Next Article in Journal
The Effects of Berberine on the Gut Microbiota in Apc min/+ Mice Fed with a High Fat Diet
Previous Article in Journal
Three New Iridoid Derivatives Have Been Isolated from the Stems of Neonauclea reticulata (Havil.) Merr. with Cytotoxic Activity on Hepatocellular Carcinoma Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Molybdenum Trioxide: Efficient Nanosorbent for Removal of Methylene Blue Dye from Aqueous Solutions

1
Chemistry Department, College of Science, Taibah University, Al-Madinah 30002, Saudi Arabia
2
Département de Chimie, Faculté des Sciences Dhar El Mahraz, Université Sidi Mohamed Ben Abdellah, B. P. 1796 (Atlas), Fès 30003, Morocco
3
Community College, Taibah University-Al-Mahd Branch, Al-Mahd 42112, Saudi Arabia
*
Author to whom correspondence should be addressed.
Molecules 2018, 23(9), 2295; https://doi.org/10.3390/molecules23092295
Submission received: 11 August 2018 / Revised: 2 September 2018 / Accepted: 6 September 2018 / Published: 8 September 2018

Abstract

:
Nano Molybdenum trioxide (α-MoO3) was synthesized in an easy and efficient approach. The removal of methylene blue (MB) in aqueous solutions was studied using this material. The effects of various experimental parameters, for example contact time, pH, temperature and initial MB concentration on removal capacity were explored. The removal of MB was significantly affected by pH and temperature and higher values resulted in increase of removal capacity of MB. The removal efficiency of Methylene blue was 100% at pH = 11 for initial dye concentrations lower than 150 ppm, with a maximum removal capacity of 152 mg/g of MB as gathered from Langmuir model. By comparing the kinetic models (pseudo first-order, pseudo second-order and intraparticle diffusion model) at various conditions, it has been found that the pseudo second-order kinetic model correlates with the experimental data well. The thermodynamic study indicated that the removal was endothermic, spontaneous and favorable. The thermal regeneration studies indicated that the removal efficiency (99%) was maintained after four cycles of use. Fourier Transform Infrared (FTIR) and Scanning Electron Microscopy (SEM) confirmed the presence of the MB dye on the α-MoO3 nanoparticles after adsorption and regeneration. The α-MoO3 nanosorbent showed excellent removal efficiency before and after regeneration, suggesting that it can be used as a promising adsorbent for removing Methylene blue dye from wastewater.

Graphical Abstract

1. Introduction

Dyes are organic pollutants that have a complex chemical structure, are highly stable; resist washing, light and microbial invasions and poorly biodegradable [1,2,3,4]. They are harmful to aquatic life and humans and their removal is of significant importance [5,6,7,8].
Several methods were performed for dye removal from industrial effluents and wastewater including flocculation, coagulation, adsorption, ion exchange, membrane separation, photodegradation, extraction, chemical oxidation and biological treatment [9,10,11,12,13,14,15]. Adsorption proposes the advantages of effectiveness, simplicity and low cost from among those above-mentioned approaches. [1,16,17,18,19,20,21].
Several natural and synthetic substances were reported earlier in the literature as adsorbents for organic dyes [22,23,24,25,26,27,28,29,30,31]. The adsorption performance of biosorbents is usually restricted by the low surface area, which results in low adsorption capacities [32]. Activated carbon (AC), from agricultural and solid wastes as the nontoxic and easily available adsorbent, is considered as a general adsorbent for removing pollutants such as organic dyes from wastewater due to its porous structure, high surface areas, fast adsorption kinetics, large adsorption capacities and general material as a support for loading nanomaterials [33,34]. However, AC is still considered highly expensive based on the market price of the commercial activated carbon available. In addition, its poor mechanical and regeneration properties have limited its use in the adsorption process. [21,28,29].
Recently, nanomaterials as synthetic adsorbents have attracted a lot of research interest because of their distinctive properties such as electron conduction, large surface area, highly active sites, low mass used and the ability to modify their surface properties [35,36]. The nanomaterials are grouped in different categories such as metal oxide, carbonaceous, bio or magnetic nanomaterials. They have been widely studied as removal agents for dyes [3,5,22,27,30,35,36,37,38,39,40]. Some examples of metal oxides nanomaterials used for dyes removal are Titanium dioxide [41], Zinc oxide [42], Magnesium oxide [43] and Magnetic iron oxide [14].
The nanoparticles are synthesized by various methods, which are categorized as three types, namely chemical, physical and mechanical processes [44]. The chemical process involves the use of chemistry solutions, making this process, not suitable for large scale production, due to its high expenses and slow to manufacture [45,46,47].
Molybdenum can be found in several oxide stoichiometries, which have been used for a variety of high-value research and commercial applications [48]. Furthermore, MoO3 is a polymorph material with at least four known phases monoclinic (β-MoO3), orthorhombic (α-MoO3), high pressure monoclinic (MoO3-II) and hexagonal (h-MoO3) [49,50,51,52]. Due to the outstanding electrochemical and catalytic activities, α-MoO3 has been widely considered [48,53,54]. Thus far, a number of α-MoO3 nanostructures were synthesized including nanobelts, nanoparticles, nanosheets, flower-like hierarchical structures and nanoflakes [49,55,56,57,58,59,60,61,62,63]. However, few studies are reported on the use of Molybdenum trioxide for removing dyes. Beltran et al. [64] reported that hexagonal and orthorhombic phases of MoO3 nanoparticles synthesized using microwave radiation followed by high-energy mechanical milling were used for Methylene blue (MB) removal. Approximately a 98% of MB was removed from 20 ppm content in water, without using photon radiation in about 25 min [64]. Huge challenge is seeking to the development of nanomaterials, easily synthetized and presenting high performance criteria for removal of dyes and regeneration [22,36,65].
In our previous work, Molybdenum trioxide (α-MoO3) nanorods and stacked nanoplates were synthesized easily and efficiently at a rather low temperature with the use of a simple and economical approach [61,66]. In this study, the capacity of the materials of interest were tested to remove methylene blue dye (MB) from aqueous solutions. The methylene blue dye is classified as a prior pollutant due to its broad usage in various industrial applications, for example coloring agents for cotton, leather, wool and silk and so forth [67]. For this purpose, the effect of a variety of parameters such as adsorbent dose, contact time, pH, initial dye concentrations and temperature were evaluated. The thermodynamic and kinetic studies were performed. The experimental equilibrium data was examined using Temkin, Freundlich, Langmuir and Dubinin–Radushkevich models. Thermal regeneration of α-MoO3 nanosorbent was also studied.

2. Experimental

2.1. Preparation of Molybdenum Trioxide Nanosorbent

All chemicals were purchased from Sigma-Aldrich (St. Louis, MO, USA) and used as received without any changes, except for the methylene blue (MB) dye, which was supplied by Panreac, Barcelona, Spain.
Molybdenum trioxide nanosorbent (α-MoO3) was synthesized using the thermal decomposition of an oxalic precursor of Molybdenum gained from the reaction of oxalic acid and ammonium molybdate (NH4)6Mo7O24·4H2O in the solid state, as described in our earlier work [61]. Oxalic acid and ammonium molybdate (NH4)6Mo7O24·4H2O were mixed together in a ratio of Mo:acid of 1:3. The mixture was ground then heated on a hot plate at 160 °C. Then, the oxalic precursor was decomposed at 350 °C in a tubular furnace open on both ends.

2.2. Adsorption Experiments

The removal of MB was carried out by batch adsorption experiments [68]. The removal of MB by α-MoO3 was carried out by stirring specific amount of adsorbent into 100 mL of MB solution of known concentrations at specific temperature (T = 25, 50 and 70 °C) and at different contact times (10, 30, 60, 90 and 120 min). At the end of predetermined time intervals, the solution was filtrated with a 0.45 µm syringe filter (Whatman, Sigma-Aldrich, St. Louis, MO, USA) and examined using a UV-Visible spectrometer (Thermo Fisher Scientific, Madison, WI, USA) at λmax = 665 nm. The pH of the MB solution was adjusted by adding either 0.01 N NaOH or 0.01 N HCl solutions. The percentage of removal (%) and the removed amount of MB at equilibrium qe (mg/g) were calculated using the following relationships.
Removal   % = C i C f C i × 100
q e = ( C i C f ) M × V
where Ci and Cf represent the initial and equilibrium concentration of MB (ppm), respectively. V is the used volume of solution (L) and M is the added mass of α-MoO3 (g). The results were repeated three times and the uncertainty was about 3%.

2.3. Adsorbent Regeneration Method

For the regeneration experiments, a solution of 150 ppm was used and the removal equilibrium time was extended for 2 h. The fresh spent α-MoO3 was filtered, dried at 100 °C and calcined at 400 °C for 1 h, under air atmosphere. The calcined α-MoO3 was tested again at the same conditions. The regeneration process was repeated for three cycles.

2.4. Characterization

The powder characterization in terms of the phase composition of the synthetized α-MoO3 nanosorbent, was analyzed by XRD (X-ray diffractometer 6000, Shimadzu, Tokyo, Japan, installed with λCu-Kα = 1.5406 Ǻ and Ni filter). The specific surface area was deduced from the nitrogen isotherm adsorption and using the BET equation (DBET = 6000/d.S, where S is the specific surface area and d is the density), as reported in our previous work [61]. The specific surface area value was 41.02 m2/g.
The presence of MB dye on the α-MoO3 nanoparticles after the adsorption and regeneration studies was confirmed by FTIR spectroscopy using IR Affinity-1S Shimadzu apparatus (Shimadzu, Tokyo, Japan) in the range of 400 and 4000 cm−1 using KBr pellets. Scanning electron microscope (SEM) analysis was performed using Quanta Feg 250 (Thermo Fisher Scientific, Hillsboro, OR, USA). The concentration at equilibrium was determined using UV-Visible spectrophotometer (Thermo Scientific Genesys 10S, Madison, WI, USA).

3. Results and Discussion

3.1. Removal of MB

3.1.1. Effect of Initial Dye Concentration and Contact Time

The effect of contact time and initial dye concentration on the removal of MB dye was studied and presented in Figure 1. The removal of MB increases with the increase of contact time and reaches a maximum value of 99% at about 30 min for initial MB concentrations of 10, 20 and 30 ppm and 120 min of contact time for initial dye concentration of 40 ppm. The removal capacity was improved from 19 mg/g to 42 mg/g when the initial dye concentrations increased from 20 ppm to 50 ppm, respectively. These results can be clarified by the primarily great availability of vacant sites on the α-MoO3 surface, which steadily decreases as the sites are filled up over time as a result of the sorption process [69].

3.1.2. Effect of Adsorbent Dose and Initial Dye Concentration

The adsorbent dose is a very important parameter in the adsorption process [70]. The removal of MB using α-MoO3 was investigated by varying the adsorbent dose from 1.0 to 4.0 g/L and the initial dye concentrations from 30 to 60 ppm (Figure 2).
For lower initial concentrations less than 50 ppm, 2 g/L of adsorbent dose was needed to achieve 99% of MB removal percentage. However, for 60 ppm, 3 g/L was the minimum adsorbent needed to obtain 99% of removal efficiency.
The amount of MB removed decreased with respect to an increase of adsorbent dose and this is shown in Figure 2. This is due to the increase of the available active sites on the adsorbents’ surface area. These results can be explained by the availability of more active sites as the adsorbent dose increased [70].

3.1.3. Temperature Effect

As the temperature has a great effect on removing dyes [71], an investigation was carried out on temperature as a parameter on its own from 25 to 70 °C during the process of removing the MB dye, this can be seen in Figure 3. The percentage removal of MB (at Ci = 40 ppm) has gone up from 82% to 99% and the removal capacity has increased from 33 mg/g to 39 mg/g. In actual fact, the removal activity of the adsorbent sites enhanced as the temperature increased giving rise to the dye molecule motion [71,72].
Thermodynamic factors are important in the adsorption process [73,74]. The likelihood and the mechanism of adsorption can be projected in reference to the thermodynamic factors [73]. Thermodynamic parameters can be deduced using the following equations:
Δ G o = RTLnK d
K d = C a C e
LnK d = Δ S o R Δ H o RT
where R is the gas constant (J·mol−1·K−1), ΔG° is the free energy (KJ·mol−1), Kd is the distribution constant, T is absolute temperature (K), Ce is the equilibrium concentration (mol/L), Ca is the amount of dye adsorbed on the adsorbent at equilibrium (mol/L), ΔH° is the standard enthalpy (KJ·mol−1) and ΔS° is the standard entropy (KJ·mol−1·K). ∆S° and ∆H° values were achieved from the intercept and slope of plot lnKd versus 1/T and presented in Figure 4 (The value of the regression correlation coefficients (R2) is 0.83). ∆G° values were obtained from Equation (3) and presented in Table 1. The adsorption is favorable and spontaneous, indicated by the negative value of ∆G°. ∆H° value indicates that MB removal occurred in a physisorption process as indicated by the positive value of ∆H° (90 KJ mol−1) [75]. The increased disorder and randomness at the solid solution interface of MB and α-MoO3 is indicated by the positive values of ∆S°. The adsorbed water molecules are displaced by the adsorbate molecules and therefore more translational energy is gained than is lost, this leads the system occurring randomly [76].

3.1.4. Effect of pH

pH is an essential element that controls the removal of dyes [71]. Consequently, the effect of pH for the removal of MB using α-MoO3 nanosorbent was studied by variable pH values from 2.5 to 11 at temperature of 25 °C and initial concentration of 40 ppm. As presented in Figure 5, the MB removal is evidently pH dependent. The percentage removal increases from 47% to 99% as pH increases from 2.5 to 11. The amount of dye removed per unit mass of adsorbent at equilibrium (qe) increased from 19 to 40 mg/g by variation of pH from 2.5 to 11. At pH = 11 the hydroxyl group (OH) in solution favors the positive charge of the MB since its pKa equals 3.8 [77]. Therefore, pH = 11 was considered as the optimum value for MB removal using α-MoO3 nanosorbent.

3.1.5. Effect of MB Initial Dye Concentration and Contact Time after pH Adjustment

The removal efficiency of α-MoO3 was examined for higher concentrations of methylene blue dye at pH = 11 as presented in Figure 6. Interestingly, the percent of removal of MB was 100% after 60 min and 120 min for initial dye concentrations of 100 and 150 ppm, respectively. The removed amount of MB was 100 mg/g for initial dye concentrations of 100 ppm and 150 mg/g for initial dye concentrations of 150 and 250 ppm.

3.2. Kinetic Study

The kinetic models based on the removal capacity were fitted to experimental data to determine the rates of adsorption for MB dye molecules and to investigate the mechanism of the removal process [78].
The data obtained from the kinetics of removing MB using 0.1 g of α-MoO3 nanosorbent at room temperature and pH = 11 was analyzed by pseudo first-order (PFO), pseudo second-order (PSO) and intraparticle diffusion (IPD) kinetic models. The equations of the studied models are given in Table 2.
The three model parameters, pseudo first, pseudo second and intra-particle diffusion are tabulated in Table 3 and presented in Figure 7, Figure 8 and Figure 9 respectively. The three models differ in their regression correlation coefficients (R2). Pseudo first ranges from 0.995 to 0.997, whereas Pseudo second is 0.998 to 1.000 and intra-particle is 0.832 to 0.910, with their different concentrations used. The R2 for pseudo second-order is close to 1 and hence this model fitted well the experimental data.

3.3. Adsorption Isotherms

To optimize the design of a removal system for the MB molecules, various isotherm equations have been used to describe the equilibrium characteristics of the removal process [81]. Four adsorption models were investigated, namely Freundlich, Langmuir, Temkin isotherm and Dubinin–Radushkevich models. The equations for the four tested models are summarized in Table 4.
Langmuir, Freundlich, D–R isotherm and Temkin models were applied to fit the experimental data. The values of the regression correlation coefficients (R2) and the model parameters are included within Table 5 and shown in Figure 10. Langmuir equation showed the highest value of R2 (1.000) and D–R model showed the lowest value of R2 (0.939), whereas intermediate values were achieved for Temkin and Freundlich (0.989 and 0.997 respectively). Langmuir model fits wells with the experimental data and the MB removal took place on homogenous surface forming a monolayer on the α-MoO3 adsorbent, with high adsorption capacity of 152 mg/g. MB dye removal by α-MoO3 is favorable which is indicated by the separation factor RL ranging from 0.0007 to 0.0090.
The comparative links between α-MoO3 and other sorbents presented in this work are shown in Table 6. The Molybdenum trioxide (α-MoO3) nanorods and stacked nanoplates synthesized easily and efficiently at rather low temperature with the use of simple and economical approach [61,66] showed high removal capacity. In addition, the molybdenum trioxide is presenting the advantage to be successfully regenerated as it will be presented in this paper. Moreover, no modification is needed for the molybdenum trioxide because it is used as prepared which is not the case when using supported gold nanoparticles or when using nanotubes. Another important point to raise is that the mass production of the MoO3 is possible as the production can be done easily at higher scale.

3.4. Regeneration and Characterization of the Nanosorbent

3.4.1. Regeneration Efficiency

The regeneration and repeatability of the adsorbent are very critical for the practical application. Many regeneration procedures were proposed in the literature survey, including thermal treatment, chemical extraction, bio-regeneration, supercritical regeneration, microwave irradiation and so forth. Thermal regeneration is often applied for regeneration of exhausted activated carbon [91]. In our case, the structure of α-MoO3 removal agent was stable and the thermal treatment method was selected in this part.
It is found that α-MoO3 could be regenerated through thermal treatment. The MB removal efficiency of α-MoO3 was maintained after three cycles of regeneration with an average of 99% as presented in Figure 11. The high removal efficiency indicated that the regeneration of the adsorbent by calcination under air atmosphere at 400 °C was highly efficient and suggesting an excellent reusability.

3.4.2. Fourier-Transform Infrared Spectroscopy

In order to fully recognize the MB removal process by α-MoO3 nanosorbent, the materials exposed to MB were studied by IR spectroscopy. Figure 12 shows the FTIR spectra of the α-MoO3 sample before and after removal of MB dye. As seen, the characteristic stretching and flexing vibrations of the metal–oxygen bonds at 991, 880, 820, 513, 486 and a broad centered at 623 cm−1, corresponded to Molybdenum trioxide [92]. The FTIR spectrum of pure MB exhibited bands between 1700 and 1000 cm−1 [93]. While, the FTIR spectrum of α-MoO3 after adsorption of MB (MoO3-MB1) exhibited additional bands located at 1600 cm−1, related to C=C stretching of MB, due to the presence of MB attached to the active sites of α-MoO3 [94]. The FTIR spectrum of the regenerated α-MoO3 (MoO3-R) after thermal treatment was similar to the fresh α-MoO3. The reused sample (MoO3-MB2) exhibited again all bands characteristic of the MB [93]. The obtained spectrum confirmed the efficiency of the reused adsorbent.

3.4.3. Scanning Electron Microscope (SEM) Analysis

It is interesting to follow up the evolution of the α-MoO3 morphology at different steps of the adsorption test. The SEM micrograph in Figure 13A indicated that the α-MoO3 particles exhibited sponge like structure, of dimensions varying from 5 to 10 microns. After removal of MB molecules, the sponge-like structure vanished and the pores were stuffed by the removed molecules (Figure 13B). Figure 13C,D indicated that the morphology of the sample was not altered after regeneration and the first reuse. In both cases the particles are less agglomerated with aggregates less than 1 micron in size. In overall, the morphology of α-MoO3 was not significantly modified even after the second reuse in Figure 13E.

3.4.4. Removal Mechanism of MB

It was found that the removal of MB by α-MoO3 nanoparticles was by adsorption mechanism. In fact, the FTIR spectroscopy indicated that the removed MB cations caused by adsorption process, without chemical decomposition of MB and no intermediate compounds were detected. In addition, the increase on the effectiveness of the removal of MB using α-MoO3 nanoparticles by increasing the pH until 11 could be attributed to the basic media. From this establishment, a mechanism could be suggested (Figure 14). In fact, in the first step at pH = 11, the positive charge of the MB is maintained since its pKa is equal to 3.8 [77]. In addition, the hydroxyl groups (OH) in the solution react with α-MoO3 to produce the ion molybdate (MoO42−) without intermediate compounds [95]. Thus, the adsorption is governed by strong electrostatic interactions between the negatively surface charge of molybdate (MoO42−) and the positively charged MB cations.
The specific surface area of α-MoO3 deduced from the monolayer capacity (qm) at natural pH and has been calculated from the following equation:
Specific Surface Area (SSA) = qm × N × A
where qm is the monolayers capacity in moles per gram; N is Avogadro number (6.019 × 1023) and A is area per molecule on the surface.
The value of (57 m2/g) was slightly higher than the value deduced from the BET equation (42 m2/g), using the N2 adsorption isotherm. The difference between these values was related to the mechanism of adsorption related to nitrogen and MB molecules [96]. In the N2 absorption method, the molecules are attracted to the surface by van der Waals forces (physisorption) and multiple layers may form. However, in the case of MB used as probe molecule, there is a high bonding energy (ionic Coulombian attraction—chemisorption) and it is generally limited to a monolayer [97].

4. Conclusions

Nanocrystalline α-MoO3, synthesized through a simple method, was tested as a Nanosorbent for the removal of cationic Methylene blue dye from aqueous solution. The material exhibited higher removal efficiency (99%) at pH = 11 and a maximum removal capacity of 152 mg/g. The adsorbent was easily regenerated by calcination and the removal efficiency was 99% after three regeneration/removal cycles. Considering the easy and low-cost of α-MoO3 synthesis process, the high removal efficiency and its regeneration after several cycles, the synthesized α-MoO3 adsorbent will be proposed as promising candidate for the removal of MB from aqueous solutions.

Author Contributions

Conceptualization, S.R., H.O.H.; Methodology, S.R., H.O.H., F.K., and M.A.; Validation, S.R., H.O.H., F.K., A.M., A.A., and M.A.; Formal Analysis, H.O.H., M.A., F.K., A.A., A.M., and F.A.W.; Investigation, S.R., H.O.H., F.K., M.A., A.M., and F.A.W.; Resources, M.A., F.A.W., and A.A.; Data Curation, S.R., H.O.H., M.A., A.M., and F.A.W.; Writing-Original Draft Preparation, S.R., H.O.H., F.K., A.A., M.A., A.M., and F.A.W.; Writing-Review & Editing, H.O.H., S.R., and F.K.; Visualization, S.R., H.O.H., A.M., M.A., F.K., A.A. and F.A.W.; Supervision, S.R., H.O.H.; Project Administration, S.R., and H.O.H.; Funding Acquisition, M.A., F.A.W., and A.A.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, L.; Zhang, J.; Wang, A. Removal of methylene blue from aqueous solution using chitosan-g-poly (acrylic acid)/montmorillonite super adsorbent nanocomposite. Colloids Surf. A 2008, 32, 47–53. [Google Scholar] [CrossRef]
  2. Mohammed, M.A.; Shitu, A.; Ibrahim, A. Removal of Methylene Blue Using Low Cost Adsorbent: A Review. Res. J. Chem. Sci. 2014, 4, 91–102. [Google Scholar]
  3. Ulson de Souza, S.M.A.G.; Forgiarini, E.; Ulson de Souza, A.A. Toxicity of textile dyes and their degradation by the enzyme horseradish peroxidase (HRP). J. Hazard Mater. 2007, 147, 1073–1078. [Google Scholar] [CrossRef] [PubMed]
  4. Sucharita, A. Textile Dyes: Its Impact on Environment and its Treatment. J. Bioremed. Biodeg. 2014, 5, 1. [Google Scholar]
  5. Madrakian, T.; Afkhami, A.; Ahmadi, M.; Bagheri, H. Removal of some cationic dyes from aqueous solutions using magnetic modified multi-walled carbon nanotubes. J. Hazard Mater. 2011, 196, 109–114. [Google Scholar] [CrossRef] [PubMed]
  6. Yang, N.; Zhu, S.; Zhang, D.; Xu, S. Synthesis and properties of magnetic Fe3O4-activated carbon nanocomposite particles for dye removal. Mater. Lett. 2008, 62, 645–647. [Google Scholar] [CrossRef]
  7. Elemen, S.; Kumbasar, E.P.A.; Yapar, S. Modeling the adsorption of textile dye on organoclay using an artificial neural network. Dyes Pigment. 2012, 95, 102–111. [Google Scholar] [CrossRef]
  8. Solis, M.; Solis, A.; Perez, H.I.; Manjarrez, N.; Flores, M. Microbial decolouration of azo dyes: A review. Process Biochem. 2012, 47, 1723–1748. [Google Scholar] [CrossRef]
  9. Turgay, O.; Ersoz, G.; Atalay, S.; Forss, J.; Welander, U. The treatment of azo dyes found in textile industry wastewater by anaerobic biological method and chemical oxidation. Sep. Purif. Techol. 2011, 79, 26–33. [Google Scholar] [CrossRef]
  10. Verma, A.K.; Dash, R.R.; Bhunia, P. A review on chemical coagulation/flocculation technologies for removal of colour from textile wastewaters. J. Environ. Manag. 2012, 93, 154–168. [Google Scholar] [CrossRef] [PubMed]
  11. Greluk, M.; Hubicki, Z. Effect of basicity of anion exchangers and number and positions of sulfonic groups of acid dyes on dyes adsorption on macroporous anion exchangers with styrenic polymer matrix. Chem. Eng. J. 2013, 215–216, 731–739. [Google Scholar] [CrossRef]
  12. Kanagaraj, J.; Senthilvelan, T.; Panda, R.C. Degradation of azo dyes by laccase: Biological method to reduce pollution load in dye wastewater. Clean Technol. Environ. Policy 2015, 17, 1443–1456. [Google Scholar] [CrossRef]
  13. Vanhulle, S.; Trovaslet, M.; Enaud, E.; Lucas, M.; Taghavi, S.; van der Lelie, D.; van Aken, B.; Foret, M.; Onderwater, R.C.A.; Wesenberg, D.; et al. Decolorization, cytotoxicity and genotoxicity reduction during a combined ozonation/fungal treatment of dye-contaminated wastewater. Environ. Sci. Technol. 2008, 42, 584–589. [Google Scholar] [CrossRef] [PubMed]
  14. Cornelia, P.; Oana, P.; Robert, I.; Simona, G.M. Effective removal of methylene blue from aqueous solution using a new magnetic iron oxide nanosorbent prepared by combustion synthesis. Clean Technol. Environ. Policy 2016, 18, 705–715. [Google Scholar]
  15. Forgacs, E.; Cserhati, T.; Oros, G. Removal of synthetic dyes from wastewaters: A review. Environ. Int. 2004, 30, 953–971. [Google Scholar] [CrossRef] [PubMed]
  16. Miyah, Y.; Lahrichi, A.; Idrissi, M.; Khalil, A.; Zerrouq, F. Adsorption of methylene blue dye from aqueous solutions onto walnut shells powder: Equilibrium and kinetic studies. Surf. Interface 2018, 11, 74–81. [Google Scholar] [CrossRef]
  17. Kang, S.; Zhao, Y.; Wang, W.; Zhang, T.; Chen, T.; Yi, H.; Rao, F.; Song, S. Removal of methylene blue from water with montmorillonite nanosheets/chitosan hydrogels as adsorbent. Appl. Surf. Sci. 2018, 448, 203–211. [Google Scholar] [CrossRef]
  18. Chen, Y.H. Synthesis, characterization and dye adsorption of ilmenite nanoparticles. J. Non-Cryst. Solids 2011, 357, 136–139. [Google Scholar] [CrossRef]
  19. Ozdemir, U.; Ozbay, I.; Ozbay, B.; Veli, S. Application of economical models for dye removal from aqueous solutions: Cash flow, cost–benefit and alternative selection methods. Clean Technol. Environ. Policy 2014, 16, 423–429. [Google Scholar] [CrossRef]
  20. Sadhukhan, B.; Mondal, N.K.; Chattoraj, S. Biosorptive removal of cationic dye from aqueous system: A response surface methodological approach. Clean Technol. Environ. Policy 2014, 16, 1015–1025. [Google Scholar] [CrossRef]
  21. George, Z.; Kyzas, J.F.; Kostas, A.M. The Change from Past to Future for Adsorbent Materials in Treatment of Dyeing Wastewaters. Materials 2013, 6, 5131–5158. [Google Scholar] [Green Version]
  22. Oudghiri-Hassani, H.; Rakass, S.; Abboudi, M.; Mohmoud, A.; Al Wadaani, F. Preparation and Characterization of α-Zinc Molybdate Catalyst: Efficient Sorbent for Methylene Blue and Reduction of 3-Nitrophenol. Molecules 2018, 23, 1462. [Google Scholar] [CrossRef] [PubMed]
  23. Qian, W.C.; Luo, X.P.; Wang, X.; Guo, M.; Li, B. Removal of methylene blue from aqueous solution by modified bamboo hydrochar. Ecotoxicol. Environ. Saf. 2018, 157, 300–306. [Google Scholar] [CrossRef] [PubMed]
  24. Low, S.K.; Tan, M.C. Dye adsorption characteristic of ultrasound pre-treated pomelo peel. J. Environ. Chem. Eng. 2018, 6, 3502–3509. [Google Scholar] [CrossRef]
  25. Mounia, L.; Belkhiri, L.; Bollinger, J.C.; Bouzaza, A.; Assadi, A.; Tirri, A.; Dahmoune, F.; Madani, K.; Remini, H. Removal of Methylene Blue from aqueous solutions by adsorption on Kaolin: Kinetic and equilibrium studies. Appl. Clay Sci. 2018, 153, 38–45. [Google Scholar] [CrossRef]
  26. Bentahar, S.; Dbik, A.; El Khomri, M.; El Messaoudi, N.; Lacherai, A. Removal of a cationic dye from aqueous solution by natural clay. Groundw. Sustain. Dev. 2018, 6, 255–262. [Google Scholar] [CrossRef]
  27. Sadeghzadeh-Attar, A. Efficient photocatalytic degradation of methylene blue dye by SnO2 nanotubes synthesized at different calcination temperatures. Sol. Energy Mater. Sol. Cells 2018, 183, 16–24. [Google Scholar] [CrossRef]
  28. Zhang, Y.; Li, G.; Liu, J.; Wang, T.; Wang, X.; Liu, B.; Liu, Y.; Huo, Q.; Chu, Z. Synthesis of hierarchical hollow sodium titanate microspheres and their application for selective removal of organic dyes. J. Colloid Interface Sci. 2018, 528, 109–115. [Google Scholar] [CrossRef] [PubMed]
  29. Oliva, J.; Martinez, A.I.; Oliva, A.I.; Garcia, C.R.; Martinez-Luevanos, C.; Garcia-Lobato, M.; Ochoa-Valiente, M.; Berlanga, A. Flexible graphene composites for removal of methylene blue dye-contaminant from water. Appl. Surf. Sci. 2018, 436, 739–746. [Google Scholar] [CrossRef]
  30. Bayat, M.; Javanbakht, V.; Esmaili, J. Synthesis of zeolite/nickel ferrite/sodium alginate bionanocomposite via a co-precipitation technique for efficient removal of water-soluble methylene blue dye. Int. J. Biol. Macromol. 2018, 116, 607–619. [Google Scholar] [CrossRef] [PubMed]
  31. Kanakaraju, D.; Shahdad, N.R.M.; Lim, Y.C.; Pace, A. Magnetic hybrid TiO2/Alg/FeNPs triads for the efficient removal of methylene blue from water. Sustain. Chem. Pharm. 2018, 8, 50–62. [Google Scholar] [CrossRef]
  32. Crini, G. Non-conventional low-cost adsorbents for dye removal: A review. Bioresour. Technol. 2006, 97, 1061–1085. [Google Scholar] [CrossRef] [PubMed]
  33. Ghaedi, M.; Tavallali, H.; Sharifi, M.; Nasiri Kokhdan, S.; Asghari, A. Preparation of low cost activated carbon from Myrtus communis and pomegranate and their efficient application for removal of Congo red from aqueous solution. Spectrochim. Acta Part A 2012, 86, 107–114. [Google Scholar] [CrossRef] [PubMed]
  34. Taghizadeh, F.; Ghaedi, M.; Kamali, K.; Sharifpour, E.; Sahraie, R.; Purkait, M.K. Comparison of nickel and/or zinc selenide nanoparticle loaded on activated carbon as efficient adsorbents for kinetic and equilibrium study of removal of Arsenazo (ΙΙΙ) dye. Powder Technol. 2013, 245, 217–226. [Google Scholar] [CrossRef]
  35. Sweet, M.J.; Chessher, A.; Singleton, I. Review: Metal-based nanoparticles; size, function and areas for advancement in applied microbiology. Adv. Appl. Microbiol. 2012, 80, 113–142. [Google Scholar] [PubMed]
  36. Tan, K.B.; Vakili, M.; Horri, B.A.; Poh, P.E.; Abdullah, A.Z.; Salamatinia, B. Adsorption of dyes by nanomaterials: Recent developments and adsorption mechanisms. Sep. Purif. Technol. 2015, 150, 229–242. [Google Scholar] [CrossRef]
  37. Ai, L.; Zhang, C.; Liao, F.; Wang, Y.; Li, M.; Meng, L.; Jiang, J. Removal of methylene blue from aqueous solution with magnetite loaded multi-wall carbon nanotube: Kinetic, isotherm and mechanism analysis. J. Hazard. Mater. 2011, 198, 282–290. [Google Scholar] [CrossRef] [PubMed]
  38. Patil, M.R.; Shrivastava, V.S. Adsorption removal of carcinogenic acid violet 19 dye from aqueous solution by polyaniline-Fe2O3 magnetic nano-composite. J. Mater. Environ. Sci. 2015, 6, 11–21. [Google Scholar]
  39. Vîrlan, C.; Ciocârlan, R.G.; Roman, T.; Gherca, D.; Cornei, N.; Pui, A. Studies on adsorption capacity of cationic dyes on several magnetic nanoparticles. Acta Chem. Iasi. 2013, 21, 19–30. [Google Scholar] [CrossRef]
  40. Chang, P.R.; Zheng, P.; Liu, B.; Anderson, D.P.; Yu, J.; Ma, X. Characterization of magnetic soluble starch-functionalized carbon nanotubes and its application for the adsorption of the dyes. J. Hazard. Mater. 2011, 186, 2144–2150. [Google Scholar] [CrossRef] [PubMed]
  41. Lee, C.K.; Liu, S.S.; Juang, L.C.; Wang, C.C.; Lyu, M.D.; Hung, S.H. Application of titanate nanotubes for dyes adsorptive removal from aqueous solution. J. Hazard. Mater. 2007, 148, 756–760. [Google Scholar] [CrossRef] [PubMed]
  42. Salehi, R.; Arami, M.; Mahmoodi, N.M.; Bahrami, H.; Khorramfar, S. Novel biocompatible composite (Chitosan–zinc oxide nanoparticle): Preparation, characterization and dye adsorption properties. Colloids Surf. B 2010, 80, 86–93. [Google Scholar] [CrossRef] [PubMed]
  43. Li, X.; Xiao, W.; He, G.; Zheng, W.; Yu, N.; Tan, M. Pore size and surface area control of MgO nanostructures using a surfactant-templated hydrothermal process: High adsorption capability to azo dyes. Colloids Surf. A 2012, 408, 79–86. [Google Scholar] [CrossRef]
  44. Ealias, A.M.; Saravanakumar, M.P. A review on the classification, characterisation, synthesis of nanoparticles and their application. IOP Conf. Ser. Mater. Sci. Eng. 2017, 263, 032019. [Google Scholar] [Green Version]
  45. Xiao, X.; Song, H.; Lin, S.; Zhou, Y.; Zhan, X.; Hu, Z.; Zhang, Q.; Sun, J.; Yang, B.; Li, T.; et al. Scalable Salt-Templated Synthesis of Two-Dimensional Transition Metal Oxides. Nat. Commun. 2016, 7, 11296. [Google Scholar] [CrossRef] [PubMed]
  46. Ji, F.; Ren, X.; Zheng, X.; Liu, Y.; Pang, L.; Jiang, J.; Liu, S. 2D MoO3 Nanosheets for Superior Gas Sensors. Nanoscale 2016, 8, 8696–8703. [Google Scholar] [CrossRef] [PubMed]
  47. Oudghiri-Hassani, H. Synthesis, characterization and catalytic performance of iron molybdate Fe2(MoO4)3 nanoparticles. Catal. Commun. 2015, 60, 19–22. [Google Scholar] [CrossRef]
  48. de Castro, I.A.; Datta, R.S.; Ou, J.Z.; Castellanos-Gomez, A.; Sriram, S.; Daeneke, T.; Kalantar-zadeh, K. Molybdenum Oxides—From Fundamentals to Functionality. Adv. Mater. 2017, 29, 1701619–1701650. [Google Scholar] [CrossRef] [PubMed]
  49. Chithambararaj, A.; Sanjini, N.S.; Bose, A.C.; Velmathi, S. Flower-like Hierarchical h-MoO3: New Findings of Efficient Visible Light Driven Nano Photocatalyst for Methylene Blue Degradation. Catal. Sci. Technol. 2013, 3, 1405–1414. [Google Scholar] [CrossRef]
  50. Negishi, H.; Negishi, S.; Kuroiwa, Y.; Sato, N.; Aoyagi, S. Anisotropic Thermal Expansion of Layered MoO3 Crystals. Phys. Rev. B 2004, 69, 064111. [Google Scholar] [CrossRef]
  51. McCarron, E.M.; Calabrese, J.C. The Growth and Single Crystal Structure of a High Pressure Phase of Molybdenum Trioxide: MoO3-II. J. Solid State Chem. 1991, 91, 121–125. [Google Scholar] [CrossRef]
  52. Parise, J.B.; McCarron, E.M.; Von Dreele, R.; Goldstone, J.A. Beta-MoO3 Produced From a Novel Freeze Drying Route. J. Solid State Chem. 1991, 93, 193–201. [Google Scholar] [CrossRef]
  53. Kim, H.S.; Cook, J.B.; Lin, H.; Ko, J.S.; Tolbert, S.H.; Ozolins, V.; Dunn, B. Oxygen Vacancies Enhance Pseudocapacitive Charge Storage Properties of MoO3−x. Nat. Mater. 2016, 16, 454–460. [Google Scholar] [CrossRef] [PubMed]
  54. Yin, H.; Kuwahara, Y.; Mori, K.; Cheng, H.; Wen, M.; Yamashita, H. High-Surface-Area Plasmonic MoO3−x: Rational Synthesis and Enhanced Ammonia Borane Dehydrogenation Activity. J. Mater. Chem. A 2017, 5, 8946–8953. [Google Scholar] [CrossRef]
  55. Alsaif, M.M.Y.A.; Chrimes, A.F.; Daeneke, T.; Balendhran, S.; Bellisario, D.O.; Son, Y.; Field, M.R.; Zhang, W.; Nili, H.; Nguyen, E.P.; et al. High-Performance Field Effect Transistors Using Electronic Inks of 2D Molybdenum Oxide Nanoflakes. Adv. Funct. Mater. 2016, 26, 91–100. [Google Scholar] [CrossRef]
  56. Truong, T.G.; Meriadec, C.; Fabre, B.; Bergamini, J.F.; de Sagazan, O.; Ababou-Girard, S.; Loget, G. Spontaneous Decoration of Silicon Surfaces with MoOx Nanoparticles for the Sunlight-Assisted Hydrogen Evolution Reaction. Nanoscale 2017, 9, 1799–1804. [Google Scholar] [CrossRef] [PubMed]
  57. Fernandes, C.I.; Capelli, S.C.; Vaz, P.D.; Nunes, C.D. Highly Selective and Recyclable MoO3 Nanoparticles in Epoxidation Catalysis. Appl. Catal. A 2015, 504, 344–350. [Google Scholar] [CrossRef]
  58. Alsaif, M.M.Y.A.; Field, M.R.; Daeneke, T.; Chrimes, A.F.; Zhang, W.; Carey, B.J.; Berean, K.J.; Walia, S.; van Embden, J.; Zhang, B.; et al. Exfoliation Solvent Dependent Plasmon Resonances in Two-Dimensional Sub-Stoichiometric Molybdenum Oxide Nanoflakes. ACS Appl. Mater. Interfaces 2016, 8, 3482–3493. [Google Scholar] [CrossRef] [PubMed]
  59. Zhang, H.; Gao, L.; Gong, Y. Exfoliated MoO3 Nanosheets for High-Capacity Lithium Storage. Electrochem. Commun. 2015, 52, 67–70. [Google Scholar] [CrossRef]
  60. Song, G.; Hao, J.; Liang, C.; Liu, T.; Gao, M.; Cheng, L.; Hu, J.; Liu, Z. Degradable Molybdenum Oxide Nanosheets with Rapid Clearance and Efficient Tumor Homing Capabilities as a Therapeutic Nanoplatform. Angew. Chem. Int. Ed. 2016, 55, 2122–2126. [Google Scholar] [CrossRef] [PubMed]
  61. Abboudi, M.; Oudghiri-Hassani, H.; Wadaani, F.; Rakass, S.; Al Ghamdi, A.; Messali, M. Enhanced catalytic reduction of para-nitrophenol using α-MoO3 molybdenum oxide nanorods and stacked nanoplates as catalysts prepared from different precursors. J. Taibah Univ. Sci. 2018, 12, 133–137. [Google Scholar] [CrossRef]
  62. Wang, Y.; Zhang, X.; Luo, Z.; Huang, X.; Tan, C.; Li, H.; Zheng, B.; Li, B.; Huang, Y.; Yang, J.; et al. Liquid Phase Growth of Platinum Nanoparticles on Molybdenum Trioxide Nanosheets: An Enhanced Catalyst with Intrinsic Peroxidase-like Catalytic Activity. Nanoscale 2014, 6, 12340–12344. [Google Scholar] [CrossRef] [PubMed]
  63. Bai, H.; Yi, W.; Li, J.; Xi, G.; Li, Y.; Yang, H.; Liu, J. Direct Growth of Defect-Rich MoO3−X Ultrathin Nanobelts for Efficiently Catalyzed Conversion of Isopropyl Alcohol to Propylene under Visible Light. J. Mater. Chem. A 2016, 4, 1566–1571. [Google Scholar] [CrossRef]
  64. Santos-Beltran, M.; Paraguay-Delgado, F.; Garcıa, R.; Antunez-Flores, W.; Ornelas-Gutierrez, C.; Santos-Beltran, A. Fast methylene blue removal by MoO3 nanoparticles. J. Mater. Sci. Mater. Electron. 2016, 28, 2935–2948. [Google Scholar] [CrossRef]
  65. Tiwari, D.K.; Behari, J.; Sen, P. Application of Nanoparticles in Waste Water Treatment. World Appl. Sci. J. 2008, 3, 417–433. [Google Scholar]
  66. Abboudi, M.; Oudghiri-Hassani, H.; Wadaani, F.; Messali, M.; Rakass, S. Synthesis Method of Precursors to produce Molybdenum Oxide MoO3 and related Materials. U.S. Patent 9,611,152B2, 4 April 2007. [Google Scholar]
  67. Benkhaya, S.; El Harfi, S.; El Harfi, A. Classifications, properties and applications of textile dyes: A review. Appl. J. Envir. Eng. Sci. 2017, 3, 311–320. [Google Scholar]
  68. Rakass, S.; Mohmoud, A.; Oudghiri-Hassani, H.; Abboudi, M.; Kooli, F.; Wadaani, F. Modified Nigella Sativa Seeds as a Novel Efficient Natural Adsorbent for Removal of Methylene Blue Dye. Molecules 2018, 23, 1950. [Google Scholar] [CrossRef] [PubMed]
  69. Mahmoud, D.K.; Salleh, M.A.M.; Karim, W.A.W.A.; Idris, A.; Abidin, Z.Z. Batch adsorption of basic dye using acid treated kenaf fibre char: Equilibrium, kinetic and thermodynamic studies. Chem. Eng. J. 2012, 181–182, 449–457. [Google Scholar] [CrossRef]
  70. Kannan, N.; Karuppasamy, K. Low cost adsorbents for the removal of phenyl aceticacid from aqueous solution. Indian J. Environ. Protec. 1998, 18, 683–690. [Google Scholar]
  71. Wawrzkiewicz, M.; Hubicki, Z. Removal of Tartrazine from aqueous solutions by strongly basic polystyrene anion exchange resins. J. Hazard. Mater. 2009, 164, 502–509. [Google Scholar] [CrossRef] [PubMed]
  72. Krishnan, K.; Anirudhan, T.S. A Preliminary examination of the adsorption characteristics of Pb(II) ions using sulphurised activated carbon prepared from bagasse pith. Indian J. Chem. Technol. 2002, 9, 32–40. [Google Scholar]
  73. Karaer, H.; Kaya, I. Synthesis, characterization of magnetic chitosan/active charcoal composite and using at the adsorption of methylene blue and reactive blue4. Micropor. Mesopor. Mat. 2016, 232, 26–38. [Google Scholar] [CrossRef]
  74. Özcan, A.; Öncü, E.M.; Özcan, A.S. Kinetics, isotherm and thermodynamic studies of adsorption of Acid Blue 193 from aqueous solutions onto natural sepiolite. Colloids Surf. A Physicochem. Eng. Asp. 2006, 277, 90–97. [Google Scholar] [CrossRef]
  75. Patil, S.; Renukdas, S.; Patel, N. Removal of methylene blue, a basic dye from aqueous solutions by adsorption using teak tree (Tectona grandis) bark powder. Inter. J. Environ. Sci. 2011, 1, 711–726. [Google Scholar]
  76. 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]
  77. Jihyun, R.K.; Santiano, B.; Kim, H.; Kan, E. Heterogeneous Oxidation of Methylene Blue with Surface-Modified Iron-Amended Activated Carbon. Am. J. Anal. Chem. 2013, 4, 115–122. [Google Scholar]
  78. Febrianto, J.; Kosasih, A.N.; Sunarso, J.; Ju, Y.; Indraswati, N.; Ismadji, S. Equilibrium and kinetic studies in adsorption of heavy metals using biosorbent: A summary of recent studies. J. Hazard. Mater. 2009, 162, 616–645. [Google Scholar] [CrossRef] [PubMed]
  79. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  80. Furusawa, T.; Smith, J.M. Intraparticle mass transport in slurries by dynamic adsorption studies. AIChE J. 1974, 20, 88–93. [Google Scholar] [CrossRef]
  81. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  82. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef]
  83. Shahwan, T.; Erten, H.N. Temperature effects in barium sorption on natural kaolinite and chlorite-illite clays. J. Radioanal. Nucl. Chem. 2004, 260, 43–48. [Google Scholar] [CrossRef] [Green Version]
  84. Dada, A.O.; Olalekan, A.P.; Olatunya, A.M.; Dada, O. Langmuir, Freundlich, Temkin and Dubinin–Radushkevich Isotherms Studies of Equilibrium Sorption of Zn2+ Unto Phosphoric Acid Modified Rice Husk. J. Appl. Chem. 2012, 3, 38–45. [Google Scholar]
  85. Ma, J.; Yu, F.; Zhou, L.; Jin, L.; Yang, M.X.; Luan, J.S.; Tang, Y.H.; Fan, H.B.; Yuan, Z.W.; Chen, J.H. Enhanced adsorptive removal of methyl orange and methylene blue from aqueous solution by alkali-activated multiwalled carbon nanotubes. ACS Appl. Mater. Interfaces 2012, 4, 5749–5760. [Google Scholar] [CrossRef] [PubMed]
  86. Zhang, Y.R.; Wang, S.Q.; Shen, S.L.; Zhao, B.X. A novel water treatment magnetic nanomaterial for removal of anionic and cationic dyes under severe condition. Chem. Eng. J. 2013, 233, 258–264. [Google Scholar] [CrossRef]
  87. Xiong, L.; Yang, Y.; Mai, J.X.; Sun, W.L.; Zhang, C.Y.; Wei, D.P.; Chen, Q.; Ni, J.R. Adsorption behavior of methylene blue onto titanate nanotubes. Chem. Eng. J. 2010, 156, 313–320. [Google Scholar] [CrossRef]
  88. Roosta, M.; Ghaedi, M.; Daneshfar, A.; Sahraei, R.; Asghari, A. Optimization of the ultrasonic assisted removal of methylene blue by gold nanoparticles loaded on activated carbon using experimental design methodology. Ultrason. Sonochem. 2014, 21, 242–252. [Google Scholar] [CrossRef] [PubMed]
  89. Ghaedi, M.; Heidarpour, S.; Nasiri Kokhdan, S.; Sahraie, R.; Daneshfar, A.; Brazesh, B. Comparison of silver and palladium nanoparticles loaded on activated carbon for efficient removal of Methylene blue: Kinetic and isotherm study of removal process. Powder Technol. 2012, 228, 18–25. [Google Scholar] [CrossRef]
  90. Xie, Y.; Qian, D.; Wu, D.; Ma, X.F. Magnetic halloysite nanotubes/iron oxide composites for the adsorption of dyes. Chem. Eng. J. 2011, 168, 959–963. [Google Scholar] [CrossRef]
  91. Chen, A.S.C.; Sorg, T.J.; Wang, L. Regeneration of iron-based adsorptive media used for removing arsenic from groundwater. Water Res. 2015, 77, 85–97. [Google Scholar] [CrossRef] [PubMed]
  92. Seguin, L.; Figlarz, M.; Cavagnat, R.; Lassègues, J.C. Infrared and Raman spectra of MoO3 molybdenum trioxides and MoO3·xH2O molybdenum trioxide hydrates. Spectrochim. Acta A Mol. Biomol. Spect. 1995, 51, 1323–1344. [Google Scholar] [CrossRef]
  93. Ahmed, F.; Dewani, R.; Pervez, M.K.; Mahboob, S.J.; Soomro, S.A. Non-destructive FT-IR analysis of mono azo dyes. Bulg. Chem. Commun. 2016, 48, 71–77. [Google Scholar]
  94. Etman, A.S.; Abdelhamid, H.N.; Yuan, Y.; Wang, L.; Zou, X.; Sun, J. Facile Water-Based Strategy for Synthesizing MoO3−x Nanosheets: Efficient Visible Light Photocatalysts for Dye Degradation. ACS Omega 2018, 3, 2201–2209. [Google Scholar] [CrossRef]
  95. Aracena, A.; Sannino, A.; Jerez, O. Dissolution kinetics of molybdite in KOH media at different temperatures. Trans. Nonferrous Met. Soc. China 2018, 28, 177–185. [Google Scholar] [CrossRef]
  96. Li, F.; Wu, X.; Ma, S.; Xu, Z.; Liu, W.; Liu, F. Adsorption and Desorption Mechanisms of Methylene Blue Removal with Iron-Oxide Coated Porous Ceramic Filter. J. Water Resour. Prot. 2009, 1, 35–40. [Google Scholar] [CrossRef]
  97. Santamarina, J.C.; Klein, K.A.; Wang, Y.H.; Prencke, E. Specific surface: Determination and relevance. Can. Geotech. J. 2002, 39, 233–241. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compounds molybdenum trioxide (α-MoO3) are available from the authors.
Figure 1. Effect of initial dye concentration and contact time on removal efficiency of methylene blue (MB) using α-MoO3 (madsorbent = 0.1 g, T = 25 °C, pH = 5.5).
Figure 1. Effect of initial dye concentration and contact time on removal efficiency of methylene blue (MB) using α-MoO3 (madsorbent = 0.1 g, T = 25 °C, pH = 5.5).
Molecules 23 02295 g001
Figure 2. Adsorbent dose effect and initial dye concentration for the efficiency of MB removal using α-MoO3 for 30 min, T = 25 °C, pH = 5.5.
Figure 2. Adsorbent dose effect and initial dye concentration for the efficiency of MB removal using α-MoO3 for 30 min, T = 25 °C, pH = 5.5.
Molecules 23 02295 g002
Figure 3. Effect of temperature on the removal efficiency of 40 ppm of MB solution using α-MoO3 (t = 30 min, pH = 5.5).
Figure 3. Effect of temperature on the removal efficiency of 40 ppm of MB solution using α-MoO3 (t = 30 min, pH = 5.5).
Molecules 23 02295 g003
Figure 4. Von’t Hoff plot showing the temperature effect for the removal of MB by α-MoO3.
Figure 4. Von’t Hoff plot showing the temperature effect for the removal of MB by α-MoO3.
Molecules 23 02295 g004
Figure 5. Effect of pH on the removal efficiency of 40 ppm of MB solution using α-MoO3 (mads = 0.1 g, T = 25 °C, t = 30 min).
Figure 5. Effect of pH on the removal efficiency of 40 ppm of MB solution using α-MoO3 (mads = 0.1 g, T = 25 °C, t = 30 min).
Molecules 23 02295 g005
Figure 6. Effect of initial dye concentration contact time on the removal efficiency of MB using α-MoO3 at pH 11 (mads = 0.1 g, T = 25 °C).
Figure 6. Effect of initial dye concentration contact time on the removal efficiency of MB using α-MoO3 at pH 11 (mads = 0.1 g, T = 25 °C).
Molecules 23 02295 g006
Figure 7. Pseudo first-order model plot showing the effect of contact time and initial dye concentration of MB removal by α-MoO3.
Figure 7. Pseudo first-order model plot showing the effect of contact time and initial dye concentration of MB removal by α-MoO3.
Molecules 23 02295 g007
Figure 8. Pseudo Second order model plot showing the effect of contact time and initial dye concentration of MB removal by α-MoO3.
Figure 8. Pseudo Second order model plot showing the effect of contact time and initial dye concentration of MB removal by α-MoO3.
Molecules 23 02295 g008
Figure 9. Intra-particle diffusion model plot showing the effect of contact time and initial dye concentration of MB removal by α-MoO3.
Figure 9. Intra-particle diffusion model plot showing the effect of contact time and initial dye concentration of MB removal by α-MoO3.
Molecules 23 02295 g009
Figure 10. Freundlich (a) and Langmuir (b) isotherm model plots showing the effect of initial dye concentration for the removal of MB by α-MoO3.
Figure 10. Freundlich (a) and Langmuir (b) isotherm model plots showing the effect of initial dye concentration for the removal of MB by α-MoO3.
Molecules 23 02295 g010
Figure 11. Recycled efficiency of α-MoO3 for removal of Methylene blue.
Figure 11. Recycled efficiency of α-MoO3 for removal of Methylene blue.
Molecules 23 02295 g011
Figure 12. Fourier transform infrared (FTIR) spectra of MoO3, MoO3-MB1, MoO3-R, MoO3-MB2 and MB.
Figure 12. Fourier transform infrared (FTIR) spectra of MoO3, MoO3-MB1, MoO3-R, MoO3-MB2 and MB.
Molecules 23 02295 g012
Figure 13. Scanning electron microscopy (SEM) Micrographs of the starting (A) Molybdenium trioxide (α-MoO3) (magnification of ×5000, scale bar of 10 μm), (B) after MB dye removed (magnification of ×5000, scale bar of 10 μm), (C) relates to the regenerated α-MoO3 (magnification of ×60,000, scale bar of 1 μm) and (D) after first regeneration/removal cycle of MB dye (magnification of ×60,000, scale bar of 1 μm), (E) shows the morphology of α-MoO3 after second regeneration process (magnification of ×60,000, scale bar of 1 μm).
Figure 13. Scanning electron microscopy (SEM) Micrographs of the starting (A) Molybdenium trioxide (α-MoO3) (magnification of ×5000, scale bar of 10 μm), (B) after MB dye removed (magnification of ×5000, scale bar of 10 μm), (C) relates to the regenerated α-MoO3 (magnification of ×60,000, scale bar of 1 μm) and (D) after first regeneration/removal cycle of MB dye (magnification of ×60,000, scale bar of 1 μm), (E) shows the morphology of α-MoO3 after second regeneration process (magnification of ×60,000, scale bar of 1 μm).
Molecules 23 02295 g013
Figure 14. Schematic mechanism of the MB removal using the Molybdenum trioxide nanosorbent.
Figure 14. Schematic mechanism of the MB removal using the Molybdenum trioxide nanosorbent.
Molecules 23 02295 g014
Table 1. Thermodynamic parameters for removal of MB by α-MoO3.
Table 1. Thermodynamic parameters for removal of MB by α-MoO3.
AdsorbentAdsorbate∆H° (KJ·mol−1)∆S° (KJ·mol−1·K)∆G° (KJ·mol−1)
α-MoO3MB900.316298 K323 K343 K
−3.741−11.643−12.305
Table 2. Kinetic models’ equations.
Table 2. Kinetic models’ equations.
ModelEquationParameters
Pseudo first-order (PFD) [79] Ln ( q e q t ) = Lnq e + K 1 t qt: the removal capacity at time t (mg/g); qe: the removal capacity at equilibrium (mg/g); K1: the rate constant of pseudo first-order adsorption (1/min)
Pseudo second-order (PSD) [79] t q t = 1 K 2 q e 2 + t q e qt: the removal capacity at time t (mg/g); qe: the removal capacity at equilibrium (mg/g); K2: the pseudo second-order rate constant (g·mg−1·min−1)
Intraparticle diffusion (IPD) [80]. q t = K I t 0.5 + l I (mg/g) and KI (mg/(g·min0.5)) are the intraparticle diffusion constants, qt: the removal capacity (mg/g) at time t; t: the contact time (min)
Table 3. Kinetic parameters for removal of MB using α-MoO3.
Table 3. Kinetic parameters for removal of MB using α-MoO3.
Dye Ci mg/LPseudo First-OrderPseudo Second-OrderIntra-Particle-Diffusion Model
qexp (mg/g)qe (mg/g)k1 (1/min)R12qe (mg/g)k2 (g/mg min)R22I (mg/g)ki (mg/g min0.5)R32
10099.82810.0970.9971110.000970.99860.204.420.832
130129.53210.0970.9961360.001470.99993.154.030.834
150149.62250.0450.9952000.00017134.3511.750.910
Where qexp is the removal capacity (mg/g) at 120 min.
Table 4. Adsorption Isotherm model equations for removal of MB using α-MoO3.
Table 4. Adsorption Isotherm model equations for removal of MB using α-MoO3.
ModelEquationParameters
Freundlich [81] Lnq e = Lnq F + 1 n LnC e n: the heterogeneity factor (g/L); qF: the Freundlich constant (mg(1−1/n)·L1/n·g−1); Ce: concentration of MB at equilibrium (ppm); qe: the MB dye amount adsorbed by α-MoO3 at equilibrium (mg/g)
Langmuir [82] C e q e = 1 q m K L + C e q m Ce: concentration of MB at equilibrium (ppm); qe: the MB dye amount adsorbed by α-MoO3 at equilibrium (mg/g); KL: Langmuir constant of adsorption (L/mg); qm: the maximum amount of MB dye removed by α-MoO3 (mg/g)
R L = 1 1 + K L C i KL: the Langmuir constant; Ci: the initial concentration of MB; RL: values indicate that the removal of MB could be linear (RL = 1), irreversible (RL = 0), favorable (0 < RL < 1), or unfavorable (RL > 1).
Dubinin–Radushkevich (D-R) [83] Lnq e = Lnq m K ε 2 ε: the Polanyi potential; K: constant for the sorption energy (mol2/kJ2); R: the Universal gas constant (8.314 J.mol-1 K−1); T : the temperature (K); Ce: the equilibrium concentration of the MB dye left in the solution (ppm); qm: the theoretical saturation capacity.
ε = R T L n ( 1 + 1 C e )
Temkin [84] q e = B T LnA T + B T LnC e BT = RT/bT; bT: the Temkin constant related to heat of sorption (J/mol); AT: the Temkin isotherm constant (L/g); R: the gas constant (8.314 J/mol K); T: the absolute temperature (K)
Table 5. Isotherm parameters for removal of MB using α-MoO3.
Table 5. Isotherm parameters for removal of MB using α-MoO3.
LangmuirFreundlichTemkinDubinin–Radushkevich
qm (mg/g)KL (L/mg)R2Range RLqF (mg(1−1/n)·L1/n·g−1)1/nR2AT (L/g)BT (J/mol)R2qm (mg/g)R2E (KJ/mol)
1529.5810.0007–0.00901610.3010.99774.836.560.9891520.93916
Table 6. Earlier reports for the highest amount of MB removed (qm).
Table 6. Earlier reports for the highest amount of MB removed (qm).
NanosorbentQmax (mg/g)Reference
Magnetic iron oxide nanosorbent25.54[14]
Alkali-activated multiwalled carbon nanotubes399.00[85]
Fe3O4 magnetic nanoparticles modified with 3-glycidoxypropyltrimethoxysilane and glycine158.00[86]
Calcined titanate nanotubes133.33[87]
Gold nanoparticles loaded on activated carbon104.00–185.00[88]
Silver nanoparticles loaded on activated carbon71.43[89]
Palladium nanoparticles loaded on activated carbon75.40[89]
Magnetic halloysite nanotubes/iron oxide composites18.44[90]
Zinc molybdate nanoparticles217.86[22]
Molybdenum trioxide nanoparticles (hexagonal and orthorhombic phases)122.50[64]
Molybdenum trioxide nanorods and stacked nanoplates152.00This work

Share and Cite

MDPI and ACS Style

Rakass, S.; Oudghiri Hassani, H.; Abboudi, M.; Kooli, F.; Mohmoud, A.; Aljuhani, A.; Al Wadaani, F. Molybdenum Trioxide: Efficient Nanosorbent for Removal of Methylene Blue Dye from Aqueous Solutions. Molecules 2018, 23, 2295. https://doi.org/10.3390/molecules23092295

AMA Style

Rakass S, Oudghiri Hassani H, Abboudi M, Kooli F, Mohmoud A, Aljuhani A, Al Wadaani F. Molybdenum Trioxide: Efficient Nanosorbent for Removal of Methylene Blue Dye from Aqueous Solutions. Molecules. 2018; 23(9):2295. https://doi.org/10.3390/molecules23092295

Chicago/Turabian Style

Rakass, Souad, Hicham Oudghiri Hassani, Mostafa Abboudi, Fethi Kooli, Ahmed Mohmoud, Ateyatallah Aljuhani, and Fahd Al Wadaani. 2018. "Molybdenum Trioxide: Efficient Nanosorbent for Removal of Methylene Blue Dye from Aqueous Solutions" Molecules 23, no. 9: 2295. https://doi.org/10.3390/molecules23092295

Article Metrics

Back to TopTop