Next Article in Journal
Evaluation of Different Methods to Assess the Hydraulic Behavior in Horizontal Treatment Wetlands
Previous Article in Journal
Does Data Availability Constrain Temperature-Index Snow Models? A Case Study in a Humid Boreal Forest
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Aluminum-Doped Cobalt Ferrite as an Efficient Photocatalyst for the Abatement of Methylene Blue

1
Institute of Chemical Sciences, Bahauddin Zakariya University, Multan 60800, Pakistan
2
Department of Physics, Bahauddin Zakariya University, Multan 60800, Pakistan
3
School of Energy and Power Engineering, Xi’an Jiaotong University, 28 West Xianning Road, Xi’an 710049, Shaanxi, China
4
School of Chemical and Process Engineering, University of Leeds, Leeds LS2 9JT, UK
5
C2P2 (CNRS-UMR 5265), Université Claude Bernard Lyon 1, ESCPE Lyon, 43 Boulevard du 11 Novembre 1918, CEDEX, 69626 Villeurbanne, France
6
Department of Applied Chemistry and Chemical Technology, University of Karachi, Karachi 75270, Pakistan
*
Authors to whom correspondence should be addressed.
Water 2020, 12(8), 2285; https://doi.org/10.3390/w12082285
Submission received: 24 July 2020 / Revised: 10 August 2020 / Accepted: 11 August 2020 / Published: 14 August 2020
(This article belongs to the Section Wastewater Treatment and Reuse)

Abstract

:
The present study is aimed to access the photodegradation efficiency of methylene blue dye using CoFe2O4 and Co0.1Al0.03Fe0.17O0.4 nanoparticles. The synthesis of spinel ferrites nanoparticles was performed by a facile sol-gel method. The synthesized nanoparticles were characterized by FTIR, XRD, SEM, EDS, Nitrogen adsorption/desorption and UV–Visible spectroscopy. The XRD studies confirmed the spinel cubic structure of ferrite. It was also found that the crystallinity increases at an annealing temperature of 800 °C. The application of these nanoparticles for methylene blue’s photocatalytic degradation was explored and also the optimization of several parameters involving dye’s concentration, amount of catalyst and pH of the solution was done. Photocatalytic degradation of methylene blue showed that at pH 11, using 200 W visible light bulb and in 120 min; 93% methylene blue dye was degraded by using 0.1 g of Co0.1Al0.03Fe0.17O0.4.

1. Introduction

Pollution is unquestionably among the most serious and prevailing dangers of the modern world that have been threatening the biosphere for decades and to overcome this menace, immediate preventive measures are necessary [1,2]. Water pollution caused by dyes is a global concern and their removal through cost-effective and eco-friendly techniques is a challenging task [3,4]. Several industries that employ dyeing procedures, e.g., textile, leather and paper, contribute significantly to water pollution due to the excessive use of toxic compounds like dyes, which eventually enter the ecosystem, affecting the surrounding flora and fauna either directly or indirectly [5]. They can also affect our drinking water supplies since they are not easily degraded through natural or biological methods. Additionally, their toxicity poses a danger to the environment, they can form a layer above the sea-water (or freshwater) surface and block the sunlight, which is one of the most relevant components for the survival of aquatic life. This is due to the inefficiency of the dyeing process in which the dyes are still present in their waste due to their high stability in water [6].
Methylene blue (MB), C16H18N3SCl, an organic dye used in the industry, is well-known to cause water pollution [7]. It is linked to numerous health issues in humans—such as irritation of digestive tract resulting in anaemia, nausea, vomiting, and diarrhoea—and enhance eutrophication [8] which results in a misbalance of the aquatic ecosystem that may cause catastrophic effects to fauna and flora [9]. Previously, chemical coagulation, biological, and adsorption methods were used to treat industrial waste but these approaches do not completely degrade MB in water, and nowadays the advanced oxidative processes (AOPs) have become the most commonly used methods for wastewater treatment [10]. In this process, the generation of active radical species (OH) takes place from common oxidants such as O3, H2O2, and O2. For the improvement in the dye removal efficiency, AOPs are modified with the doping of transition metal salts and by irradiation with different light sources (e.g., ultraviolet, visible light, and γ rays) as well as ultrasonic waves with the aim of increasing the generation rate of OH.
In this study, the efficient photocatalytic degradation method is used to degrade MB dye by using aluminum-doped spinel ferrite nanoparticles as photocatalysts [11] because they are stable and present high catalytic activity due to their small bandgap [12]. The wide use of ferrite as nanocatalyst [13] is due to their cost-effectiveness, stability, facile preparation, stability towards harsh conditions, and its docile nature towards further modifications [14,15]. The ferrite nanocatalysts are generally nontoxic or slightly phytotoxic only if freshly prepared NPs are released into water [16,17]. Then, both doped and undoped ferrites have gained a great interest in catalysis [18], and for this study, undoped CoFe2O4 and doped Co0.1Al0.03Fe0.17O0.4 were synthesized, characterized and used for the photocatalytic degradation of MB dye in the presence of visible light. Several parameters affecting the catalytic activity were also optimized.

2. Experimental

2.1. Synthesis of Undoped CoFe2O4 and Doped CoAlxFe2−xO4

Aluminum-doped cobalt nanoferrites, CoAlxFe2−xO4, in which x is 0.1, 0.2, 0.3, 0.4 and 0.5 have been synthesized by the sol-gel combustion method [19]. All the reagents used for the synthesis were of analytical grade and were used as received. The 100 mL of each aqueous solution of the nitrates [0.1 M Co (NO3)2·6H2O, 0.5 M Fe (NO3)2·9H2O, and 0.2 M Al (NO3)2·9H2O] was prepared and mixed followed by the slow addition of 9 g citric acid (chelating agent). To precipitate out nanoferrites, the pH was adjusted to 8 by using a 32% ammonia solution. Afterward, the water was evaporated and the solution was converted into a wet gel by constant heating at 70 °C. After some time, the gel underwent self-ignition, transforming the sample to ash, which was ground and annealed at 800 °C for 5 h, using a heating rate of 20 °C min−1.
All materials were characterized using standard techniques and were used for MB dye degradation.

2.2. Characterization

2.2.1. X-ray Diffraction (XRD)s

XRD served to determine the phases and crystallinity of the synthesized nanoferrite samples over a 2θ range of 10° to 80°. XRD was performed on a Raigaku D/Max 3A Model 1981 using CuKα radiation at room temperature. Debye–Scherrer formula was applied for the calculation of average crystallite size given in Equation (1) [20]:
D = (0.9λ)/(βcosθ)
where “D” presents average crystallite size, “θ” is Bragg’s angle, “β” serves for full width half maximum (FWHM) and λ is the wavelength of the X-ray (CuKα is 1.54051 A). The lattice parameter of the spinel ferrites was determined by Equation (2):
a = d(h2 + k2 + l2)1/2
where “a” represents lattice constant, “d” is interplanar spacing, and h, k and l are Miller indices. Unit cell volume is found out by Equation (3):
V = a3
where “a” is the lattice parameter and V is the cell’s volume.

2.2.2. SEM-EDS Analysis

SEM-EDS analysis was performed using NOVA NANO Scanning Electron Microscope equipped with an Energy-Dispersive Spectrophotometer (EDS). Particle size and their distribution were calculated using the Image J and Origin Pro software, while the elemental analysis of CoFe2O4 and Co0.1Al0.3Fe1.7O0.4 with different compositions was determined through EDS.

2.2.3. Nitrogen Adsorption/Desorption Studies

The surface area was calculated through the N2 sorption isotherms on a Nova Station A. Around 0.3 g of each sample were used and analysis was performed at liquid nitrogen bath temperature.

2.3. Degradation of MB Dye Using Synthesized Spinel Ferrites

For the photocatalytic degradation assays, a dark photocatalytic reactor was used to avoid outside illumination and, when required, a reflective surface was used inside. Visible light bulbs containing tungsten filament (Philips 100 W and 200 W) were applied as a source of photons; it was connected at the top of the reactor chamber. The reaction chamber was a simple wooden box whose inside walls were covered with aluminum foil (Figure 1). The photocatalytic degradation of MB dye with undoped (CoFe2O4) and doped (Co0.1Al0.03Fe0.17O0.4) cobalt ferrite nanoparticles was estimated by quantification of the MB dye using a UV–Vis spectrophotometer. Photocatalysts were taken into a beaker containing the MB solution and put inside the photoreactor. After attaining the adsorption equilibrium (Ao) (about 30 min), the light was turned on to study the photocatalytic conversion of the MB dye solution. After a time-lapse of 30 min, an aliquot was taken using a centrifuge machine. The MB dye’s concentration after irradiation was determined by measuring the absorbance intensity of each aliquot at the dye maximum absorption λ, i.e., 665 nm, by using an ultraviolet-visible spectrophotometer (SHIMADZU 1800). The dye’s % degradation was then found by using Equation (4):
% Degradation = (Ao − A/Ao) × 100
where
  • Ao is MB dye’s initial absorbance
  • A is MB dye’s final absorbance, after irradiation.

2.4. Optimization of Different Variables Affecting Degradation of MB Dye

2.4.1. Effect of Time

Illumination time is considered a prime factor affecting the dye’s degradation. The role of illumination time on dye’s percentage (%) degradation was studied at different time periods till equilibrium was established while keeping other variables constant, i.e., catalyst amount 0.5 g and 10 mg/L MB dye solution.

2.4.2. Effect of pH

The effect of pH on methylene blue dye photocatalytic degradation was considered in the pH range of 3–11 using 100 mL of 10 mg/L of the initial concentration of MB dye and a dose of 0.5 g under visible light for 120 min. MB solution’s pH was adjusted either through 0.1 M HCl or 0.1 M NaOH. Moreover, the pH of the MB dye solution was found out before and after the photocatalytic degradation process.

2.4.3. Effect of dose of Catalyst

To determine the role of the amount of catalyst, dosages of nanoferrites were varied from 0.05 g to 0.2 g. Different amounts of catalysts were taken in beakers containing 100 mL MB dye having a concentration of 10 mg/L and pH 11. This sample was kept in the photoreactor for a time duration of 120 min, and stirred by a magnetic stirrer (400 rpm). A plot of % degradation of methylene blue dye for various catalyst doses vs. time was drawn.

2.4.4. Effect of Initial Concentration of MB Dye Solution

The effect of the initial concentration on photodegradation process was studied by using the fixed photocatalyst dose of 0.1 g at varying initial concentrations of MB dye ranging from 5 mg/L to 30 mg/L, from a stock solution of 100 mg/L with pH 11 under visible light for 120 min. Later, the graph was plotted as the % degradation vs. time.

2.4.5. Effect of Intensity of Light

For this purpose, the variation in light intensity with respect to the changing distance between the light source and the solution containing the catalyst was investigated. 100 Watts and 200 Watts visible light bulbs were utilized for photocatalytic degradation by using 10 mg/L dye concentration, at pH 11 with 0.1 g catalyst dose for 120 min.

2.5. Photostability Studies

Photostability studies of the CoAlFe2O4 were carried out to determine its durability following the procedure described by Bashir et al. [21]. The FTIR spectrum of the unused photocatalyst was taken in the first step. In the next step, photocatalyst was employed for the degradation of dye, then washed and reused in the same manner for 5 consecutive cycles. The FTIR spectrum of the washed photocatalyst after 5 runs was taken again.

3. Results and Discussion

3.1. Synthesis of CoFe2O4 and Doped CoAlFe2O4

Synthesized catalyst was of a grayish-black powder and magnetic in nature, whose physical appearance is shown in Figure 2.

3.2. Characterization

3.2.1. SEM Analysis

Figure 3 and Figure 4 present the SEM images of the synthesized undoped and doped samples. From Figure 3, it can be observed that the particles show a uniform pattern in the doped sample as compared to the undoped sample. Moreover, the undoped sample is not in a well-ordered form while the doped sample is in a well-ordered form. Further from Figure 5, it can be observed that the average grain size slightly decreases upon doping of Al3+ ion from 46.51 nm for CoFe2O4 [22] to 36.21 nm for Co0.1Al0.03Fe0.17O0.4.

3.2.2. XRD Analysis

XRD patterns of cobalt ferrite CoFe2O4 is shown in Figure 6. All the peaks were indexed within a single-phase spinel structure with few percentages of α-Fe2O3. X-Powder Software was used to study the XRD pattern and crystalline phases were identified by correlation with reference data from JCPDS Card No. 22108 for Cobalt ferrites (CoFe2O4). The spectrum showed peaks at 2θ (23.48, 30.12, 35.46, 36.74, 43.07, 53.49, 56.81, 63.03) which are attributed to hkl values (111), (220), (311), (222), (400), (422), (511), (440) and proves the ferrite’s spinel cubic structure. The additional XRD peaks that can be seen in Figure 6 and Figure 7 were assigned to α-Fe2O3 (JCPDS Card No. 33–664) as an associated phase [23].
Figure 7 presents the XRD diffraction spectrum of Co0.1Al0.03Fe0.17O4. Generally, pure aluminum-doped cobalt ferrites exhibit seven peaks, which were located in the 2θ range of (10°–80°). It can also be observed that the most intense peak appeared at 2θ = 35° (311), which suggests the spinel structure of ferrites. The analysis using X-Powder software shows the crystalline phases of spinel ferrites have a perfect match with reference data of JCPDS Card No. 86-2267, and a small percentage of α-Fe2O3, which proves that cubic spinel lattice has been successfully doped with Al3+ ions. The XRD pattern shows peaks at 2 θ(23.74, 29.89, 35.52, 43.14, 54.16, 57.10, 62.99), which are accredited to (111), (220), (311), (400), (422), (511) and (440) of the spinel cubic structure of ferrite [24]. The peak (311) is the highly diffracted peak of cubic spinel ferrite and its intensity determines the crystallinity of this cubic structure. From the Debye–Scherrer formula, the average crystallite size of CoFe2O4 was 37 nm and 21.6 nm for Co0.1Al0.03Fe0.17O0.4, shown in Table 1. The lattice parameter decreases with the insertion of Al on the structure, from 8.38615 Å for the undoped to 8.3387 Å to the doped one. This effect is classic when a larger atom, Fe3+ (0.67 Å), is substituted by a smaller one, Al3+ (0.51 Å), thus changing all cell parameters.

3.2.3. Energy-Dispersive X-ray Spectroscopy

Figure 8 is representative of the EDS spectra showing the presence of Co, Al, Fe, and oxygen.

3.2.4. Nitrogen Adsorption/Desorption Studies

Nitrogen adsorption/desorption studies were performed to determine the surface area, pore volume, and pore radius (Table 2, Figure 9).
From the data in Table 2, it can be observed that a slight increase in the surface area occurred in the case of doped ferrite as compared to the undoped material.

3.2.5. FTIR of Cobalt Ferrite

The FTIR spectrum of cobalt ferrite is shown in Figure 10. The frequency bands at 443.91 cm−1 (95.04% T) and 530.33 cm−1 (93.23% T) were due to the stretching vibrations of metal oxides in octahedral group complex Co (II)-O2− and tetrahedral group complex Fe (III)-O2− of cobalt ferrite, which proves the spinel structure of ferrite. The frequency band at 787.72 cm−1 in the FTIR spectrum of cobalt ferrite was attributed to Co-O vibrations where Co ions were in a tetrahedral position. The band at 906.97 cm−1 is related to CoOOH hydration of water. The band at 1058.46 cm−1 and 1196.95 cm−1 was characteristic of the CoFe2O4 system and it was due to residual FeOOH. Bands that were present between 1500 cm−1 and 3000 cm−1 were due to the bending of the absorbed water molecules on the structure of spinel ferrite, i.e., 1584.24 cm−1 and 2105.27 cm−1. It was observed that differences in band positions were expected because of different bond lengths of Fe3+-O2−ions which were distributed at octahedral and tetrahedral sites [25].

3.2.6. FTIR of Aluminum-Doped Cobalt Ferrite

The FTIR spectrum of aluminum-doped cobalt ferrite shown in Figure 10 confirms the spinel structure of ferrite nanoparticles. The transmission spectrum of aluminum-doped cobalt ferrite shows two bands (υ1, υ2). The characteristic peak which was in the range of 400–600 cm−1 represents the presence of the M-O stretching vibration band. These bands were very susceptible to changes in the interaction between cations and oxygen in tetrahedral and octahedral sites [26]. The high-frequency band at 533.68 cm−1 shows stretching vibrations of tetrahedral groups. Lower frequency band near 406.01 cm−1 shows the stretching mode of octahedral M-O groups in ferrites. The reason behind the difference in the location of both frequency bands was due to different distances of Fe3+ ion in octahedral and tetrahedral complexes. Figure 11 illustrates that aluminum doping shifts lower frequency bands towards higher frequency bands.

3.3. Adsorption of MB on the Surface of Synthesized Catalysts

The adsorption phenomenon of MB on the surface of Co0.5Al0.3Fe1.7O0.4 is shown in Figure 12. Results, clearly, indicate that adsorption equilibrium was obtained in 30 min.

3.4. Photocatalytic Activity of Synthesized Catalysts

The photocatalytic activity of different synthesized catalysts is shown in Figure 13.
From Figure 13 it can be observed that undoped cobalt ferrite CoFe2O4 showed only 45% degradation efficiency while samples with 15% aluminum (Co0.1Al0.03Fe0.17O0.4) doped cobalt ferrite presented maximum degradation efficiency, however a further increase of aluminum doping decreases the degradation efficiency of the catalyst which might be due to distortion in the crystal structure of the catalyst. As Co0.1Al0.03Fe0.17O0.4 showed maximum efficiency, so aluminum-doped cobalt ferrite is taken for further optimization. Results were compared with the efficiency of other catalysts for degradation of MB but Co0.1Al0.03Fe0.17O0.4 showed a 10–15% higher efficiency [27,28,29].

3.5. Optimization of Different Variables Affecting the Degradation of MB Dye

3.5.1. Effect of Time

Illumination time is a key factor in an efficient photodegradation of dye. The impact of illumination time on dye’s percentage degradation was studied at a different time interval at pH 11 with a 0.5 g catalyst (shown in Figure 14). The addition of Al favors the formation of hole–electron pairs, which is linked to the formation of the radical intermediates, thus mediating the photodegradation process. Results confirmed that 120 min was enough to degrade dyes up to a maximum of 83% using the doped catalyst which was higher than that of undoped one (58%) [30,31].

3.5.2. Effect of pH

The pH of the solution plays a significant role in photodegradation as it affects the charge on the catalyst’s surface which in turn influences the adsorption of dye [32]. From Figure 15, it can be observed that maximum photocatalytic degradation occurred in alkaline pH and minimum in acidic medium. The maximum percentage of degradation was observed at pH 11. As discussed in the literature, it is due to the cationic nature of methylene blue dye that its surface becomes protonated in acidic pH which happens because of the adsorption of H+ ions. Therefore, it results in repulsion between dye molecules and photocatalyst. The increase in pH from acidic to alkaline causes a reduction in the repulsion between catalyst and dye molecules. This results in bringing the dye close to the catalytic surface, hence increasing photocatalytic degradation of dye. However, at pH higher than 11, the negative charge appears again on the surface of the catalyst and increase in repulsion between the catalyst’s surface and the dye may occur [33,34].
The role of pH in photodegradation of dye is associated with multiple phenomena [35]. The adsorption characteristics of the dye also change with pH and affect the surface characteristics of semiconductors [36]. Positive holes react with hydroxide ions which can form hydroxyl radicals. Hence, positive holes are the prime oxidation species at low pH, whereas at neutral or high pH levels, hydroxyl radicals are considered as the predominant species [37]. In alkaline solution, OH is easily generated by hydroxide ions present on the semiconductor surface [38]. It is also necessary to neutralize the pH of treated water by using different chemical methods.

3.5.3. Effect of Catalyst Dose

In photodegradation of dye; optimization of catalyst dose is shown in Figure 16. The optimum dose was calculated for the efficient removal of dye. The amount of catalyst required will depend upon the configuration of the reactor, light intensity, type of the dye targeted, type and morphology of particles of the catalyst [39,40].
From Figure 16, it is, clearly, observed that 0.1 g catalyst amount showed maximum degradation. Further increase in catalyst dose caused a decrease in photocatalytic degradation of dye. The decreasing efficiency with an increase in dose might be because the light penetration to all the catalyst sites will become difficult due to an increase in turbidity caused by an increasing amount of photocatalyst and light scattering effect.

3.5.4. Effect of Initial Concentration of MB Dye Solution

The optimization of the effect of the initial concentration of MB dye was also evaluated and results are presented in Figure 17.
It is obvious that 10 mg/L dye solution gave maximum degradation, however, above this, the increase in concentration caused a decreased degradation of methylene blue which might be due to unavailability of active sites for adsorption and decrease in hydroxyl OH radicals due to an increase in turbidity and as a result degradation decreased [41,42].

3.5.5. Effect of Intensity of Light

The results presented in Figure 18 describe that the photodegradation increases with an increase in the intensity of light [43,44,45] which might be due to the increase of OH radicals with high energy photons. The high intensity also results in a temperature rise which further helps the formation of free radicals.
Moreover, the photostability studies confirmed the durability of Co0.1Al0.03Fe0.17O0.4 as shown in Figure 19. Even after 5 consecutive cycles, no prominent change in the FTIR spectra taken before and after the cycles were observed which confirm its reusability. The most characteristic band appeared in a range of 400–600 cm−1 that represents the M-O stretching vibration band and can be observed in both the spectra. A minor band also emerged in Spectrum 2 which might be due to the presence of some minute amount of impurity left adsorbed on the catalyst’s surface.
Different photocatalysts have been reported in the literature for the purpose of photodegradation of methylene blue. The competency of aluminum-doped cobalt ferrite was found comparable to those already synthesized modified spinel ferrite. A comparison of the efficiency of the different photocatalyst is given in Table 3.

3.5.6. Kinetic Modeling

According to Zhang et al., the pseudo-first-order or second-order kinetic models are not only facile but can also accommodate any rate law expression [50]. Therefore, these two models were followed to determine the rate of reaction. Figure 20 presents the graphs obtained for both kinetic models. The regression coefficient (R2) value for the pseudo-first-order was 0.974, while for the second order, the obtained value was 0.998. It can hence be inferred from the regression R2 value, that the degradation followed second-order kinetics.

4. Conclusions

A series of aluminum-doped cobalt ferrites with compositional formula CoAlxFe2−xO4 (x = 0.1 to 0.5) was prepared and studied for the photodegradation of dye. The homogenous nanosized spinel ferrites were obtained through the sol-gel method. Due to Al3+ doping crystallite size and lattice parameter decreases and crystallite size found to be 21.6 nm to 37 nm. An FTIR spectrum confirms the cubic spinel structure of ferrites. Cobalt ferrite possessed a surface area of 9.217 m2/g whereas that of aluminum-doped cobalt ferrite was 14.865 m2/g. The SEM histograms confirmed that synthesis on nanoparticles ranging from 10 to 120 nm having maximum size distribution 40–50 nm EDS data confirm the presence of Co, Fe, O and Al participating cations. For maximum efficiency of doped aluminum spinel ferrite, 10 mg/L solution of MB dye, 0.1 g catalyst at pH 11 in 200 watts was required for 120 min. The synthesized NP may be used in wastewater treatment of dyes industry by designing different sieves using such adsorbents on the pathway of their wastewater.

Author Contributions

N.A. and S.M. supervised the research project. N.R. and N.S. carried out the experimental work as part of their Mphil theses. M.I.K., J.F.G., I.B.A., M.T., Z.A. and G.Y. contributed to the conceptualization, data analysis and interpretation, article editing and proof reading, as well as provision of resources for chemical characterization. All authors have read and agreed to the published version of the manuscript.

Funding

The author is grateful to Bahauddin Zikriya University, Multan, Pakistan and University of Leeds, UK for meeting the expenses related to this publication.

Conflicts of Interest

There are no potential conflict of interest required to be reported.

References

  1. Liyanage, C.P.; Yamada, K. Impact of population growth on the water quality of natural water bodies. Sustainability 2017, 9, 1405. [Google Scholar] [CrossRef] [Green Version]
  2. Vieira, R.H.; Volesky, B. Biosorption: A solution to pollution. Int. Microbial. 2000, 3, 17–24. [Google Scholar]
  3. Christie, R.M. Environmental Aspects of Textile Dyeing, 1st ed.; Woodhead Publishing Limited and CRC Press: Cambridge, UK, 2007. [Google Scholar]
  4. Kant, R. Textile dyeing industry an environmental hazard. Nat. Sci. 2012, 4, 22–26. [Google Scholar] [CrossRef] [Green Version]
  5. Ali, N.; Zada, A.; Zahid, M.; Ismail, A.; Rafiq, M.; Riaz, A.; Khan, A. Enhanced photodegradation of methylene blue with alkaline and transition-metal ferrite nanophotocatalysts under direct sun light irradiation. J. Chin. Chem. Soc. 2019, 66, 402–408. [Google Scholar] [CrossRef]
  6. Sulak, M.T.; Yatmaz, H.C. Removal of textile dyes from aqueous solutions with eco-friendly biosorbent. Desalin. Water Treat. 2012, 37, 169–177. [Google Scholar] [CrossRef]
  7. Arora, C.; Soni, S.; Sahu, S.; Mittal, J.; Kumar, P.; Bajpai, P.K. Iron based metal organic framework for efficient removal of methylene blue dye from industrial waste. J. Mol. Liq. 2019, 284, 343–352. [Google Scholar] [CrossRef]
  8. Jack Clifton, I.I.; Leikin, J.B. Methylene blue. Am. J. Ther. 2003, 10, 289–291. [Google Scholar] [CrossRef]
  9. Ekambaram, S.P.; Perumal, S.S.; Rajendran, D.; Samivel, D.; Khan, M.N. New approach of dye removal in textile effluent: A cost-effective management for clean up of toxic dyes in textile effluent by water hyacinth. In Toxicity and Biodegradation Testing; Humana Press: New York, NY, USA, 2018; pp. 241–267. [Google Scholar]
  10. Ge, L.; Liu, J. Efficient visible light-induced photocatalytic degradation of methyl orange by QDs sensitized CdS-Bi2WO6. Appl. Catal. B Environ. 2011, 105, 289–297. [Google Scholar] [CrossRef]
  11. Sobczyński, A.; Dobosz, A. Water purification by photocatalysis on semiconductors. Pol. J. Environ. Stud. 2001, 10, 195–205. [Google Scholar]
  12. Sanpo, N.; Wen, C.; Berndt, C.C.; Wang, J. New Approaches to the Study of Spinel Ferrite Nanoparticles for Biomedical Applications; Springer International Publishing: Cham, Switzerland, 2015. [Google Scholar]
  13. Sugimoto, M. The past, present, and future of ferrites. J. Am. Ceram. Soc. 1999, 82, 269–280. [Google Scholar] [CrossRef]
  14. Singh, A.K.; Zaidi, M.G.H.; Saxena, R.; Suman, R.; Maheshwari, P.; Gangwar, H.P.; Verma, A. Effect of surface treatment on tribological characteristic of ferrite nanoparticles epoxy composites. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2020; Volume 802, p. 012007. [Google Scholar]
  15. Harraz, F.A.; Mohamed, R.M.; Rashad, M.M.; Wang, Y.C.; Sigmund, W. Magnetic nanocomposite based on titania–silica/cobalt ferrite for photocatalytic degradation of methylene blue dye. Ceram. Int. 2014, 40, 375–384. [Google Scholar] [CrossRef]
  16. Cuiping, L.; Haodi, W.; Shijia, L.; Chen, L.; Binbin, C.; Qishe, Y. Preparation of magnetic Co0.5Zn0.5Fe2O4/AgBr hybrids for the visible-light-driven degradation of methyl orange. Mater. Sci. Semicond. Proc. 2018, 73, 67–71. [Google Scholar]
  17. López, J.; Camacho, M.; Solís, D.; González, C.; Carrillo, G.; Martinez, V.; Mijangos, R.; Cuevas, D. Toxicity assessment of cobalt ferrite nanoparticles on wheat plants. J. Toxicol. Environ. Health Part A 2018, 81, 604–619. [Google Scholar] [CrossRef] [PubMed]
  18. Kharisov, B.I.; Dias, H.R.; Kharissova, O.V. Mini-review: Ferrite nanoparticles in the catalysis. Arab. J. Chem. 2019, 12, 1234–1246. [Google Scholar] [CrossRef] [Green Version]
  19. Zayat, M.; Levy, D. Blue CoAl2O4 particles prepared by the sol−gel and citrate−gel methods. Chem. Mater. 2000, 12, 2763–2769. [Google Scholar] [CrossRef]
  20. Patterson, A.L. The Scherrer formula for X-ray particle size determination. Phys. Rev. 1939, 56, 978. [Google Scholar] [CrossRef]
  21. Bashir, A.; Hanif, F.; Yasmeen, G.; Mabood, F.; Hussain, A.; Abbas, N.; Manzoora, S. Polyaniline based magnesium nanoferrite composites as efficient photocatalysts for the photodegradation of Indigo Carmine in aqueous solutions. Desalin. Water Treat. 2019, 164, 368–377. [Google Scholar] [CrossRef]
  22. Mathew, D.S.; Juang, R.S. An overview of the structure and magnetism of spinel ferrite nanoparticles and their synthesis in microemulsions. Chem. Eng. J. 2007, 129, 51–65. [Google Scholar] [CrossRef]
  23. Gore, S.K.; Jadhav, S.S.; Jadhav, V.V.; Patange, S.M.; Naushad, M.; Mane, R.S.; Kim, K.H. The structural and magnetic properties of dual phase cobalt ferrite. Sci. Rep. 2017, 7, 1–9. [Google Scholar] [CrossRef]
  24. Mathubala, G.; Manikandan, A.; Antony, S.A.; Ramar, P. Photocatalytic degradation of methylene blue dye and magneto-optical studies of magnetically recyclable spinel NixMn1-xFe2O4 (x = 0.0–1.0) nanoparticles. J. Mol. Struct. 2016, 1113, 79–87. [Google Scholar] [CrossRef]
  25. Kambale, R.C.; Song, K.M.; Koo, Y.S.; Hur, N. Low temperature synthesis of nanocrystalline Dy3+ doped cobalt ferrite: Structural and magnetic properties. J. Appl. Phys. 2011, 110, 053910. [Google Scholar] [CrossRef]
  26. Bhavyashri, P.S.; Raveendra, R.S.; Jayasheelan, A.; Prakash, C.S.; Prasad, B.D.; Rewathkar, K.G.; Nagabhushana, B.M. Effect of temperature on electrical properties of auto combustion derived nano calcium ferrite. Int. J. Adv. Res. 2015, 5, 153–161. [Google Scholar]
  27. Kalam, A.; Al-Sehemi, A.G.; Assiri, M.; Du, G.; Ahmad, T.; Ahmad, I.; Pannipara, M. Modified solvothermal synthesis of cobalt ferrite (CoFe2O4) magnetic nanoparticles photocatalysts for degradation of methylene blue with H2O2/visible light. Results Phys. 2018, 8, 1046–1053. [Google Scholar] [CrossRef]
  28. Mishra, D.; Senapati, K.K.; Borgohain, C.; Perumal, A. CoFe2O4 Fe3O4 Magnetic nanocomposites as photocatalyst for the degradation of methyl orange dye. J. Nanotechnol. 2012, 1, 1–6. [Google Scholar] [CrossRef] [Green Version]
  29. Paliwal, A.; Banu, R.; Ameta, R.; Ameta, S.C. Photocatalytic degradation of methylene blue using undoped and Co-doped bismuth ferrite. J. Appl. Chem. 2017, 6, 967–975. [Google Scholar]
  30. Cabrera, A.F.; Torres, C.R.; Marchetti, S.G.; Stewart, S.J. Degradation of methylene blue dye under dark and visible light conditions in presence of hybrid composites of nanostructured MgFe2O4 ferrites and oxygenated organic compounds. J. Environ. Chem. Eng. 2020, 8, 104274. [Google Scholar] [CrossRef]
  31. Kefeni, K.K.; Mamba, B.B. Photocatalytic application of spinel ferrite nanoparticles and nanocomposites in wastewater treatment. SMT 2020, 23, e00140. [Google Scholar] [CrossRef]
  32. Fox, M.A.; Dulay, M.T. Heterogeneous photocatalysis. Chem. Rev. 1993, 93, 341–357. [Google Scholar] [CrossRef]
  33. Alene, A.N.; Abate, G.Y.; Adere, T.H. Bioadsorption of Basic Blue Dye from Aqueous Solution onto Raw and Modified Waste Ash as Economical Alternative Bioadsorbent. J. Chem. 2020, 2020, 8746035. [Google Scholar] [CrossRef]
  34. Zaini, M.A.A.; Sudi, R.M. Valorization of Human hair as mythelene blue dye adsorbents. Green Process Synth. 2017, 7, 344–352. [Google Scholar] [CrossRef]
  35. Konstantinou, I.K.; Albanis, T.A. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: Kinetic and mechanistic investigations: A review. Appl. Catal. B Environ. 2004, 49, 1–14. [Google Scholar] [CrossRef]
  36. Sharma, R.; Singhal, S. Photodegradation of textile dye using magnetically recyclable heterogeneous spinel ferrites. J. Chem. Technol. Biotechnol. 2015, 90, 955–962. [Google Scholar] [CrossRef]
  37. Tunesi, S.; Anderson, M. Influence of chemisorption on the photodecomposition of salicylic acid and related compounds using suspended titania ceramic membranes. J. Phys. Chem. 1991, 95, 3399–3405. [Google Scholar] [CrossRef]
  38. Tang, W.Z.; An, H. UV/TiO2 photocatalytic oxidation of commercial dyes in aqueous solutions. Chemosphere 1995, 31, 4157–4170. [Google Scholar] [CrossRef]
  39. Saquib, M.; Muneer, M. Semiconductor mediated photocatalysed degradation of an anthraquinone dye, Remazol Brilliant Blue R under sunlight and artificial light source. Dyes Pigm. 2002, 53, 237–249. [Google Scholar] [CrossRef]
  40. Reddy, E.P.; Davydov, L.; Smirniotis, P. TiO2-loaded zeolites and mesoporous materials in the sonophotocatalytic decomposition of aqueous organic pollutants: The role of the support. Appl. Catal. B Environ. 2003, 42, 1–11. [Google Scholar] [CrossRef]
  41. Ilinoiu, E.C.; Pode, R.; Manea, F.; Colar, L.A.; Jakab, A.; Orha, 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]
  42. Kansal, S.K.; Singh, M.; Sud, D. Studies on TiO2/ZnO photocatalysed degradation of lignin. J. Hazard. Mater. 2008, 153, 412–417. [Google Scholar] [CrossRef]
  43. Curcó, D.; Giménez, J.; Addardak, A.; Cervera-March, S.; Esplugas, S. Effects of radiation absorption and catalyst concentration on the photocatalytic degradation of pollutants. Catal. Today 2002, 76, 177–188. [Google Scholar] [CrossRef]
  44. Karunakaran, C.; Senthilvelan, S. Photooxidation of aniline on alumina with sunlight and artificial UV light. Catal. Commun. 2005, 6, 159–165. [Google Scholar] [CrossRef]
  45. Faisal, M.; Harraz, F.A.; Jalalah, M.; Alsaiari, M.A.; Al-Sayari, S.A.; Al-Assiri, M.S. Polythiophene doped ZnO nanostructures synthesized by modified sol-gel and oxidative polymerization for efficient photodegradation of methylene blue and gemifloxacin antibiotic. Mater. Today Commun. 2020, 24, 101048. [Google Scholar] [CrossRef]
  46. Arifin, M.; Karim, K.; Abdullah, H.; Khan, M. Synthesis of Titania Doped Copper Ferrite Photocatalyst and Its Photoactivity towards Methylene Blue Degradation under Visible Light Irradiation. Bull. Chem. React. Eng. Catal. 2019, 14, 219–227. [Google Scholar] [CrossRef] [Green Version]
  47. Javed, H.; Rehman, A.; Mussadiq, S.; Shahid, M.; Khan, M.A.; Shakir, I.; Warsi, M.F. Reduced graphene oxide-spinel ferrite nano-hybrids as magnetically separable and recyclable visible light driven photocatalyst. Synth. Met. 2019, 254, 1–9. [Google Scholar] [CrossRef]
  48. Harifi, T.; Montazer, M. A novel magnetic reusable nanocomposite with enhanced photocatalytic activities for dye degradation. Sep. Purif. Technol. 2014, 134, 210–219. [Google Scholar] [CrossRef]
  49. Ciocarlan, R.G.; Seftel, E.M.; Mertens, M.; Pui, A.; Mazaj, M.; Tusar, N.N.; Cool, P. Novel magnetic nanocomposites containing quaternary ferrites systems Co0. 5Zn0. 25M0. 25Fe2O4 (M = Ni, Cu, Mn, Mg) and TiO2-anatase phase as photocatalysts for wastewater remediation under solar light irradiation. Mater. Sci. Eng. B 2018, 230, 1–7. [Google Scholar] [CrossRef]
  50. Zhang, H.; Liu, D.; Ren, S.; Zhang, H. Kinetic studies of direct blue photodegradation over flower-like TiO2. Res. Chem. Intermed. 2017, 43, 1529–1542. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of photocatalytic degradation of dye.
Figure 1. Schematic representation of photocatalytic degradation of dye.
Water 12 02285 g001
Figure 2. The final appearance of the catalyst.
Figure 2. The final appearance of the catalyst.
Water 12 02285 g002
Figure 3. SEM image of undoped CoFe2O4 at a resolution of 5 µm.
Figure 3. SEM image of undoped CoFe2O4 at a resolution of 5 µm.
Water 12 02285 g003
Figure 4. SEM images of Co0.1Al0.03Fe0.17O0.4 at a resolution of 5 µm.
Figure 4. SEM images of Co0.1Al0.03Fe0.17O0.4 at a resolution of 5 µm.
Water 12 02285 g004
Figure 5. Histogram of distribution of particles of CoFe2O4 and Co0.1Al0.03Fe0.17O0.4.
Figure 5. Histogram of distribution of particles of CoFe2O4 and Co0.1Al0.03Fe0.17O0.4.
Water 12 02285 g005
Figure 6. X-ray diffraction spectra of CoFe2O4.
Figure 6. X-ray diffraction spectra of CoFe2O4.
Water 12 02285 g006
Figure 7. X-ray diffraction spectra of Co0.1Al0.03Fe0.17O0.4.
Figure 7. X-ray diffraction spectra of Co0.1Al0.03Fe0.17O0.4.
Water 12 02285 g007
Figure 8. Energy-dispersive X-ray spectra of CoFe2O4 and Co0.1Al0.03Fe0.17O0.4.
Figure 8. Energy-dispersive X-ray spectra of CoFe2O4 and Co0.1Al0.03Fe0.17O0.4.
Water 12 02285 g008
Figure 9. Nitrogen adsorption/desorption studies of CoFe2O4 and Co0.1Al0.03Fe0.17O0.4.
Figure 9. Nitrogen adsorption/desorption studies of CoFe2O4 and Co0.1Al0.03Fe0.17O0.4.
Water 12 02285 g009
Figure 10. FTIR of cobalt ferrite CoFe2O4.
Figure 10. FTIR of cobalt ferrite CoFe2O4.
Water 12 02285 g010
Figure 11. FTIR of Co0.5Al0.3Fe1.7O0.4.
Figure 11. FTIR of Co0.5Al0.3Fe1.7O0.4.
Water 12 02285 g011
Figure 12. Adsorption of methylene blue (MB) on Co0.5Al0.3Fe1.7O0.4.
Figure 12. Adsorption of methylene blue (MB) on Co0.5Al0.3Fe1.7O0.4.
Water 12 02285 g012
Figure 13. The plot of % Photodegradation of MB dye by different photocatalysts.
Figure 13. The plot of % Photodegradation of MB dye by different photocatalysts.
Water 12 02285 g013
Figure 14. % degradation of MB dye by undoped and doped photocatalysts.
Figure 14. % degradation of MB dye by undoped and doped photocatalysts.
Water 12 02285 g014
Figure 15. Percentage degradation of MB dye at different pH in varying time intervals using Co0.1Al0.03Fe0.17O0.4.
Figure 15. Percentage degradation of MB dye at different pH in varying time intervals using Co0.1Al0.03Fe0.17O0.4.
Water 12 02285 g015
Figure 16. The plot of percentage degradation of MB by varying catalyst amounts in different time intervals.
Figure 16. The plot of percentage degradation of MB by varying catalyst amounts in different time intervals.
Water 12 02285 g016
Figure 17. Percentage photodegradation of MB at varying concentrations of dye with respect to time.
Figure 17. Percentage photodegradation of MB at varying concentrations of dye with respect to time.
Water 12 02285 g017
Figure 18. Percentage photodegradation of MB at various light intensities; in the presence of 100 and 200 watts visible light bulb.
Figure 18. Percentage photodegradation of MB at various light intensities; in the presence of 100 and 200 watts visible light bulb.
Water 12 02285 g018
Figure 19. FTIR spectra of Co0.1Al0.03Fe0.17O4 before (1) and after 5 cycles (2) using as photocatalyst.
Figure 19. FTIR spectra of Co0.1Al0.03Fe0.17O4 before (1) and after 5 cycles (2) using as photocatalyst.
Water 12 02285 g019
Figure 20. Plots of (a) pseudo-first-order (b) second-order kinetic models for photocatalytic degradation of methylene blue using Co0.1Al0.03Fe0.17O0.4.
Figure 20. Plots of (a) pseudo-first-order (b) second-order kinetic models for photocatalytic degradation of methylene blue using Co0.1Al0.03Fe0.17O0.4.
Water 12 02285 g020aWater 12 02285 g020b
Table 1. XRD results of CoFe2O4 and Co0.1Al0.03Fe0.17O0.4.
Table 1. XRD results of CoFe2O4 and Co0.1Al0.03Fe0.17O0.4.
Sr. No.SampleLattice Parametera (Å)Average Crystalline Size (nm)Volume of Cell (Å3)Inter Planar Spacing (d)
1CoFe2O48.3861537589.772.515
2Co0.1Al0.03Fe0.17O0.48.338721.6579.822.501
Table 2. Nitrogen adsorption/desorption results of undoped and doped ferrites.
Table 2. Nitrogen adsorption/desorption results of undoped and doped ferrites.
SampleBET Surface Area
(m2/g)
Average Pore Radius
(Å)
Total Pore Volume
(cm3/g)
CoFe2O49.21717.800.013
Co0.1Al0.03Fe0.17O0.414,86527.691.261.26
Table 3. A comparison of different modified spinel ferrite with Co0.1Al0.3Fe1.7O0.4.
Table 3. A comparison of different modified spinel ferrite with Co0.1Al0.3Fe1.7O0.4.
Types of Synthesized PhotocatalystSize (nm)Pollutant/spHContact Time (min)Catalyst Dose (g/L)Pollutant (mg/L)Light SourceRemoval CapacityCycleRef.
CuFe2O4@TiO264MB-1800.520Vis40%3[46]
Ni0.65Zn0.35Fe2O4-MB-5615UV–Vis55%-[47]
Fe3O4/TiO235MB-2400.210UV50%3[48]
Co0.5Zn0.25Ni0.25Fe2O4–TiO2-MB-80112.79UV–Vis12.66%-[49]
Co0.1Al0.3Fe1.7O0.440MB111200.510Vis83%5Current work

Share and Cite

MDPI and ACS Style

Abbas, N.; Rubab, N.; Sadiq, N.; Manzoor, S.; Khan, M.I.; Fernandez Garcia, J.; Barbosa Aragao, I.; Tariq, M.; Akhtar, Z.; Yasmin, G. Aluminum-Doped Cobalt Ferrite as an Efficient Photocatalyst for the Abatement of Methylene Blue. Water 2020, 12, 2285. https://doi.org/10.3390/w12082285

AMA Style

Abbas N, Rubab N, Sadiq N, Manzoor S, Khan MI, Fernandez Garcia J, Barbosa Aragao I, Tariq M, Akhtar Z, Yasmin G. Aluminum-Doped Cobalt Ferrite as an Efficient Photocatalyst for the Abatement of Methylene Blue. Water. 2020; 12(8):2285. https://doi.org/10.3390/w12082285

Chicago/Turabian Style

Abbas, Naseem, Nida Rubab, Natasha Sadiq, Suryyia Manzoor, Muhammad Imran Khan, Javier Fernandez Garcia, Isaias Barbosa Aragao, Muhammad Tariq, Zeeshan Akhtar, and Ghazala Yasmin. 2020. "Aluminum-Doped Cobalt Ferrite as an Efficient Photocatalyst for the Abatement of Methylene Blue" Water 12, no. 8: 2285. https://doi.org/10.3390/w12082285

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