of “bare” TiO2 alone in photocatalytic reactors: (i) the relatively wide band gap of TiO2 (3.2 eV), which absorbs only 3–5% energy of the solar spectrum and restricts its applications to UV excitation sources [6]; and (ii), when used in slurry or suspension mode, a post-treatment recovery step for the TiO2 catalyst is necessary, which is normally difficult, energy intensive, and expensive [7]. Moreover, TiO2 powder is prone to form aggregates, especially at high concentrations, and this is a major limitation to the application in continuous flow systems [8]. This also leads to undesired scattering of incident UV light by the suspended particles [9,10].
Therefore, several approaches have been explored in order to extend the optical absorption of TiO2 into the visible light region and, thus, to enhance its photocatalytic activity [11]. The doping with transition metals, such as Cr, Fe, Co, Mo, Mn, and V, or their ions [12,13,14,15,16,17,18,19], broadens the visible light absorption range by reducing the band gap or introducing additional intra-band gap energy states. Moreover, the immobilization of TiO2 powder on solid supports, such as ceramic [20], silicates [21,22,23,24], activated carbon [25], clay [26,27,28], sands [29,30], or zeolites [31,32,33,34], is a good strategy to enhance the dispersion of the TiO2 species and the light harvesting for an effective photodegradation. It is also a convenient method to solve the problem of the catalyst powder recovery after use, facilitating the recycling of the photocatalyst in long-term applications. Among them, zeolites have especially attracted an increasing interest, thanks to their high specific surface area, tunable hydrophobic/hydrophilic properties, tailored adsorption features, and environmentally friendly nature [35,36,37]. In addition, since zeolites possess relevant ionic exchange properties, they are, in principle, particularly suitable solids for the adsorption and subsequent catalytic degradation of positively-charged dye molecules [38].
Although many research teams have reported on the preparation of zeolite-supported or metal ion-doped TiO2 catalysts separately, few studies have investigated the effect of the two above-mentioned modifications simultaneously. Miaoliang et al. have studied the photocatalytic degradation of methyl orange by platinum-modified TiO2 on natural zeolite (Pt-TiO2/zeolites) [39]. It was found that the doping of Pt can improve significantly the degradation of methyl orange with 86.2% of the decolorization rate under 30 min irradiation time and the optimum Pt doping amount was 1.5 wt%. Lazau et al. [40] have compared the photodegradation of humic acid over undoped TiO2/natural zeolite (TiO2/zeolite) and the same TiO2/zeolite system doped with silver ions (Ag-TiO2/zeolite). The photocatalytic activity and the mineralization of humic acid under Ultraviolet-Visible (UV-VIS) irradiation was much more efficient in the presence of the Ag-TiO2/zeolite composite. The physical adsorption of humic acid by the composites could in addition accelerate its degradation rate. Khatamian et al. [41] evaluated the photocatalytic activity of supported Ni-TiO2 and Fe-TiO2 composites on the ZSM-5 (Zeolite Socony Mobil-5) zeolite during the transformation of Yellow GX pigment in aqueous solution under ambient conditions. The Fe-TiO2/ZSM-5 system showed to be a more efficient catalyst than Ni-TiO2/ZSM-5 and TiO2/ZSM-5 for the degradation of dye molecules. Its higher photocatalytic activity was attributed to the effective separation of charge carriers due to the fact that Fe3+ ions decrease the recombination process of electron–hole pairs. An Fe-TiO2 system deposited on a natural clay was also used as a catalyst for the photo-treatment of wastewaters [26]. However, the solid was properly active only under UV-A irradiation and in the presence of peroxide-containing oxidants.
In the present work, we have followed two strategies: (i) supporting evenly-dispersed TiO2 species onto a proton-containing microporous molecular sieve zeolite with variable oxide loadings, with the dual aim of avoiding the formation of TiO2 aggregates, as well as of optimizing the amount of TiO2 content for a maximal degradation of pollutants; and (ii) supporting Fe-doped TiO2 species onto a HY zeolite in order to have a photocatalyst with high catalytic activity in the oxidation and degradation of pollutants under visible light. Thanks to a cooperative positive effect due to the presence of a zeolite support with enhanced adsorption features, a photocatalytically-active, highly-dispersed TiO2 phase and a dopant (Fe species) with good properties in the catalytic degradation of pollutants, an improved performance of Fe + TiO2 on zeolite systems is expected with respect to the current state of the art for similar photocatalysts. The metal ion-doped photocatalysts were synthesized by a simple and conventional sol-gel method. The photocatalytic performance of all of the prepared catalysts in decolorization reactions was then evaluated in the degradation of methylene blue, MB, as a model polluting dye. Throughout the work, a constant attention was paid to optimizing the reaction under very mild conditions, with a simple experimental apparatus and under conventional visible light irradiation.

2. Results and Discussion

2.1. Adsorption Studies

Adsorption studies were initially carried out over a set of the most readily available zeolites, i.e., HY (H-faujasite), Hβ, HMOR (H-mordenite), and HZSM-5 zeolites in the presence of methylene blue, MB. The MB removal capability of the solids is shown in Figure 1. In all cases, after an initial rapid adsorption of MB within the first 10 min, the extent of MB removal attained a steady state in 30 min and no further significant decrease in MB concentration was detected even after 24 h, suggesting that adsorption equilibrium was reached. In order to be sure a steady situation is fully attained, 60 min was chosen as the optimum time for the adsorption equilibrium and all of the further photodegradation experiments were carried out after this time.
The lower adsorption property of Hβ (30%), HMOR (25%), and HZSM-5 (19%), compared to HY (38%), is attributed to the sterical constraints of the structure of zeolites. Actually, HY presents a network of interconnected cages with openings of 7.4 Å and a high specific surface area (880 m2 g−1). The faujasite structure of HY is, on the contrary, more accessible than the one of Hβ, which is composed of two straight pore channels (7.6 × 6.4 Å) together with a third zig-zag pore channel (5.6 × 5.6 Å), reducing the possibility of adsorption and diffusion of the adsorbate. Then, H-ZSM-5, which has a low specific surface area (411 m2 g−1) and the smallest pore size (5.1 × 5.5 Å), was the solid where the entry of MB was most difficult. Thus, the adsorption trend followed the order: HY > Hβ > HMOR > HZSM-5 (Figure 1).
In conclusion, HY zeolite showed the highest adsorption property for MB and, hence, it was chosen as the most suitable support for the rest of the MB photodegradation studies.

2.2. Effect of TiO2 Loading on HY Zeolite

2.2.1. Characterization of the Photocatalysts

Starting from HY zeolite as a support, a series of TiO2/HY catalysts was prepared with a TiO2 content spanning from 3 wt% to 15 wt%. In order to confirm the crystalline structure of the solids, the samples were studied by XRD method. Figure 2 shows the XRD patterns of the TiO2/HY catalysts, after calcination at 450 °C. The diffraction peaks at 2θ = 25°, 38°, 48°, 55°, 63°, 68°, 70°, and 75° were matched to the spectrum of the anatase TiO2 phase. The measurements showed that the thermal treatment at 450 °C had successfully transformed the amorphous TiO2 phase into a 100% crystalline TiO2 anatase phase. The gradual increase in the intensity of the anatase peaks (2θ = 25°, 38°, and 48°) of TiO2 and the corresponding relative decrease in the intensity of the peak (2θ = 16°) of HY indicates that the presence of anatase increased with the amount of TiO2 loaded onto the HY support [42]. On the other hand, no peaks assigned to the rutile or brookite phase were observed in the XRD patterns in any of the TiO2/HY catalysts. This result indicates that, for the TiO2/HY catalysts, the anatase was the kinetically metastable phase under the synthesis conditions tested here [43]. As there is no significant shift in the peaks for TiO2/HY in comparison with those of the pristine HY zeolite, it could be concluded that no change occurred in the zeolite crystal structure during the synthesis of TiO2/HY [44].
Figure 3 shows the FTR spectra of TiO2 and supported TiO2. For all of these samples, the region from 1600 cm−1 to 3700 cm−1 indicates the presence of water molecules and OH group in the zeolite [45]. The strongest absorption peak at 1047 cm−1 is assigned to the framework stretching vibration band of Si (Al)–O in the tetrahedral Si (Al) O4 units of the zeolite, indicating that the zeolite structure was kept. The structural bands at 450–900 cm−1 are attributed to the stretching vibrations of T–O, T–O–T, and O–T–O bonds in tetrahedral SiO4 and AlO4 units (T = Si or Al atoms). At the same time, no evident bands were observed in the region of 950–960 cm−1, commonly assigned to the antisymmetric stretching vibration of the Ti–O–Si bond [42,46]. As it is suggested that no strong chemical interaction takes place between the TiO2 species and the zeolite, TiO2 is likely dispersed on the surface of the zeolite or partially encapsulated within the zeolite cavities [47,48]. Furthermore, the peaks at 920, 895, and 860 cm−1 that are characteristic of Ti–O, Ti=O, and Ti–O–Ti, respectively, from bulk TiO2 are not evident in these samples. This is a further clue that the Ti species are small and homogeneously dispersed within the zeolite framework [49,50].
The values for the BET specific surface area of the HY zeolite and TiO2 containing zeolite photocatalysts are reported in Figure 4. The BET area of the HY samples gradually decreased when the loading of TiO2 increased. These diminutions in the surface area of the TiO2-containing zeolite solids, in comparison to the corresponding parent zeolite, are consistent with a gradual hindering and partial blocking of the pores of the zeolite by the TiO2 particles [51]. It is, however, worth highlighting that the BET surface area decreases linearly, with a small negative constant slope. This is a further clue that the TiO2 species are evenly and properly dispersed inside the support, as a function of the oxide loading.
The UV-VIS diffuse reflectance spectra of TiO2/HY and TiO2 catalysts are shown in Figure 5 and provide an insight about the interaction of the TiO2 species with the zeolite. The position of the absorption edge for pure anatase is around 400 nm. On the contrary, for the TiO2/HY systems, the edge is red-shifted and the shift increases as the TiO2 loading increases. This is likely due to the high degree of dispersion of the anatase aggregates within the zeolite [52,53,54].

2.2.2. Evaluation of the Photocatalytic Activity

The photocatalysis tests were performed over the series of 3, 7, 10, and 15 wt% TiO2-loaded HY solids in order to understand their influence on the degradation efficiency, and to optimize the amount of TiO2 loading for a maximal degradation of MB under UV irradiation (Figure 6). The results revealed that the use of the HY zeolite as a support in the TiO2/HY system played a key role in increasing and improving the degradation rate. First, the deposition of TiO2 onto the surface of HY zeolite enhances the electron transfer from the photo-excited semiconductor to the guest molecules of HY, as an electron acceptor and, consequently, the zeolite prevents the electron–hole recombination [55]. In addition, the presence of the zeolitic molecular sieve improves the photodegradation efficiency of TiO2 particles thanks to the synergistic presence of various positive factors: its high specific surface area, its good adsorption properties, and its capability to keep TiO2 moieties as dispersed as possible. For these reasons, the degradation efficiency increased by increasing the TiO2 loading on the support up to 10 wt%. However, above 10 wt%, the maximum degradation percentage was lower and such behavior can be attributed to a more extensive pore blocking of HY zeolite by titania species. The incipient pore blocking not only negatively affected the adsorption capacity, but also increased the shielding of the incident light by TiO2 particles, since an increased turbidity of the suspension reduces the solar light penetration [56]. Moreover, at higher TiO2 loading, the excited particles may not be in close proximity to the zeolite surface and, hence, the conduction band electrons of the TiO2 are less prone to be delocalized over the zeolite. As a result, there is a higher occurrence of the electron-hole recombination, giving rise to a lower degradation efficiency. Only an optimal tuning of the TiO2 loading on the zeolite surface could, therefore, lead to a more efficient degradation.
Taking into account these observations, in the present study, the optimum loading of TiO2 on HY was found to be 10 wt% and this value was chosen as a starting point for the further doping of the solids by iron species, in order to obtain Fe-doped TiO2/HY systems.

2.3. Effect of Fe Doping onto TiO2/HY Zeolite

2.3.1. Characterization of the Photocatalysts

The diffuse reflectance spectroscopy shows the changes in the optical properties of the supported titania after Fe doping by iron species, also revealed by the adjusted Kubelka-Munk transformation (see the Experimental Details section) aimed at determining the band-gap energy of the doped oxide (Figure 7). The wavelength of the spectral onset increased by increasing the dopant amount (Figure 7a). Therefore, the band diminution in band gap with the increase in Fe doping is to be attributed to the overlap of Fe 3d, Ti 3d, and O 2p states, which eventually induces the absorption of higher wavelength photons in titania [57]. Figure 7b shows the indirect transition Kubelka-Munk plot as a function of the photon energy allowing the estimation at the intercept of the threshold energies, namely, 3.12, 2.81, and 2.55 eV for TiO2, 3% Fe-TiO2/HY, and 10% Fe-TiO2/HY, respectively. The gradual decrease in the overall band gap energy with the amount of iron doping testifies to a shift towards the visible region.
Thanks to these modified absorption properties, it was possible to test the Fe-doped TiO2/HY catalysts under visible light irradiation. The UV fraction of the emission of the lamp used in the experimental apparatus was cut by means of a UV filter in order to allow only the visible fraction to reach the photoreactor.

2.3.2. Photocatalytic Studies

Figure 8a shows the degradation curves of MB solutions catalyzed by Fe-doped TiO2/HY photocatalysts at different Fe(III) dosages under visible light irradiation. It can be observed that the decolorization rate of MB solutions over these systems was higher than that one obtained in the presence of the undoped TiO2/HY systems. The degradation rate increased as the iron concentration grew from 3 wt% to 10 wt%. Then it decreased back at Fe(III) concentrations higher than 10 wt% (Figure 8b). This result is fully consistent with some previous observations [58,59]. An appropriate concentration of iron ions can introduce trapping species in the transfer between for photo-generated holes and electrons, thus inhibiting the recombination of electron-hole pairs [60]. Indeed, the recombination rate decreases as the dopant concentration grows, because the distance between trapping sites in a particle decreases with a higher number of dopants. However, if the concentration of trapping species is too high (>12 wt%), the overall process is slowed down and a lower MB removal is obtained (Figure 8). On the contrary, if the iron concentration is very low, there are fewer trapping sites available, and this leads to an overall diminution of the photocatalytic activity [61].
The removal capacity of Fe-TiO2/HY with 10 wt% of Fe content has been compared to the one of other similar catalytic systems reported in the literature and summarized in Table S1 (Supplementary Materials). The degradation performance is reported as percentage of dye removal (%). It should be noted that the comparison has a relative meaning only, because of different testing conditions (concentration and nature of dye, catalyst amount, concentration, and the nature of dopant, experimental setup, etc.) and the type of TiO2 support [25,38,39,41,62,63,64,65,66,67]. Nevertheless, if we look closely at the degradation of the dye pollutants, the present photocatalyst based on Fe-TiO2/HY attained very high decolorization percentages (>98%) at shorter times (60 min) and under milder conditions, avoiding the use of UV radiation or additional oxidizing reactants. The conceptually simple and reliable Fe-TiO2/HY system can, therefore, be considered as a promising catalyst for a potential application in the elimination of organic dye pollutants from wastewater.

3. Materials and Experimental Details

3.1. Materials

The commercial zeolites used in this study were obtained from Zeolyst International (Delfzijl, The Netherlands): HY (Si/Al = 30), HMOR (Si/Al = 20), NH4-β (Si/Al = 13.5), and NH4-ZSM-5 (Si/Al = 17). In order to obtain the protonic form (Hβ and HZSM-5), the ammonium form of the zeolite was calcined under dry air at 100 °C for 2 h and at 550 °C for 5 h using a heating rate of 1 °C min−1. Methylene blue (MB: basic blue 9, C16H18N3SCl; MW, 319.85 g mol−1) was used as the model pollutant and obtained from Merck Eurolab S.A. (Figure 9). MB has a maximum visible absorbance at a wavelength of 666 nm. Solutions were prepared by dissolving an appropriate amount of the dye in doubly-distilled water.

3.2. Photocatalyst Preparation

The zeolite-supported TiO2 and Fe-doped TiO2 catalysts (designated as TiO2/zeolite and Fe-TiO2/zeolite) were prepared according to a sol–gel method as described hereafter [68,69].
(1)
TiO2/zeolite: titanium(IV) tetraisopropoxide, TIPT (Aldrich, 97%, Darmstadt, Germany), was added to the zeolite-suspended isopropanol solution and the mixture was stirred for 2 h. Water was added to the resulting slurry to hydrolyze titanium(IV) tetraisopropoxide adsorbed on zeolite particles and then the resulting slurry was left under continuous magnetic stirring until a gel was obtained. The gel was dried at 90 °C for 12 h and calcinated at 450 °C for 3 h to give rise to TiO2 supported on zeolites. Finally, the catalysts were ground into a fine powder and stored in the dark. The amount of TiO2 loaded on zeolite was in the range from 0 to 15 wt%, namely, 3, 7, 10 and 15 wt% TiO2 on HY.
(2)
Fe-doped TiO2/zeolite: HY zeolite powders (specific surface area = 880 m2 g−1) were dispersed in 10-fold distilled water and heated up to 70 °C. Subsequently, TIPT solution and Fe (NO3)3·9H2O solution (the nominal weight ratios of Fe/Ti were 3.0, 6.0, 8.0, 10, 12% w/w, respectively) were added dropwise into the beaker and the pH of the suspension solution was adjusted to a value of about 3. The system was vigorously stirred at 70 °C for 4 h. After ageing for 12 h, the products were repeatedly washed with deionized water and then dried at 80 °C for 2 h. These samples were finally calcined at 450 °C for 3 h in a muffle furnace and then the photocatalysts were obtained. For the purpose of comparative photocatalysis study, a reference sample of titanium (IV) dioxide, TiO2 was also prepared using the same procedure except for the addition of zeolite.

3.3. Characterization of Photocatalyst

The crystalline structure of the photocatalyst was measured using X-ray diffraction (XRD, X Pert Pro Panalytical, Almelo, The Netherlands) with a Cu Kα source. The samples were analyzed in a 2θ range of 10–80° to identify the crystalline phase. IR spectra of the samples were recorded using a Shimadzu (Kyoto, Japan) FTIR spectrometer with KBr pellets. The samples were analyzed in the wave number range of 400–4000 cm−1. The BET specific surface area was measured by N2 adsorption at low temperature on a surface area analyzer (ASAP 2000, Micromeritics, Norcross, GA, USA). The diffuse reflectance spectra of dry powders were measured using a Varian CARY 4000 (Mulgrave, Australia) UV-VIS) equipped with a DRA900 integrated sphere. The values of band gap energy (Eg) were calculated using Equation (1):
(αhυ)½ = a (hυEg)
where α, υ, Eg, and a represent the absorption coefficient, frequency vibration, band gap, and proportional constant, respectively.

3.4. Adsorption Study

Prior to photocatalytic experiments, an adsorption study of MB onto the catalyst surface was carried out by mixing 100 mL of aqueous solution of MB with a fixed weight of the catalyst (100 mg). Aliquots were withdrawn at specific time intervals and the change in MB concentration was measured by using a UV-160A UV-VIS spectrophotometer (ThermoFisher, Shanghai, China) at 666 nm, the λmax of MB. The extent of equilibrium adsorption was determined from the decrease in MB concentration.

3.5. Photocatalysis Experiments

The experimental setup (Figure 10) was based on a rectangular photocatalytic reactor with a workable area of 9 cm × 45 cm, made out of Plexiglass® (Decoplast, Blida, Algeria). The wastewater stream to be treated entered the top of the chamber as a thin film onto the photocatalyst nanoparticles immobilized on the glass plate (8 cm × 44 cm). The sample to be treated (1000 mL of dye solution) was stored in a reservoir and continuously circulated in the system by a peristaltic pump at an adjustable flow rate. The reservoir was open to the air to ensure sufficient oxygenation. The solution in the reservoir was continuously stirred to keep the solution as homogeneous as possible. The whole reactor was thermostated at 25 °C.
Artificial irradiation was provided by a 15 W UV lamp with a maximum peak intensity at 254 nm, positioned above the reactor at a distance of 15 cm. The lamp was switched on at the beginning of each experiment. The immobilized photocatalyst was cleaned after every experiment by circulating deionized water with a few drops of aqueous hydrogen peroxide (30 wt%) under UV irradiation for regenerating the photocatalyst.
For visible light measurements, all the experimental parameters were the same except for the fact that the UV part of the light spectrum was cut by means of a 3 mm thick GG420 filter (Schott, Suzhou, China) placed on the photoreactor. The distance between the UV-VIS source and the solution was adjusted according to the experimental conditions. At different reaction times obtained with the experimental design, 2.5 mL samples were taken and the remaining BM was determined via spectrophotometer evaluation at λmax = 665 nm and a calibration curve. The dye removal efficiency (in percentage) was calculated by Equation (2):
degradation   efficiency   ( % ) = 100 % * ( C 0 C t ) C 0
where C0 is the initial concentration of the dye and Ct is its concentration at the time t.

4. Conclusions

The photocatalytic degradation of methylene blue, MB, was investigated using TiO2 species supported onto HY zeolite either in the presence or in the absence of Fe(III) doping species. XRD and UV-VIS results showed that the TiO2 species deposited on HY zeolite were in the anatase form and were mainly located on the outer surface of zeolite. The synergetic co-presence of the zeolite support, as an adsorbent, and TiO2 species, as catalytically-active sites, led to an improved light harvesting capability and to increased adsorption of the polluting substrate. This resulted in an enhanced degradation of the MB dye. Under UV irradiation, the best photocatalytic degradation performance was achieved for TiO2/HY at 10 wt% of TiO2 loading.
The further addition of Fe(III) species, as a dopant, led to an improved photocatalyst that could be simply used under visible light irradiation. Over this system, the Fe-doped TiO2/HY sample, with 10 wt% of Fe content, showed the best photocatalytic performance. A photo-assisted degradation of MB close to 100% was achieved after 60 min, under very mild conditions and visible light. Thanks to an appropriate tuning of the Fe(III) dopant concentration (10 wt%) within the TiO2 species, the unproductive recombination of photogenerated electron and hole pairs could be successfully inhibited, giving rise to an efficient visible-light responsive photocatalyst for abatement of polluting dyes.

Supplementary Materials

The following are available online at www.mdpi.com/2073-4344/7/11/344/s1, Table S1: Comparison of the efficiency in the removal of pollutant dyes from aqueous media catalyzed by different photocatalytic systems.

Acknowledgments

The authors would like to acknowledge the support provided by Abderrahmane Mira University of Bejaia, Algeria for funding this work.

Author Contributions

A.S. conceived and designed experiments; G.F. and N.C. performed the experiments; K.K. and D.R. analyzed the data; and A.S. and M.G. analyzed the data and wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schirmer, R.H.; Adler, H.; Pickhardt, M.; Mandelkow, E. “Lest we forget you—Methylene blue…”. Neurobiol. Aging 2011, 32, 2325.e7–2325.e16. [Google Scholar] [CrossRef] [PubMed]
  2. Malik, P.K. Use of activated carbons prepared from sawdust and rice-husk for adsorption of acid dyes: A case study of Acid Yellow 36. Dyes Pigments 2003, 56, 239–249. [Google Scholar] [CrossRef]
  3. Kovacic, P.; Somanathan, R. Toxicity of imine-iminium dyes and pigments: Electron transfer, radicals, oxidative stress and other physiological effects. J. Appl. Toxicol. 2014, 34, 825–834. [Google Scholar] [CrossRef] [PubMed]
  4. Vadivelan, V.; Kumar, K.V. Equilibrium, kinetics, mechanism, and process design for the sorption of methylene blue onto rice husk. J. Colloid Interface Sci. 2005, 286, 90–100. [Google Scholar] [CrossRef] [PubMed]
  5. Ravikumar, K.; Deebika, B.; Balu, K. Decolourization of aqueous dye solutions by a novel adsorbent: Application of statistical designs and surface plots for the optimization and regression analysis. J. Hazard. Mater. 2005, 122, 75–83. [Google Scholar] [CrossRef] [PubMed]
  6. Hamal, D.B.; Klabunde, K. Synthesis, characterization, and visible light activity of new nanoparticles photocatalysis based on silver, carbon, and sulfur-doped TiO2. J. Colloid Interface Sci. 2007, 311, 514–522. [Google Scholar] [CrossRef] [PubMed]
  7. Nawi, M.A.; Jawad, A.H.; Sabar, S.; Wan Ngah, W.S. Immobilized bilayer TiO2/chitosan system for the removal of phenol under irradiation by a 45 Watt compact fluorescent lamp. Desalination 2011, 280, 288–296. [Google Scholar] [CrossRef]
  8. Caratto, V.; Locardi, F.; Alberti, S.; Villa, S.; Sanguineti, E.; Martinelli, A.; Balbi, T.; Canesi, L.; Ferretti, M. Different sol-gel prepartion of iron-doped TiO2 nanoparticles: Characterization, photocatalytic and cytotoxicity. J. Sol.-Gel Sci. Technol. 2016, 80, 152–159. [Google Scholar] [CrossRef]
  9. Zhang, Q.; Fan, W.; Gao, L. Anatase TiO2 nanoparticles immobilized on ZnO tetrapods as a highly efficient and easily recyclable photocatalyst. Appl. Catal. B 2007, 76, 168–173. [Google Scholar] [CrossRef]
  10. Andronic, L.; Duta, A. TiO2 thin films for dyes photodegradation. Thin Solid Films 2007, 515, 6294–6297. [Google Scholar] [CrossRef]
  11. Myilsamy, M.; Mahalakshmi, M.; Subha, N.; Rajabhuvaneswari, A.; Murugesan, V. Visible light responsive mesoporous graphene-Eu2O3/TiO2 nanocomposites for the efficient photocatalytic degradation of 4-chlorophenol. RSC Adv. 2016, 6, 35024–35035. [Google Scholar] [CrossRef]
  12. Izadyar, S.; Fatemi, S. Fabrication of X Zeolite Based Modified Nano TiO2 Photocatalytic Paper for Removal of VOC Pollutants under Visible Light. Ind. Eng. Chem. Res. 2013, 52, 10961–10968. [Google Scholar] [CrossRef]
  13. Zhu, J.F.; Deng, Z.G.; Chen, F.; Zhang, J.L.; Chen, H.J.; Anpo, M.; Huang, J.Z.; Zhang, L.Z. Structures and magnetic properties of p-type Mn: TiO2 dilute magnetic semiconductor thin films. Appl. Catal. B 2006, 62, 329–335. [Google Scholar] [CrossRef]
  14. Anpo, M.; Takeuchi, M. The design and development of highly reactive titanium oxide photocatalysts operating under visible light irradiation. J. Catal. 2003, 216, 505–516. [Google Scholar] [CrossRef]
  15. Maeda, K.; Shimodaira, Y.; Lee, B.; Teramura, K.; Lu, D.; Kobayashi, H.; Domen, K. Studies on TiNxOyFz as a Visible-Light-Responsive Photocatalyst. J. Phys. Chem. C 2007, 111, 18264–18270. [Google Scholar] [CrossRef]
  16. Carneiro, J.O.; Teixeira, V.O.; Portinha, A.; Magalhaes, A.; Coutinho, P.; Tavares, C.J.; Newton, R. Iron-doped photocatalytic TiO2 sputtered coatings on plastics for self-cleaning applications. Mater. Sci. Eng. B 2007, 138, 144–150. [Google Scholar] [CrossRef]
  17. Gota, K.R.; Suresh, S. Preparation and its application of TiO2/ZrO2 and TiO2/Fe photocatalysts: A perspective study. Asian J. Chem. 2014, 26, 7087–7101. [Google Scholar]
  18. Zhang, Y.; Fan, W.; Du, H.Q.; Zhao, Y.W.; Song, R.G.; Xiang, N. Microstructure and photocatalytic property of TiO2 and Fe3+:TiO2 films produced by micro-arc oxidation. Surf. Coat. Technol. 2017, 315, 196–204. [Google Scholar] [CrossRef]
  19. Sunada, F.; Heller, A. Effects of water, salt water, and silicone over coating of the TiO2 photocatalyst on the rates and products of photocatalytic oxidation of liquid 3-octanol and 3-octanone. Environ. Sci. Technol. 1998, 32, 282–288. [Google Scholar] [CrossRef]
  20. Shi, J.; Kuwahara, Y.; An, T.; Yamashita, H. The fabrication of TiO2 supported on slag-made calcium silicate as low-cost photocatalyst with high adsorption ability for the degradation of dye pollutants in water. Catal. Today 2017, 281, 21–28. [Google Scholar] [CrossRef]
  21. Strauss, M.; Pastorello, M.; Sigoli, F.A.; de Souza e Silva, J.M.; Mazali, I.O. Singular effect of crystallite size on the charge carrier generation and photocatalytic activity of nano-TiO2. Appl. Surf. Sci. 2014, 319, 151–157. [Google Scholar] [CrossRef]
  22. Chen, Y.; Wang, K.; Lou, L. Photodegradation of dye pollutants on silica gel supported TiO2 particles under visible light irradiation. J. Photochem. Photobiol. A Chem. 2004, 163, 281–287. [Google Scholar] [CrossRef]
  23. Xu, Y.; Zheng, W.; Liu, W. Enhanced photocatalytic activity of supported TiO2: Dispersing effect of SiO2. J. Photochem. Photobiol. A Chem. 1999, 122, 57–60. [Google Scholar] [CrossRef]
  24. Hermann, J.; Matos, J.; Didier, J.; Guillard, C.; Laine, J.; Malato, S.; Blanco, J. Solar photocatalytic degradation of 4-chlorophenol using the synergistic effect between titania and activated carbon in aqueous suspension. Catal. Today 1999, 54, 255–265. [Google Scholar] [CrossRef]
  25. Bel Hadjltaief, H.; Ben Zina, M.; Da Costa, P.; Galvez, M.E. Heterogeneous TiO2–Fe-plate catalyst for the discoloration and mineralization of aqueous solutions of cationic and anionic dyes. Desalination Water Treat. 2016, 57, 13505–13517. [Google Scholar] [CrossRef]
  26. Ooka, C.; Yoshida, H.; Suzuki, K.; Hattori, T. Highly hydrophobic TiO2 pillared clay for photocatalytic degradation of organic compounds in water. Microporous Mesoporous Mater. 2004, 67, 143–150. [Google Scholar] [CrossRef]
  27. Sun, Z.; Chen, Y.; Ke, Q.; Yang, Y.; Yuan, J. Photocatalytic degradation of cationic azo dye by TiO2/bentonite nanocomposite. J. Photochem. Photobiol. A Chem. 2002, 149, 169–174. [Google Scholar] [CrossRef]
  28. Kang, M.G.; Park, H.S.; Kim, K.J. Effect of improved crystallinity of titanium silicalite-2 on photodecomposition of simple aromatic hydrocarbons. J. Photochem. Photobiol. A Chem. 2002, 149, 175–181. [Google Scholar] [CrossRef]
  29. Hanaor, D.A.H.; Sorrell, C.C. Sand supported mixed-phase TiO2 photocatalysts for water decontamination applications. Adv. Eng. Mater. 2014, 16, 248–254. [Google Scholar] [CrossRef]
  30. Luo, M.; Bowden, D.; Brimblecombe, P. Removal of Dyes from Water Using a TiO2 Photocatalyst Supported on Black Sand. Water Air Soil Pollut. 2009, 198, 233–241. [Google Scholar] [CrossRef]
  31. Bhattacharyya, A.; Kawi, S.; Ray, M.B. Photocatalytic degradation of orange II by TiO2 catalysts supported on adsorbents. Catal. Today 2004, 98, 431–439. [Google Scholar] [CrossRef]
  32. Suwarnkar, M.B.; Kadam, A.N.; Khade, G.V.; Gavade, N.L.; Garadkar, K.M. Modification of TiO2 nanoparticles by HZSM-5 for the enhancement in photodegradation of Acid Green 25. J. Mater. Sci. Mater. Electron. 2016, 27, 843–851. [Google Scholar] [CrossRef]
  33. Ilinoiu, E.C.; Pode, R.; Manea, F.; Colar, L.A.; Jakab, A.; Orha, C.; Ratiu, C.; Lazau, C.; Sfarloaga, P. Photocatalytic activity of a nitrogen-doped TiO2 modified zeolite in the degradation of Reactive Yellow 125 azo dye. J. Taiwan Inst. Chem. Eng. 2013, 44, 270–278. [Google Scholar] [CrossRef]
  34. Wang, S.; Peng, Y. Natural zeolites as effective adsorbents in water and wastewater treatment. Chem. Eng. J. 2010, 156, 11–24. [Google Scholar] [CrossRef]
  35. Mahalakshmi, M.; Vishnu Priya, S.; Arabindoo, B.; Palanichamy, M.; Murugesan, M. Photocatalytic degradation of aqueous propoxur solution using TiO2 and Hβ zeolite-supported TiO2. J. Hazard. Mater. 2009, 161, 336–343. [Google Scholar] [CrossRef] [PubMed]
  36. Wang, S.; Li, H.; Xie, S.; Liu, S.; Xu, L. Physical and chemical regeneration of zeolitic adsorbents for dye removal in wastewater treatment. Chemosphere 2006, 65, 82–87. [Google Scholar] [CrossRef] [PubMed]
  37. Corma, A. From microporous to mesoporous molecular sieve materials and their use in catalysis. Chem. Rev. 1997, 97, 2373–2419. [Google Scholar] [CrossRef] [PubMed]
  38. Nezamzadeh-Ejhieh, A.; Karimi-Shamsabadi, M. Comparison of photocatalytic efficiency of supported CuO onto microand nano particles of zeolite X in photodecolorization of methylene blue and methyl orange aqueous mixture. Appl. Catal. A Gen. 2014, 477, 83–92. [Google Scholar] [CrossRef]
  39. Miaoliang, H.; Xu, C.; Wu, Z.; Huang, Y.; Lin, J.; Wu, J. Photocatalytic discolorization of methyl orange solution by Pt modified TiO2 loaded on natural zeolite. Dyes Pigments 2008, 77, 327–334. [Google Scholar]
  40. Lazau, C.; Ratiu, C.; Orha, C.; Pode, R.; Manea, F. Photocatalytic activity of undoped and Ag-doped TiO2-supported zeolite for humic acid degradation and mineralization. Mater. Res. Bull. 2011, 46, 1916–1921. [Google Scholar] [CrossRef]
  41. Khatamian, M.; Hashemian, S.; Yavari, A.; Saket, M. Preparation of metal ion (Fe3+ and Ni2+) doped TiO2 nanoparticles supported on ZSM-5 zeolite and investigation of its photocatalytic activity. Mater. Sci. Eng. B 2012, 177, 1623–1627. [Google Scholar] [CrossRef]
  42. Kim, Y.; Yoon, M. TiO2/Y-Zeolite encapsulating intramolecular charge transfer molecules: A new photocatalyst for photoreduction of methyl orange in aqueous medium. J. Mol. Catal. A Chem. 2001, 168, 257–263. [Google Scholar] [CrossRef]
  43. Hanaor, D.A.H.; Sorrell, C.C. Review of the anatase to rutile phase transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef]
  44. Shankar, M.V.; Cheralathan, K.K.; Arabindoo, B.; Palanichamy, M.; Murugesan, V. Enhanced photocatalytic activity for the destruction of monocrotophos pesticide by TiO2/Hβ. J. Mol. Catal. A Chem. 2004, 223, 195–200. [Google Scholar] [CrossRef]
  45. Hassan, H.; Hameed, B.H. Oxidative decolorization of Acid Red 1 solutions by Fe–zeolite Y type catalyst. Desalination 2011, 276, 45–52. [Google Scholar] [CrossRef]
  46. Petkowicz, D.I.; Brambilla, R.; Radtke, C.; Da Silva, C.; Da Rocha, Z.N.; Pergher, S.B.C.; Dos Santos, J.H.Z. Photodegradation of methylene blue by in situ generated titania supported on a NaA zeolite. Appl. Catal. A Gen. 2009, 357, 125–134. [Google Scholar] [CrossRef]
  47. Cheng-Cai, W.; Chung-Kung, L.; Meng-Du, L.; Lain-Chuen, J. Photocatalytic degradation of C.I. Basic Violet 10 using TiO2 catalysts supported by Y zeolite: An investigation of the effects of operational parameters. Dyes Pigments 2008, 76, 817–824. [Google Scholar]
  48. Aramendía, M.A.; Colmenares, J.C.; López-Fernández, S.; Marinas, A.; Marinas, J.M.; Urbano, F.J. Screening of different zeolite-based catalysts for gas-phase selective photooxidation of propan-2-ol. Catal. Today 2007, 129, 102–109. [Google Scholar] [CrossRef]
  49. Liu, X.; Iu, K.K.; Thomas, J.K. Encapsulation of TiO2 in zeolite Y. Chem. Phys. Lett. 1992, 195, 163–168. [Google Scholar] [CrossRef]
  50. Anpo, M.; Yamashita, H.; Ichihashi, Y.; Fujii, Y.; Honda, M. Photocatalytic Reduction of CO2 with H2O on Titanium Oxides Anchored within Micropores of Zeolites: Effects of the Structure of the Active Sites and the Addition of Pt. J. Phys. Chem. B 1997, 101, 2632–2636. [Google Scholar] [CrossRef]
  51. Cheralathan, K.K.; Sudarsan Kumar, I.; Palanichamy, M.; Murugesan, V.J. Liquid phase alkylation of phenol with 4-hydroxybutan-2-one in the presence of modified zeolite HBEA. Appl. Catal. A Gen. 2003, 241, 247–260. [Google Scholar] [CrossRef]
  52. Durga Kumari, V.; Subrahmanyam, M.; Subba Rao, K.V.; Ratnamala, A.; Noorjahan, M.; Tanaka, K. An easy and efficient use of TiO2 supported HZM-5 and TiO2+HZSM-5 zeolite combinate in the photodegradation of aqueous phenol and p-chlorophenol. Appl. Catal. A Gen. 2002, 234, 155–165. [Google Scholar] [CrossRef]
  53. Zheng, S.; Gao, L.; Zhang, Q.; Zhang, W.; Guo, J. Preparation, characterization and photocatalytic properties of singly and doubly titania-modified mesoporous silicate MCM-41 by varying titanium precursors. J. Mater. Chem. 2001, 11, 578–583. [Google Scholar] [CrossRef]
  54. Babic, B.; Zarubic, A.; Minovic Arsic, T.; Pantic, J.; Jokic, B.; Abazovic, N.; Matovic, B. Iron doped anatase for application in photocatalysis. Eur. Ceram. Soc. 2016, 36, 2991–2996. [Google Scholar] [CrossRef]
  55. Sasikala, R.; Shirole, A.R.; Sudarsan, V.; Jagannath; Sudakar, C.; Naik, R.; Rao, R.; Bharadwaj, S.R. Enhanced photocatalytic activity of indium and nitrogen co-doped TiO2–Pd nanocomposites for hydrogen generation. Appl. Catal. A Gen. 2010, 377, 47–54. [Google Scholar] [CrossRef]
  56. Phanikrishna Sharma, M.V.; Lalitha, K.; Durgakumari, V.; Subrahmanyam, M. Solar photocatalytic mineralization of isoproturon over TiO2/HY composite systems. Sol. Energy Mater. Sol. Cells 2008, 92, 332–342. [Google Scholar] [CrossRef]
  57. Qamara, M.; Merzougui, B.; Anjum, D.; Hakeem, A.S.; Yamani, Z.H.; Bahnemann, D. Synthesis and photocatalytic activity of mesoporous nanocrystalline Fe-doped titanium Dioxide. Catal. Today 2014, 230, 158–165. [Google Scholar] [CrossRef]
  58. Lei, J.; Li, X.; Li, W.; Sun, F.; Lu, D.; Lin, Y. Photocatalytic degradation of methyl orange on arrayed porous iron-doped anatase TiO2. J. Solid State Electrochem. 2012, 16, 625–632. [Google Scholar] [CrossRef]
  59. Navio, J.A.; Testa, J.J.; Djedjeian, P.; Padron, J.R.; Rodriguez, D.; Litter, M.I. Iron doped titania powders prepared by a sol–gel method. Part II: Photocatalytic properties. Appl. Catal. A 1999, 178, 191–203. [Google Scholar] [CrossRef]
  60. Tong, T.Z.; Zhang, J.L.; Tian, B.Z.; Chen, F.; He, D.N. Preparation of Fe3+-doped TiO2 catalysts by controlled hydrolysis of titanium alkoxide and study on their photocatalytic activity for methyl orange degradation. J. Hazard. Mater. 2008, 155, 572–579. [Google Scholar] [CrossRef] [PubMed]
  61. Litter, M.A.; Navio, J.A. Photocatalytic properties of iron-doped titania semiconductors. J. Photochem. Photobiol. A 1996, 98, 171–181. [Google Scholar] [CrossRef]
  62. Priyanka, K.P.; Sankararaman, S.; Balakrishna, K.M.; Varghese, T. Enhanced visible light photocatalysis using TiO2/phthalocyanine nanocomposites for the degradation of selected industrial dyes. J. Alloys Compd. 2017, 720, 541–549. [Google Scholar] [CrossRef]
  63. Miranda, L.D.L.; Bellato, C.R.; Milagres, J.A.L.; Moura, L.G.; Mounteer, A.H.; de Almeida, M.F. Hydrotalcite TiO2 magnetic iron oxide intercalated with the anionic surfactant dodecylsulfate in the photocatalytic degradation of methylene blue dye. J. Environ. Manag. 2015, 156, 225–235. [Google Scholar] [CrossRef] [PubMed]
  64. Guo, Z.; Chen, B.; Mu, J.; Zhang, M.; Zhang, P.; Zhang, Z.; Wang, J.; Zhang, X.; Sun, Y.; Shao, C.; et al. Iron phthalocyanine/TiO2 nanofiber heterostructures with enhanced visible photocatalytic activity assisted with H2O2. J. Hazard. Mater. 2012, 219–220, 156–163. [Google Scholar] [CrossRef] [PubMed]
  65. Mesgaria, Z.; Gharagozloub, M.; Khosravic, A.; Gharanjig, K. Synthesis, characterization and evaluation of efficiency of new hybrid Pc/Fe-TiO2 nanocomposite as photocatalyst for decolorization of methyl orange using visible light irradiation. Appl. Catal. A Gen. 2012, 411–412, 139–145. [Google Scholar] [CrossRef]
  66. Shende, T.P.; Bhanvase, B.A.; Rathod, A.P.; Pinjari, D.V.; Sonawane, S.H. Sonochemical synthesis of Graphene-Ce-TiO2 and Graphene-Fe-TiO2 ternary hybrid photocatalyst nanocomposite and its application in degradation of Crystal Violet Dye. Ultrason. Sonochem. 2018, 41, 582–589. [Google Scholar] [CrossRef]
  67. Rosu, M.C.; Coros, M.; Pogacean, F.; Magerusan, L.; Socaci, C.; Turza, A.; Pruneanu, S. Azo dyes degradation using TiO2-Pt/graphene oxide and TiO2-Pt/reduced graphene oxide photocatalysts under UV and natural sunlight irradiation. Solid State Sci. 2017, 70, 13–20. [Google Scholar] [CrossRef]
  68. Zhang, W.; Wang, K.; Yu, Y.; He, H. TiO2/HZSM-5 nano-composite photocatalyst: HCl treatment of NaZSM-5 promotes photocatalytic degradation of methyl orange. Chem. Eng. J. 2010, 163, 62–67. [Google Scholar] [CrossRef]
  69. Wang, C.; Shi, H.; Li, Y. Synthesis and characteristics of natural zeolite supported Fe3+-TiO2 photocatalysts. Appl. Surf. Sci. 2011, 257, 6873–6877. [Google Scholar] [CrossRef]
Figure 1. Adsorption of MB by different zeolites: HY (∎), Hβ (∆), HMOR (●), and HZSM-5 (◯). Experimental conditions: dye concentration = 20 mg L−1, zeolite amount = 1 g L−1, T = 25 °C.
Figure 1. Adsorption of MB by different zeolites: HY (∎), Hβ (∆), HMOR (●), and HZSM-5 (◯). Experimental conditions: dye concentration = 20 mg L−1, zeolite amount = 1 g L−1, T = 25 °C.
Catalysts 07 00344 g001
Figure 2. XRD patterns of TiO2, HY and 3, 7, 10, and 15 wt% TiO2/HY, from bottom to top, respectively.
Figure 2. XRD patterns of TiO2, HY and 3, 7, 10, and 15 wt% TiO2/HY, from bottom to top, respectively.
Catalysts 07 00344 g002
Figure 3. FTIR spectra of HY, 3, 7, 10, and 15 wt% TiO2/HY from bottom to top, respectively.
Figure 3. FTIR spectra of HY, 3, 7, 10, and 15 wt% TiO2/HY from bottom to top, respectively.
Catalysts 07 00344 g003
Figure 4. BET specific surface area vs. TiO2 loading (wt%) on TiO2/HY zeolite solids.
Figure 4. BET specific surface area vs. TiO2 loading (wt%) on TiO2/HY zeolite solids.
Catalysts 07 00344 g004
Figure 5. UV-VIS spectra of HY, TiO2, and the series of 3, 7, 10, and 15 wt% TiO2/HY systems.
Figure 5. UV-VIS spectra of HY, TiO2, and the series of 3, 7, 10, and 15 wt% TiO2/HY systems.
Catalysts 07 00344 g005
Figure 6. Photocatalytic degradation of MB over TiO2 supported HY zeolite at different TiO2 loadings under UV irradiation (λmax = 254 nm, dye concentration = 20 mg L−1, zeolite amount = 1 g L−1, T = 25 °C).
Figure 6. Photocatalytic degradation of MB over TiO2 supported HY zeolite at different TiO2 loadings under UV irradiation (λmax = 254 nm, dye concentration = 20 mg L−1, zeolite amount = 1 g L−1, T = 25 °C).
Catalysts 07 00344 g006
Figure 7. (a) UV-VIS diffuse reflectance spectra before and after doping with Fe; and (b) the corresponding Kubelka-Munk-transformed diffuse reflectance, as a function of hυ.
Figure 7. (a) UV-VIS diffuse reflectance spectra before and after doping with Fe; and (b) the corresponding Kubelka-Munk-transformed diffuse reflectance, as a function of hυ.
Catalysts 07 00344 g007
Figure 8. Photocatalytic degradation of MB solution catalyzed by undoped TiO2/HY and doped TiO2/HY with different amounts of iron dopant. Tests were run under visible light irradiation. (a) Photocatalytic degradation rate vs. irradiation time; and (b) the influence of the ion dopant on dye removal (time = 60 min). Experimental conditions: dye concentration = 20 mg L−1, catalyst amount = 1 g L−1, T = 25 °C.
Figure 8. Photocatalytic degradation of MB solution catalyzed by undoped TiO2/HY and doped TiO2/HY with different amounts of iron dopant. Tests were run under visible light irradiation. (a) Photocatalytic degradation rate vs. irradiation time; and (b) the influence of the ion dopant on dye removal (time = 60 min). Experimental conditions: dye concentration = 20 mg L−1, catalyst amount = 1 g L−1, T = 25 °C.
Catalysts 07 00344 g008
Figure 9. Methylene blue.
Figure 9. Methylene blue.
Catalysts 07 00344 g009
Figure 10. The experimental setup of the photocatalytic reactor.
Figure 10. The experimental setup of the photocatalytic reactor.
Catalysts 07 00344 g010

Share and Cite

MDPI and ACS Style

Foura, G.; Chouchou, N.; Soualah, A.; Kouachi, K.; Guidotti, M.; Robert, D. Fe-Doped TiO2 Supported on HY Zeolite for Solar Photocatalytic Treatment of Dye Pollutants. Catalysts 2017, 7, 344. https://doi.org/10.3390/catal7110344

AMA Style

Foura G, Chouchou N, Soualah A, Kouachi K, Guidotti M, Robert D. Fe-Doped TiO2 Supported on HY Zeolite for Solar Photocatalytic Treatment of Dye Pollutants. Catalysts. 2017; 7(11):344. https://doi.org/10.3390/catal7110344

Chicago/Turabian Style

Foura, Ghania, Nawel Chouchou, Ahcène Soualah, Kahina Kouachi, Matteo Guidotti, and Didier Robert. 2017. "Fe-Doped TiO2 Supported on HY Zeolite for Solar Photocatalytic Treatment of Dye Pollutants" Catalysts 7, no. 11: 344. https://doi.org/10.3390/catal7110344

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop