Next Article in Journal
Comparative Analysis of CO2 Adsorption Performance of Bamboo and Orange Peel Biochars
Previous Article in Journal
Completing the Spectral Mosaic of Chloromethane by Adding the CHD2Cl Missing Piece Through the Interplay of Rotational/Vibrational Spectroscopy and Quantum Chemical Calculations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Magnesium Molybdate: An Efficient Nanosorbent for Methylene Blue Cationic Dye Removal from Aqueous Solutions

1
Chemistry Department, Faculty of Science, Taibah University, Madinah 30002, Saudi Arabia
2
Laboratory of Applied Organic Chemistry (LCOA), Chemistry Department, Faculty of Sciences and Techniques, Sidi Mohamed Ben Abdellah University, Imouzzer Road, P.O. Box 2202, Fez 30000, Morocco
3
Engineering Laboratory of Organometallic, Molecular Materials, and Environment (LIMOME), Chemistry Department, Faculty of Sciences, Sidi Mohamed Ben Abdellah University, P.O. Box 1796, Fez 30000, Morocco
4
Chemistry Department, Faculty of Science, Islamic University of Madinah, Madinah 42351, Saudi Arabia
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(7), 1606; https://doi.org/10.3390/molecules30071606
Submission received: 23 February 2025 / Revised: 23 March 2025 / Accepted: 28 March 2025 / Published: 3 April 2025
(This article belongs to the Section Analytical Chemistry)

Abstract

:
The removal of methylene blue (MB) cationic dye from aqueous solutions was investigated by applying magnesium molybdate (β-MgMoO4) as a nanosorbent. The β-MgMoO4 was synthesized through a simple, rapid, and efficient method. The MB dye removal process was optimized by evaluating various parameters such as temperature, contact time, nanosorbent dosage, pH, and initial cationic dye concentration. The optimal conditions for MB removal were found to be pH 3, with a 99% removal efficiency achieved in just 10 min of contact time, when using an MB cationic dye concentration of 160 ppm. Magnesium molybdate (β-MgMoO4) showed a maximum adsorption capacity of 356 mg/g, according to Langmuir model-based calculations. The MB dye removal process occurred spontaneously while being favorable and endothermic. The kinetic investigation showed that the pseudo-second-order model accurately represented the reaction kinetics. The thermal regeneration test results indicated that the removal efficiency remained stable even after three consecutive rounds of reuse. A Fourier Transform Infrared (FTIR) spectroscopic analysis confirmed the adsorption and desorption of MB on β-MgMoO4 and its regeneration. Overall, these results indicate that a β-MgMoO4 nanosorbent is a favorable and robust adsorbent for the removal of MB cationic dye from wastewater at its maximum capacity.

Graphical Abstract

1. Introduction

Manufactured synthetic dyes [1], which are widely used in, but are not limited to, industries such as leather tanning [2], mining [3], electroplating [4], cement [5], chemical fertilizer [6], photoelectrochemical cells [7], textiles [8], cosmetics [9], paper production [10], and coloring agents, for various products, i.e., cotton, food [11], wool, leather, and silk products [12], contain numerous toxic substances [13]. These substances include bases [14], acids [15], dissolved solids, and different colored dyes. Dyes are major sources of water pollution, and, with their complex molecular structures, are highly stable [16], resistant to washing [17], and difficult to degrade [18]. As a result, dyes are considered to be poisonous [19], and they pose a substantial threat to aquatic life, freshwater sources, and human health owing to their toxicity, carcinogenicity, and potential to cause considerable environmental damage [19]. On the other hand, the increasing demand for freshwater, primarily driven by agricultural and industrial activities, has led to global concern regarding the scarcity of this vital resource [20]. To reduce dependency on freshwater sources, using recycled water has been suggested as an alternative solution. This necessitates effective wastewater purification and dye removal methods [21].
In response to this need, various dye removal techniques have been developed, including flocculation, ion exchange, ozone oxidation, chemical coagulation, extraction, membrane separation, biological treatment, electrochemical methods, photodegradation, adsorption, and hypochlorite oxidation [22,23,24,25,26,27,28,29]. Amongst these techniques, adsorption is considered efficient, simple, and low-cost, making it a popular alternative to other expensive treatment procedures [30].
Numerous efforts have been made to develop synthetic or natural materials with high efficiency in removing and degrading the harmful dyes in industrial wastewater using biological, chemical, and physical techniques [31,32].
Activated carbon is universally used as a natural adsorbent as a result of its large surface area [24,25] and high adsorption capacity. However, its use is limited due to its production expenses, regeneration abilities, phase separation difficulties, and low-quality mechanical properties [26]. Thus, research is being conducted to come up with novel adsorbents with high adsorption and regeneration abilities.
Recently, binary metal oxides have attracted great interest because of their prospective performances in various materials [27]. Particularly, metal molybdates (MMoO4 or M2(MoO4)3) have drawn attention for their applications in catalysis [33,34], the oxidative dehydrogenation of light alkanes [35,36,37,38,39], and the partial oxidation of hydrocarbons [40], and also for environmental applications such as the photocatalytic oxidation of dyes [41,42,43,44], the oxidation of methylene blue (MB) dye [45], and the removal of phosphate [46].
Magnesium molybdate compounds have drawn great interest for their possible applications as catalysts [47,48,49,50], as phosphors [51,52,53,54,55,56,57,58], as supercapacitors [59], and in lithium-ion batteries [60].
Magnesium molybdate belongs to the wolframite-type phases and exists in two polymorphs: the ß form, which is stable in the temperature range of 500 °C up to 1100 °C, and the cupro scheelite α form, which is generally synthesized using high pressures of about ~5 GPa [14].
So far, different methods of MgMoO4 synthesis have been mentioned in the literature, including impregnation synthesis [61], solid-state reaction synthesis [51,52,53,54,55], mechanochemical activation synthesis [62,63], hydrothermal synthesis [56,57,58], coprecipitation synthesis [49,50], an ultrasonic spray pyrolysis method [47], sol–gel synthesis [64], and a citrate method [48]. Different molybdate materials have been explored for the removal of dyes from aqueous solutions, such as nickel molybdate, zinc molybdate, and iron molybdate [43,49,58]. However, until now, the use of ß-MgMoO4 for the removal of dyes from aqueous solutions and other environmental applications has not yet been explored.
The current study focused on using ß magnesium molybdate as a cationic dye adsorbent. Several parameters, for instance, temperature, nanosorbent dosage, initial cationic dye concentration, solution pH, and contact time, were all studied to determine their specific effects on MB cationic dye removal activity. An evaluation of adsorption isotherms and kinetics was carried out. Additionally, the removal efficiency, after regenerating the used nanosorbent by calcination at high temperatures, was also investigated.

2. Results and Discussion

2.1. MB Removal

2.1.1. pH Effect

pH is a necessary parameter in regard to managing dye removal [65]. Therefore, its effect on MB removal using a magnesium molybdate (β-MgMoO4) nanosorbent was explored in this study by changing its pH values from 3 to 11 at a stable room temperature (20 °C) with the cationic dye’s initial concentration at 100 ppm. Figure 1 shows that there is a distinctive effect of pH on MB cationic dye removal. For example, a decrease in the MB percentage from 97% to 36% was seen with increasing pH from 3 to 11. A similar trend was observed for the amount of cationic dye removed per unit mass of adsorbent at equilibrium (qe), which decreased from 97 to 36 mg/g. This finding highlights the critical role of acidic conditions.
The presence of excess H+ ions is expected to protonate the positive charge on methylene blue dye (MBH2+), as it is unlikely that these ions would protonate the negative charges on the surface of β-MgMoO4. The surface area electrostatic fields of β-MgMoO4 and MB were raised by protonation of the surface area charges by excess H+ ions. This was mirrored by the higher color removal percentage at lower pH. Thus, the maximum removal of the MB cationic dye using β-MgMoO4 nanosorbent was achieved at pH = 3. Similar conclusions were reported by Oudghiri Hassani et al. [66] in a study of an α-zinc molybdate catalyst.

2.1.2. Effect of Adsorbent Dose

Adsorbent dosage is considered the most vital parameter in adsorption methods [67,68]. MB dye removal using β-MgMoO4, with the initial dye concentration at 160 ppm, was discovered to increase upon altering the dosage of the nanosorbent from 0.0001 to 0.1 g/L. This is described in Figure 2, the percentage (%) of removal of MB cationic dye is increased, while the surface concentration (Γ) (mg/m2) decreases if the nanosorbent dosage increases from 0.0001 to 0.1 g/L. This indicates that the available surface area and the adsorption sites were not fully covered by the present MB molecules [69].

2.1.3. Initial Cationic Dye Concentration and Contact Time Effect

Experiments on the effect of adsorption contact time for MB cationic dye solution on β-MgMoO4 were conducted at pH = 3 with various controlled contact time intervals varying from 0 min to 120 min. The interaction between the MB cationic dye and the availability of active sites on the β-MgMoO4 nanosorbent influences the initial dye concentration.
The results shown in Figure 3 indicate that the removal of the MB cationic dye by magnesium molybdate attained its highest value of 99% at 10 min, at a concentration of 160 ppm. However, at a concentration greater than 160 ppm, the MB removal percentage reached/decreased to 93%, 92%, and 90% with initial concentrations of 170, 180, and 200 ppm, respectively, at 120 min of contact time.
Furthermore, the actual adsorption capacity results gained at equilibrium increased notably from 317 mg/g to 360 mg/g, with the initial cationic dye concentration rising from 160 mg/L to 200 mg/L. This phenomenon is attributed to the many empty sites on the β-MgMoO4 surface. Immediately after the sites of the β-MgMoO4 surface adsorption area were fully occupied by MB cationic dye, no more β-MgMoO4 surface area sites were available for adsorption; hence, the maximum adsorption capacity was achieved at this stage [70].
In summary, the optimal adsorption contact time for maximum MB removal using β-MgMoO4 is 10 min, and the equilibrium adsorption capacity also increases with an increasing initial cationic dye concentration up to a certain limit, after which it starts to decrease.

2.1.4. Temperature Effect

The experimental process for removing the MB cationic dye was examined at different temperatures, ranging from 20 to 60 °C. The temperature effect results, as confirmed in Figure 4, indicate that the MB removal percentage decreases from 82% to 46% with the initial cationic dye concentration of 200 ppm as the temperature increases. The adsorption capacity also decreases from 326 mg/g to 184 mg/g, demonstrating that exothermic activity controls the adsorption of MB on β-MgMoO4.
In summary, the temperature is a key parameter that significantly affects the removal of dyes [71], in this case, using β-MgMoO4 as an adsorbent. These results indicate that an exothermic process governs the adsorption of MB onto β-MgMoO4, with lower temperatures enhancing removal efficiency.
An important parameter in the adsorption processes is thermodynamic factors [72]. The possibility and the adsorption mechanism are likely to be predicted based on the thermodynamic parameters [73]. These can be calculated via the below-mentioned formulas:
Δ G o = R T L n K d
K d = C a C e
L n K d = Δ S o R Δ H o R T
Kd stands for the distribution constant, T stands for the absolute temperature (K), R stands for the gas constant (J·mol−1·K−1), ΔG° stands for the free energy, Ca stands for the amount of cationic dye adsorbed at equilibrium, Ce stands for the equilibrium concentration (mol/L), and ΔS° and ΔH° stand for entropy and enthalpy, respectively.
Kd is dimensionless: Kd = Ca/Ce in(L/g) ((mg/g)/(mg/L)). To convert it, the obtained value of Kd, expressed in (L/g), is multiplied by (the molecular weight of the methylene blue (319.85 g/mol) × (the number of moles of water per L of solution (55 moles/L).
The calculated values of ∆S° and ∆H° were obtained from the intercept, as was the slope of the plot ln Kd versus 1/T ( L n K d = 2.09 + 3904.58 T (Figure 5)). The ∆G° values were determined using Formula (1). All calculated data are presented in Table 1.
The negative value of ∆G° indicates a favorable and a spontaneous adsorption process. However, the enthalpy (−32.46 KJ·mole−1) and entropy (−0.017 KJ·mole−1·K) of the adsorption of MB onto β-MgMoO4 are negative, showing the exothermic behavior of the process and a decrease in randomness at the solid–liquid interface during the adsorption process, respectively. The electrostatic attraction between the MB cationic dye and β-MgMoO4 (at pH = 3) improves the adsorption forces. Subsequently, the fixation, association, or immobilization of cationic dye molecules on the β-MgMoO4 surface area lowers the extent of the freedom of cationic dye molecules. Indeed, based on the ΔG_ and ΔH_ values, it can be postulated that the adsorption of MB is a physical adsorption process [69].

2.2. Study of Kinetics

A kinetics test for the removal of MB cationic dye was performed as it provides a hint about the adsorption system [74].
The kinetics data obtained from removing the MB cationic dye by applying the β-MgMoO4 nanosorbent were studied with intra-particle diffusion, pseudo-first-order, and pseudo-second-order kinetic models. The equations of the considered models are given in Table 2.
The kinetic model parameters, intra-particle diffusion, pseudo-first-order and pseudo-second-order are tabulated in Table 3 and shown in Figure 6, Figure 7, and Figure 8, respectively. They vary in terms of the values of their correlation coefficients (R2) with their linear regressions. These values were assessed for the examined concentrations and found to be as follows: 0.810 to 0.884 for intra-particle, 0.749 to 0.975 for pseudo-first-order, and 0.979 to 0.998 for pseudo-second-order. As the R2 value is equal to or close to 1 for the pseudo-second-order model, it fits perfectly with the investigational data. And hence, the kinetic parameters attained from the pseudo-second-order model agree very well with the parameters from the experimental data. The kinetic studies confirm that the pseudo-second-order model accurately describes the reaction kinetics. This can be seen from the results showing that the pseudo-second-order kinetic model was able to accurately predict the rate of the reaction. This indicates that the reaction mechanism is consistent with the pseudo-second-order kinetic model.

2.3. Adsorption Isotherms

It is crucial to study such isotherms due to their perfect explanations when planning adsorption methods [77]. In this current work, four main adsorption models were examined, specifically Dubinin–Radushkevich, Temkin, Freundlich, and Langmuir. These main adsorption models are ruled by the equations given below in Table 4.
The Freundlich, Temkin, Langmuir, and Dubinin–Radushkevich models were tested to see if they perfectly fit the investigational data. The model parameters and the regression correlation coefficients (R2) are given in Table 5, as obtained from Figure 9. The maximum value that was obtained for the Langmuir model was R2 (0.991), while the Freundlich, Temkin, and D–R models’ fits showed lower values of R2 (0.343, 0.338, and 0.210, respectively). Accordingly, Langmuir isotherm fits perfectly with that of the results obtained in the experiments, proposing that the cationic dye removal proceeds via the arrangement of the MB cationic dye monolayer onto the β-MgMoO4 nanosorbent surface area, with an adsorption capability of 356 mg/g, which leads to a homogenous surface area. However, the separation factor RL, ranging from 0.0031 to 0.0038, indicates favorable dye removal by β-MgMoO4. Hence, the explored nanosorbent has excellent removal efficiency in comparison with many other adsorbent materials (Table 6).

2.4. Magnesium Molybdate (β-MgMoO4) Regeneration and Characterization

2.4.1. Regeneration Efficacy

Various regeneration procedures have been suggested in the literature, such as bio-regeneration, microwave irradiation, thermal treatment, chemical extraction, and supercritical regeneration [83,84,85,86]. In this study, thermal treatment experimental methods were investigated for the regeneration procedure since the chemical structure of the β-MgMoO4 removal agent was chemically stable.
The results indicate that β-MgMoO4 is amenable to regeneration via thermal treatment. Figure 10 demonstrates the recycling efficacy of β-MgMoO4 for the removal of methylene blue cationic dye for three consecutive rounds. The results demonstrate that the maximum removal capacity was 246 mg/g even after three cycles of use. It was discovered to be incredibly effective in regenerating β-MgMoO4 adsorbent by calcination at 400 °C in an air atmosphere, demonstrating its exceptional reusability.
In summary, the thermal treatment method is effective in regenerating β-MgMoO4 adsorbent, and these results suggest that it has excellent reusability for the removal of MB cationic dye from aqueous solutions. These findings are promising for the practical application of β-MgMoO4 as an adsorbent for the removal of MB cationic dye from wastewater.

2.4.2. FT-IR Spectroscopy

The spectra of β-MgMoO4 before and after exposure to MB dye were compared, as displayed in Figure 11. The FT-IR spectra show clear flexing and stretching vibrational characteristics of the metal–oxygen bonds located at frequencies lower than 1100 cm−1, corresponding to beta magnesium molybdate [79]. The spectrum of pure MB cationic dye revealed bands between 1700 and 1000 cm−1 [79]. After MB removal, further bands located at 1600 cm−1 were observed in the FT-IR spectrum of (MgMoO4-MB), which are credited to MB C=C bond stretching, indicating the existence of MB due to their attachment to the β-MgMoO4 active sites.
The FTIR spectrum of regenerated MgMoO4 (MgMoO4-MB-Reg) after thermal treatment was comparable to that of fresh MgMoO4, indicating that the attached MB cationic dye molecules on the surface completely combusted. The resulting spectrum confirms the cleanliness of the regenerated adsorbent and its reusability.

2.5. MB Removal Mechanism

The data showed that there was no evidence of any intermediate molecules or any chemical breaking down of MB cationic dye during its removal following the adsorption process.
The efficacy of MB cationic dye removal using β-MgMoO4 nanoparticles was found to decrease with an increase in pH values up to 11, which could be due to the acidic media. Building on this, we can propose a mechanism by which the protonation of MB cationic dye in acidic media leads to positively charged ammonium units (–N+) [66]. The oxygen atoms of the magnesium molybdate nanoparticles and these ammonium components engage electrostatically in the following step, enabling the adsorption of MB cationic dye on the catalytic surface. This mechanism is depicted in Figure 12.
In summary, the proposed mechanism suggests that the removal of MB by β-MgMoO4 nanoparticles occurs through an electrostatic interaction involving the nanoparticles’ surface and the positively charged ammonium entities of the protonated MB. These results suggest that β-MgMoO4 nanoparticles are a promising adsorbent for removing cationic dyes from wastewater.
The SEM images clearly illustrate the structural changes in β-MgMoO4 before and after the adsorption of MB cationic dye, as well as after the regeneration and reuse of the nanosorbent. The highly porous structure of the starting material allows for better adsorption of the dye, while the filled pores in the sample after adsorption indicate that the active sites on the surface were filled by MB cationic dye molecules. The agglomeration of particles after regeneration explains the decrease in dye removal efficiency, as some of the active sites may have been blocked by agglomerated particles. However, the morphology of β-MgMoO4 remains stable and unchanged even after multiple reuses and regeneration, indicating that it is a reliable and reusable nanosorbent for removing MB cationic dye from aqueous solutions.
Note that in all of the cases, the particles were of nanoscale size, and the size values obtained with different techniques were of the same order of magnitude. However, the value of approximately 150–200 nm for particles assembled in agglomerates of about 1–2 microns obtained using the SEM technique was much higher than that of 36 nm obtained using XRD. The smaller value from the XRD analysis can be explained by the fact that XRD determines the crystallite size and not the particle size. Agglomeration of the crystallites results in the formation of particles of greater size. As a result, there will be less adsorption of MB molecules in the case of an agglomerated powder (78% after regeneration). In XRD, the crystallite size is measured even if it is agglomerated into larger particles because the crystallites are separately crystallized (Figure 13).

3. Experimental Materials and Methods

3.1. Magnesium Molybdate Preparation

The chemicals, oxalic acid H2C2O4,2H2O, magnesium nitrate Mg(NO3)2,6H2O, and ammonium molybdate (NH4)6Mo7O24,4H2O, were acquired from Sigma-Aldrich (St. Louis, MO, USA) and consumed in their original states without any modifications, except for the MB dye, which was provided by Panreac (Barcelona, Spain). The magnesium molybdate nanosorbent was synthesized using a new method, the solid-state reaction synthesis method, from the magnesium molybdenum complex [50]. This new method is very simple to conduct; the preparation is carried out without the use of any solvents in two steps and a short time. The process involved mixing and grinding together magnesium nitrate, ammonium molybdate, and oxalic acid in their solid state in a ratio of 1/7:3:1, respectively. The resulting milled combination was then heated at 160 °C on a hotplate, and then burned inside a tubular furnace at a temperature of 500 °C under static air and kept for two continuous hours.

3.2. Experimental Adsorptions

Various concentrations of solutions were prepared from a stock solution, and the removal of MB cationic dye was performed by continuously stirring a specific amount of adsorbent into a 100 mL MB cationic dye solution with specified concentrations, at various temperatures, and with different contact times. The required pH of the testing solution was controlled by adjusting with a base (0.1N NaOH) and/or an acid (0.1N HCl). Towards the end of fixed time intervals, the solution was filtered and examined with a UV–visible spectrometer at λmax = 665 nm. The percentages of MB removed, the surface concentration, and the quantity adsorbed, at equilibrium, were calculated using specific formulas.
R e m o v a l % = C o C e C o × 100
Γ = C o C e M · a s × V
q e = C o C e M × V
where C0 is the initial concentration of the MB cationic dye in solution, Ce is the equilibrium concentration of the MB cationic dye, M is the mass of the used adsorbent, a s is the specific surface area, and V is the measured solution volume.
The results were obtained successively in triplicate, and the % uncertainty was found to be less than 3%, indicating that the results are precise and reliable.

3.3. Adsorbent Regeneration Procedure

A fixed concentration of 160 ppm at pH = 3 was used for the generation experiments. The fresh β-MgMoO4 was dried after filtration at 100 °C and then calcined at 400 °C for 1 h under atmospheric air. Then, the calcined β-MgMoO4 was checked for recycling potential under the same conditions as the fresh β-MgMoO4. The recycling process was repeated for three successive rounds under the same settings. The concentration of the removed MB in the solution was analyzed using a UV-Vis spectrophotometer analytical instrument, and the MB removal percentage as well as the amount removed at equilibrium were calculated using the formulas mentioned earlier.

3.4. Characterization

The phase composition of the synthesized nanosorbent before and after its use for the removal of MB dye was analyzed by the XRD technique (X-ray diffractometer 6000, Shimadzu, Tokyo, Japan, equipped with λCu-Kα = 1.5406 Ǻ and Ni filter). The XRD pattern exhibited the characteristic reflections of pure β-MgMoO4 phase, and they were indexed based on J.C.P.D.S card number 72-2153 (Figure 14). The Scherrer equation was used to estimate the particle size from the XRD pattern of the as-prepared nanoparticles as follows: DXRD = 0.9λ/(βcos θ); here, DXRD is the average particle diameter, λ is the Cu kα wavelength, β is the full-width at half-maximum (FWHM) of the diffraction peak, and θ is the diffraction angle. The intense peak at 2 θ = 26.64o (220), which correlates to the highest d spacing, was picked to determine the crystallite size DXRD, which was discovered to be 36 nm for both cases. A Micromeritics ASAP 2020 surface area and porosity analyzer, (Micromeritics, Norcross, GA, USA), was used to measure the adsorption–desorption isotherms.
The specific surface area of the magnesium molybdate MgMoO4 prepared was estimated by the Brunauer–Emmett–Teller technique (BET). It was found to be SBET = 20.3 m2/g. FTIR spectroscopy was performed using KBr pellets and IR Affinity 1S Shimadzu equipment (Shimadzu, Tokyo, Japan) within the range of 400 to 4000 cm−1 to identify the presence of the methylene blue cationic dye on β-MgMoO4 nanosorbent following the experimental adsorption and desorption tests.
SEM studies were carried out using a Quanta Feg-250 Thermo Fisher Scientific instrument (Thermo Fisher Scientific, Hillsboro, OR, USA) to obtain micrograph images and identify the adsorption–desorption steps of the methylene blue cationic dye on the β-MgMoO4 nanosorbent. UV–visible spectrophotometer studies were performed using a Thermo Scientific Genesys 10S instrument (Thermo Fisher Scientific, Madison, WI, USA) to determine the % removal of MB and the concentration at equilibrium.

4. Conclusions

β-MgMoO4 was successfully synthesized and investigated as an MB nanosorbent in aqueous solution. It demonstrated powerful pH dependence with an accomplished removal efficiency of 99% after only 10 min of contact time, with an initial 160 ppm dye concentration of pH = 3 employed. Additionally, the kinetic investigations demonstrated that MB removal conformed to a pseudo-second-order model. Thermodynamics suggested that the Langmuir isotherm was the most appropriate model in terms of its fit with the experimental adsorption data. Remarkably, calculations of the Langmuir model displayed that the removal amount reached a maximum of 356 mg/g. Moreover, β-MgMoO4 was effectively regenerated with success after calcination at 400 °C and could be reused multiple times without a significant reduction in efficiency. Remarkably, when β-MgMoO4 was tested, it still exhibited excellent removal performance for the studied MB dye, even after three consecutive rounds of reuse, while retaining high removal efficiency. Overall, the data presented that β-MgMoO4 is indeed an effective nanosorbent that is capable of outstanding removal performance and provided consistent results over multiple recycling tests.

Author Contributions

Conceptualization, A.M., S.R. and H.O.H.; Methodology, A.M., S.R., F.K. and M.A.; Validation, S.R., H.O.H. and M.A.; Formal analysis, A.M., H.O.H., S.A.P. and F.K.; Investigation, S.R., H.O.H., F.K., E.A. and M.A.; Resources, S.R.; Data curation, A.M., S.R. and H.O.H.; Writing—original draft, A.M.; Writing—review & editing, S.R., H.O.H., S.A.P., F.K., E.A. and M.A.; Visualization, A.M., S.R. and H.O.H.; Supervision, S.R.; Project administration, S.R.; Funding acquisition, S.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study and the samples of the compounds are available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wu, T.Y.; Tsao, M.H.; Chen, F.L.; Su, S.G.; Chang, C.W.; Wang, H.P.; Lin, Y.C.; Ou-Yang, W.C.; Sun, I.W. Synthesis and Characterization of Organic Dyes Containing Various Donors and Acceptors. Int. J. Mol. Sci. 2010, 11, 329–353. [Google Scholar] [CrossRef] [PubMed]
  2. Prokein, M.; Dyes, T.; Renner, M.; Weidner, E. Waterless Leather Dyeing with Dense Carbon Dioxide as Solvent for Dyes. J. Supercrit. Fluids 2021, 178, 105377. [Google Scholar] [CrossRef]
  3. Fatimah, I.; Citradewi, P.W.; Iqbal, R.M.; Ghazali, S.A.I.S.M.; Yahya, A.; Purwiandono, G. Geopolymer from Tin Mining Tailings Waste Using Salacca Leaves Ash as Activator for Dyes and Peat Water Adsorption. S. Afr. J. Chem. Eng. 2023, 43, 257–265. [Google Scholar] [CrossRef]
  4. Sivasangari, S.; Suseendhar, S.; Suresh Kumar, K.; Vijayaprasath, N.; Thirumurugan, M. Characteristic Study of Electroplating and Dye Industrial Effluents. Int. J. Innov. Res. Sci. Eng. Technol. ISO 2007, 3297, 20810–20816. [Google Scholar]
  5. Magdy, Y.H.; Altaher, H. Kinetic Analysis of the Adsorption of Dyes from High Strength Wastewater on Cement Kiln Dust. J. Environ. Chem. Eng. 2018, 6, 834–841. [Google Scholar] [CrossRef]
  6. Rusmini, R.; Aquastin, D.; Manullang, R.R.; Daryono, D. Tingkat Kesukaan Konsumen Terhadap Serat Kenaf Organik Dengan Pewarna Alami. J. Penelit. Pertan. Terap. 2019, 19, 58–65. [Google Scholar] [CrossRef]
  7. Karimov, K.S.; Ahmad, Z.; Bhadra, J.; Nabi, J.U.; Hong, M.; Bashir, M.M.; Ahkmedov, K.M.; Murtaza, I.; Khan, M.U. Design of a Semiconductor Photoelectrochemical Cells Using Orange Dye and NaCl Aqueous Solutions. Int. J. Electrochem. Sci. 2020, 15, 3189–3195. [Google Scholar] [CrossRef]
  8. Hong, I.H.; Shen, Z.; Chen, S.C.; Chen, A.; Tsai, K.C.; Lin, Y.T. Manufacturing Parameters Optimization in Functional Textile Dyeing Processes. Procedia Manuf. 2017, 11, 619–624. [Google Scholar] [CrossRef]
  9. Guerra, E.; Llompart, M.; Garcia-Jares, C. Analysis of Dyes in Cosmetics: Challenges and Recent Developments. Cosmetics 2018, 5, 47. [Google Scholar] [CrossRef]
  10. Pinzon, I.A.P.; Razal, R.A.; Mendoza, R.C.; Carpio, R.B.; Eusebio, R.C.P. Parametric Study on Microwave-Assisted Extraction of Runo (Miscanthus Sinensis Andersson) Dye and Its Application to Paper and Cotton Fabric. Biotechnol. Rep. 2020, 28, e00556. [Google Scholar] [CrossRef]
  11. Olas, B.; Białecki, J.; Urbańska, K.; Bryś, M. The Effects of Natural and Synthetic Blue Dyes on Human Health: A Review of Current Knowledge and Therapeutic Perspectives. Adv. Nutr. 2021, 12, 2301–2311. [Google Scholar] [CrossRef] [PubMed]
  12. Cai, J.; Jiang, H.; Chen, W.; Cui, Z. Design, Synthesis, Characterization of Water-Soluble Indophenine Dyes and Their Application for Dyeing of Wool, Silk and Nylon Fabrics. Dye. Pigment. 2020, 179, 108385. [Google Scholar] [CrossRef]
  13. De, I.; Pahuja, M.; Wani, H.M.U.D.; Dey, A.; Dube, T.; Ghosh, R.; Kankan, N.; Mishra, J.; Panda, J.J.; Maruyama, T.; et al. In-Vitro Toxicity Assessment of a Textile Dye Eriochrome Black T and Its Nano-Photocatalytic Degradation through an Innovative Approach Using Mf-NGr-CNTs-SnO2 Heterostructures. Ecotoxicol. Environ. Saf. 2022, 243, 113985. [Google Scholar] [CrossRef] [PubMed]
  14. Christie, R.; Abel, A. Cationic (Basic) Dye Complex Pigments. Phys. Sci. Rev. 2021, 6, 557–567. [Google Scholar] [CrossRef]
  15. Zhu, Y.Y.; Long, J.J. Development of a Novel Method for Investigation the Uptake Behaviors of Acid Dye on Silk Fabric at Various Flow Statuses in a Coloration Circular Pipe. Arab. J. Chem. 2022, 15, 104175. [Google Scholar] [CrossRef]
  16. Takijiri, K.; Morita, K.; Nakazono, T.; Sakai, K.; Ozawa, H. Highly Stable Chemisorption of Dyes with Pyridyl Anchors over TiO2: Application in Dye-Sensitized Photoelectrochemical Water Reduction in Aqueous Media. Chem. Commun. 2017, 53, 3042–3045. [Google Scholar] [CrossRef]
  17. Gorjanc, M.; Gerl, A.; Kert, M. Screen Printing of PH-Responsive Dye to Textile. Polymers 2022, 14, 447. [Google Scholar] [CrossRef]
  18. Verma, N.; Kumar, V.; Kesari, K.K. Microbial and Lignocellulosic Biomass Based Dye Decolourization. Biomass Convers. Biorefinery 2022, 13, 16643–16666. [Google Scholar] [CrossRef]
  19. Sharma, J.; Sharma, S.; Soni, V. Classification and Impact of Synthetic Textile Dyes on Aquatic Flora: A Review. Reg. Stud. Mar. Sci. 2021, 45, 101802. [Google Scholar] [CrossRef]
  20. Pierrat, É.; Laurent, A.; Dorber, M.; Rygaard, M.; Verones, F.; Hauschild, M. Advancing Water Footprint Assessments: Combining the Impacts of Water Pollution and Scarcity. Sci. Total Environ. 2023, 870, 161910. [Google Scholar] [CrossRef]
  21. Rakass, S.; Mohmoud, A.; Hassani, H.O.; Abboudi, M.; Kooli, F.; Al 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]
  22. 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]
  23. Türgay, O.; Ersöz, 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. Technol. 2011, 79, 26–33. [Google Scholar] [CrossRef]
  24. 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]
  25. 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]
  26. Pǎcurariu, C.; Paşka, O.; Ianoş, R.; Muntean, S.G. 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] [CrossRef]
  27. 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]
  28. Forgacs, E.; Cserháti, T.; Oros, G. Removal of Synthetic Dyes from Wastewaters: A Review. Environ. Int. 2004, 30, 953–971. [Google Scholar] [CrossRef]
  29. Filice, S.; D’Angelo, D.; Libertino, S.; Nicotera, I.; Kosma, V.; Privitera, V.; Scalese, S. Graphene Oxide and Titania Hybrid Nation Membranes for Efficient Removal of Methyl Orange Dye from Water. Carbon 2015, 82, 489–499. [Google Scholar]
  30. Kandisa, R.V.; Saibaba KV, N. Dye Removal by Adsorption: A Review. J. Bioremediat Biodegrad. 2016, 07, 371. [Google Scholar] [CrossRef]
  31. Piaskowski, K.; Świderska-Dąbrowska, R.; Zarzycki, P.K. Dye Removal from Water and Wastewater Using Various Physical, Chemical, and Biological Processes. J. AOAC Int. 2018, 101, 1371–1384. [Google Scholar] [CrossRef] [PubMed]
  32. Mohmoud, A.; Rakass, S.; Hassani, H.O.; Kooli, F.; Abboudi, M.; Aoun, S. Ben Iron Molybdate Fe2(MoO4)3 Nanoparticles: Efficient Sorbent for Methylene Blue Dye Removal from Aqueous Solutions. Molecules 2020, 25, 3390. [Google Scholar] [CrossRef] [PubMed]
  33. Madeley, R.A.; Wanke, S.E. Variation of the Dispersion of Active Phases in Commercial Nickel-Molybdenum/γ-Alumina Hydrotreating Catalysts during Oxidative Regeneration. Appl. Catal. 1988, 39, 295–314. [Google Scholar] [CrossRef]
  34. Gates, B.C.; Katzer, J.R.; Schuit, G.C.A. Chemistry of Catalytic Processes; McGraw-Hill: New York, NY, USA, 1979; Volume 60. [Google Scholar]
  35. Rodriguez, J.A.; Chaturvedi, S.; Hanson, J.C.; Brito, J.L. Reaction of H2 and H2S with CoMoO4 and NiMoO4: TPR, XANES, Time-Resolved XRD, and Molecular-Orbital Studies. J. Phys. Chem. B 1999, 103, 770–781. [Google Scholar] [CrossRef]
  36. Sundaram, R.; Nagaraja, K.S. Solid State Electrical Conductivity and Humidity Sensing Studies on Metal Molybdate-Molybdenum Trioxide Composites (M = Ni2+, Cu2+ and Pb2+). Sens. Actuators B Chem. 2004, 101, 353–360. [Google Scholar] [CrossRef]
  37. Kaddouri, A.; Anouchinsky, R.; Mazzocchia, C.; Madeira, L.M.; Portela, M.F. Oxidative Dehydrogenation of Ethane on the α and β Phases of NiMoO4. Catal. Today 1998, 40, 201–206. [Google Scholar] [CrossRef]
  38. Pillay, B.; Mathebula, M.R.; Friedrich, H.B. The Oxidative Dehydrogenation of N-Hexane over Ni-Mo-O Catalysts. Appl. Catal. A Gen. 2009, 361, 57–64. [Google Scholar] [CrossRef]
  39. Mi, Y.; Huang, Z.; Hu, F.; Jiang, J.; Li, Y. Controlled Synthesis and Growth Mechanism of Alpha Nickel Molybate Microhombohedron. Mater. Lett. 2010, 64, 695–697. [Google Scholar] [CrossRef]
  40. Brito, J.L.; Barbosa, A.L.; Albornoz, A.; Severino, F.; Laine, J. Nickel Molybdate as Precursor of HDS Catalysts: Effect of Phase Composition. Catal. Lett. 1994, 26, 329–337. [Google Scholar] [CrossRef]
  41. Edrissi, M.; Samadanian-Isfahani, S.; Soleymani, M. Preparation of Cobalt Molybdate Nanoparticles; Taguchi Optimization and Photocatalytic Oxidation of Reactive Black 8 Dye. Powder Technol. 2013, 249, 378–385. [Google Scholar] [CrossRef]
  42. Ghoreishian, S.M.; Seeta Rama Raju, G.; Pavitra, E.; Kwak, C.H.; Han, Y.K.; Huh, Y.S. Controlled Synthesis of Hierarchical α-Nickel Molybdate with Enhanced Solar-Light-Responsive Photocatalytic Activity: A Comprehensive Study on the Kinetics and Effect of Operational Factors. Ceram. Int. 2019, 45, 12041–12052. [Google Scholar] [CrossRef]
  43. Ferreira, E.A.C.; Andrade Neto, N.F.; Bomio, M.R.D.; Motta, F.V. Influence of Solution PH on Forming Silver Molybdates Obtained by Sonochemical Method and Its Application for Methylene Blue Degradation. Ceram. Int. 2019, 45, 11448–11456. [Google Scholar] [CrossRef]
  44. Rashad, M.M.; Ibrahim, A.A.; Rayan, D.A.; Sanad, M.M.S.; Helmy, I.M. Photo-Fenton-like Degradation of Rhodamine B Dye from Waste Water Using Iron Molybdate Catalyst under Visible Light Irradiation. Environ. Nanotechnol. Monit. Manag. 2017, 8, 175–186. [Google Scholar] [CrossRef]
  45. Oudghiri-Hassani, H. Synthesis, Characterization and Application of Chromium Molybdate for Oxidation of Methylene Blue Dye. J. Mater. Environ. Sci. 2018, 9, 1051–1057. [Google Scholar] [CrossRef]
  46. Esmaeili, H.; Foroutan, R.; Jafari, D.; Aghil Rezaei, M. Effect of Interfering Ions on Phosphate Removal from Aqueous Media Using Magnesium Oxide@ferric Molybdate Nanocomposite. Korean J. Chem. Eng. 2020, 37, 804–814. [Google Scholar] [CrossRef]
  47. Santiago, A.A.G.; Oliveira, M.C.; Ribeiro, R.A.P.; Tranquilin, R.L.; Longo, E.; De Lázaro, S.R.; Motta, F.V.; Bomio, M.R.D. Atomistic Perspective on the Intrinsic White-Light Photoluminescence of Rare-Earth Free MgMoO4 Nanoparticles. Cryst. Growth Des. 2020, 20, 6592–6603. [Google Scholar] [CrossRef]
  48. Cadus, L.E.; Abello, M.C.; Gomez, M.F.; Rivarola, J.B. Oxidative Dehydrogenation of Propane over Mg-Mo-O Catalysts. Ind. Eng. Chem. Res. 1996, 35, 14–18. [Google Scholar] [CrossRef]
  49. Miller, J.E.; Jackson, N.B.; Evans, L.; Sault, A.G.; Gonzales, M.M. The Formation of Active Species for Oxidative Dehydrogenation of Propane on Magnesium Molybdates. Catal. Lett. 1999, 58, 147–152. [Google Scholar] [CrossRef]
  50. Al Wadaani, F. Nano-Structured ß-MgMoO4: Synthesis, Characterization and Catalytic Efficiency in the 3-Nitrophenol Reduction. Moroc. J. Chem. 2016, 4, 332–338. [Google Scholar]
  51. Xavier, C.S.; de Moura, A.P.; Longo, E.; Varela, J.A.; Zaghete, M.A. Synthesis and Optical Property of MgMoO4 Crystals. In Proceedings of the Advanced Materials Research; 2014; Volume 975. [Google Scholar]
  52. Mikhailik, V.B.; Kraus, H.; Kapustyanyk, V.; Panasyuk, M.; Prots, Y.; Tsybulskyi, V.; Vasylechko, L. Structure, Luminescence and Scintillation Properties of the MgWO 4-MgMoO4 System. J. Phys. Condens. Matter 2008, 20, 365219. [Google Scholar] [CrossRef]
  53. Zhou, L.Y.; Wei, J.S.; Yi, L.H.; Gong, F.Z.; Huang, J.L.; Wang, W. A Promising Red Phosphor MgMoO4:Eu3+ for White Light Emitting Diodes. Mater. Res. Bull. 2009, 44, 1411–1414. [Google Scholar] [CrossRef]
  54. Mo, F.; Chen, P.; Guan, A.; Zhang, W.; Zhou, L. Enhancement of Luminescence Intensity and Color Purity of MgxZn1-XMoO4:Eu3+,Bi3+ Phosphors. J. Rare Earths 2015, 33, 1064–1071. [Google Scholar] [CrossRef]
  55. Cornu, L.; Jubera, V.; Demourgues, A.; Salek, G.; Gaudon, M. Luminescence Properties and Pigment Properties of A-Doped (Zn, Mg)MoO4 Triclinc Oxides (with A=Co, Ni, Cu or Mn). Ceram. Int. 2017, 53, 13377–13378. [Google Scholar] [CrossRef]
  56. Ran, W.; Wang, L.; Yang, M.; Kong, X.; Qu, D.; Shi, J. Echanced Energy Transfer from Bi3+ to Eu3+ Ions Relying on the Criss-Cross Cluster Structure in MgMoO4 Phosphor. J. Lumin. 2017, 192, 141–147. [Google Scholar] [CrossRef]
  57. Zhang, L.; He, W.; Shen, K.; Liu, Y.; Guo, S. Controllable Synthesis of Hierarchical MgMoO4 Nanosheet-Arrays and Nano-Flowers Assembled with Mesoporous Ultrathin Nanosheets. J. Phys. Chem. Solids 2018, 115, 215–220. [Google Scholar] [CrossRef]
  58. Amberg, M.; Günter, J.R.; Schmalle, H.; Blasse, G. Preparation, Crystal Structure, and Luminescence of Magnesium Molybdate and Tungstate Monohydrates, MgMoO4 · H2O and MgWO4 · H2O. J. Solid. State Chem. 1988, 77, 162–169. [Google Scholar] [CrossRef]
  59. Yousefipour, K.; Sarraf-Mamoory, R.; Mollayousefi, S. Synthesis of Manganese Molybdate/MWCNT Nanostructure Composite with a Simple Approach for Supercapacitor Applications. RSC Adv. 2022, 12, 27868–27876. [Google Scholar] [CrossRef]
  60. Haetge, J.; Suchomski, C.; Brezesinski, T. Ordered Mesoporous β-MgMoO4 Thin Films for Lithium-Ion Battery Applications. Small Nanomicro 2013, 9, 2541–2544. [Google Scholar] [CrossRef]
  61. Lee, E.K.; Jung, K.D.; Joo, O.S.; Shul, Y.G. Catalytic Activity of Mo/MgO Catalyst in the Wet Oxidation of H2S to Sulfur at Room Temperature. Appl. Catal. A Gen. 2004, 268, 83–88. [Google Scholar] [CrossRef]
  62. Andreichikova, M.P.; Udalova, T.A. Mechanochemical Synthesis of Highly Dispersed Molybdenum from Magnesium Molybdate. Mater. Today Proc. 2020, 31, 526–528. [Google Scholar] [CrossRef]
  63. Bian, W.; Lu, X.; Wang, Y.; Zhu, H.; Chen, T.; Ta, S.; Zhang, Q. Correlations between Structure and Microwave Dielectric Properties of Co Doped MgMoO4 Ceramics. Ceram. Int. 2020, 46, 22024–22029. [Google Scholar] [CrossRef]
  64. Braziulis, G.; Janulevicius, G.; Stankeviciute, R.; Zalga, A. Aqueous Sol-Gel Synthesis and Thermoanalytical Study of the Alkaline Earth Molybdate Precursors. J. Therm. Anal. Calorim. 2013, 118, 613–621. [Google Scholar] [CrossRef]
  65. Ganguly, P.; Sarkhel, R.; Das, P. Synthesis of Pyrolyzed Biochar and Its Application for Dye Removal: Batch, Kinetic and Isotherm with Linear and Non-Linear Mathematical Analysis. Surf. Interfaces 2020, 20, 100616. [Google Scholar] [CrossRef]
  66. Oudghiri-Hassani, H.; Rakass, S.; Abboudi, M.; Mohmoud, A.; AlWadaani, F. Preparation and Characterization of α-Zinc Molybdate Catalyst: Efficient Sorbent for Methylene Blue and Reduction of 3-Nitrophenol. Molecules 2018, 23, 3390. [Google Scholar] [CrossRef]
  67. Kim, J.R.; Santiano, B.; Kim, H.; Kan, E. Heterogeneous Oxidation of Methylene Blue with Surface-Modified Iron-Amended Activated Carbon. Am. J. Anal. Chem. 2013, 04, 115–122. [Google Scholar] [CrossRef]
  68. Kooli, F.; Liu, Y.; Hbaieb, K.; Ching, O.Y.; Al-Faze, R. Characterization of Organo-Kenyaites: Thermal Stability and Their Effects on Eosin Removal Characteristics. Clay Min. 2018, 53, 91–104. [Google Scholar] [CrossRef]
  69. Popoola, S.A.; Al Dmour, H.; Rakass, S.; Fatimah, I.; Liu, Y.; Mohmoud, A.; Kooli, F. Enhancement Properties of Zr Modified Porous Clay Heterostructures for Adsorption of Basic-Blue 41 Dye: Equilibrium, Regeneration, and Single Batch Design Adsorber. Materials 2022, 15, 5567. [Google Scholar] [CrossRef]
  70. 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]
  71. Anoop 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]
  72. Sahu, S.; Pahi, S.; Tripathy, S.; Singh, S.K.; Behera, A.; Sahu, U.K.; Patel, R.K. Adsorption of Methylene Blue on Chemically Modified Lychee Seed Biochar: Dynamic, Equilibrium, and Thermodynamic Study. J. Mol. Liq. 2020, 315, 113743. [Google Scholar] [CrossRef]
  73. 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. J. Environ. Sci. Vol. 2011, 1, 711–726. [Google Scholar]
  74. Vadivelan, V.; Vasanth Kumar, K. 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]
  75. Hai Nguyen, T. Applying Linear Forms of Pseudo-Second-Order Kinetic Model for Feasibly Identifying Errors in the Initial Periods of Time-Dependent Adsorption Datasets. Water 2023, 15, 1231. [Google Scholar] [CrossRef]
  76. Furusawa, T.; Smith, J.M. Intraparticle Mass Transport in Slurries by Dynamic Adsorption Studies. AIChE J. 1974, 20, 88–93. [Google Scholar] [CrossRef]
  77. Wang, J.; Guo, X. Adsorption Kinetic Models: Physical Meanings, Applications, and Solving Methods. J. Hazard. Mater. 2020, 390, 122156. [Google Scholar] [CrossRef]
  78. Ezzati, R. Derivation of Pseudo-First-Order, Pseudo-Second-Order and Modified Pseudo-First-Order Rate Equations from Langmuir and Freundlich Isotherms for Adsorption. Chem. Eng. J. 2020, 392, 123705. [Google Scholar] [CrossRef]
  79. Hu, Q.; Zhang, Z. Application of Dubinin–Radushkevich Isotherm Model at the Solid/Solution Interface: A Theoretical Analysis. J. Mol. Liq. 2019, 277, 646–648. [Google Scholar] [CrossRef]
  80. Dada, A.O.; Olalekan, A.P.; Olatunya, A.M. Dada Langmuir, Freundlich, Temkin and Dubinin-Radushkevich Isotherms Studies of Equilibrium Sorption of Zn 2+ Unto Phosphoric Acid Modified Rice Husk. J. Appl. Chem. 2012, 3, 38–45. [Google Scholar]
  81. 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]
  82. Xiong, L.; Yang, Y.; Mai, J.; Sun, W.; Zhang, C.; Wei, D.; Chen, Q.; Ni, J. Adsorption Behavior of Methylene Blue onto Titanate Nanotubes. Chem. Eng. J. 2010, 156, 313–320. [Google Scholar] [CrossRef]
  83. 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. [Google Scholar] [CrossRef]
  84. Sadatshojaei, E.; Esmaeilzadeh, F.; Fathikaljahi, J.; Hosseynian Barzi, S.E.; Wood, D.A. Regeneration of the Midrex Reformer Catalysts Using Supercritical Carbon Dioxide. Chem. Eng. J. 2018, 343, 748–758. [Google Scholar] [CrossRef]
  85. Sun, Y.; Zhang, B.; Zheng, T.; Wang, P. Regeneration of Activated Carbon Saturated with Chloramphenicol by Microwave and Ultraviolet Irradiation. Chem. Eng. J. 2017, 320, 264–270. [Google Scholar] [CrossRef]
  86. El Gamal, M.; Mousa, H.A.; El-Naas, M.H.; Zacharia, R.; Judd, S. Bio-Regeneration of Activated Carbon: A Comprehensive Review. Sep. Purif. Technol. 2018, 197, 345–359. [Google Scholar] [CrossRef]
Figure 1. Removal efficiency of β-MgMoO4 in a 100 ppm MB cationic solution as function of pH (mads = 0.1 g; T = 20 °C; t = 30 min).
Figure 1. Removal efficiency of β-MgMoO4 in a 100 ppm MB cationic solution as function of pH (mads = 0.1 g; T = 20 °C; t = 30 min).
Molecules 30 01606 g001
Figure 2. Removal efficacy of β-MgMoO4 in 160 ppm MB cationic dye solution as function of adsorption dosage (time = 30 min; temperature = 20 °C).
Figure 2. Removal efficacy of β-MgMoO4 in 160 ppm MB cationic dye solution as function of adsorption dosage (time = 30 min; temperature = 20 °C).
Molecules 30 01606 g002
Figure 3. Removal efficacy of β-MgMoO4 for MB as a function of initial cationic dye concentration and contact time (pH = 3; madsorbent = 0.05 g; and T = 20 °C).
Figure 3. Removal efficacy of β-MgMoO4 for MB as a function of initial cationic dye concentration and contact time (pH = 3; madsorbent = 0.05 g; and T = 20 °C).
Molecules 30 01606 g003
Figure 4. Removal efficiency of β-MgMoO4 with a 200 ppm MB cationic solution, as a function of temperature (m = 0.05 g; t = 30 min; and pH = 3).
Figure 4. Removal efficiency of β-MgMoO4 with a 200 ppm MB cationic solution, as a function of temperature (m = 0.05 g; t = 30 min; and pH = 3).
Molecules 30 01606 g004
Figure 5. Van’t Hoff plot displaying effects of temperature parameters on MB cationic dye removal by β-MgMoO4.
Figure 5. Van’t Hoff plot displaying effects of temperature parameters on MB cationic dye removal by β-MgMoO4.
Molecules 30 01606 g005
Figure 6. Intra-particle diffusion model plot displaying the effects of contact time and the initial cationic dye concentration of MB removal via β-MgMoO4.
Figure 6. Intra-particle diffusion model plot displaying the effects of contact time and the initial cationic dye concentration of MB removal via β-MgMoO4.
Molecules 30 01606 g006
Figure 7. Pseudo-first-order model plot displaying the effects of the initial cationic dye concentration and contact time on MB removal via β-MgMoO4.
Figure 7. Pseudo-first-order model plot displaying the effects of the initial cationic dye concentration and contact time on MB removal via β-MgMoO4.
Molecules 30 01606 g007
Figure 8. Pseudo-second-order model plot displaying the effects of the initial cationic dye concentration and contact time of MB removal by β-MgMoO4.
Figure 8. Pseudo-second-order model plot displaying the effects of the initial cationic dye concentration and contact time of MB removal by β-MgMoO4.
Molecules 30 01606 g008
Figure 9. Graphs of (a) Freundlich and (b) Langmuir isotherms demonstrating initial cationic dye concentration effects on MB removal by β-MgMoO4.
Figure 9. Graphs of (a) Freundlich and (b) Langmuir isotherms demonstrating initial cationic dye concentration effects on MB removal by β-MgMoO4.
Molecules 30 01606 g009
Figure 10. Recycling competences of β-MgMoO4 for removal of methylene blue cationic dye (160 ppm, 0.05 g, 30 min).
Figure 10. Recycling competences of β-MgMoO4 for removal of methylene blue cationic dye (160 ppm, 0.05 g, 30 min).
Molecules 30 01606 g010
Figure 11. FT-IR spectra of MgMoO4, MgMoO4-MB, MgMoO4-MB-Reg, and MB.
Figure 11. FT-IR spectra of MgMoO4, MgMoO4-MB, MgMoO4-MB-Reg, and MB.
Molecules 30 01606 g011
Figure 12. Schematic mechanism of MB cationic dye removal employing magnesium molybdate nanosorbent.
Figure 12. Schematic mechanism of MB cationic dye removal employing magnesium molybdate nanosorbent.
Molecules 30 01606 g012
Figure 13. SEM micrographs of (A) pure magnesium molybdate (β-MgMoO4), (B) the morphology of β-MgMoO4 after MB dye removal, (C) the morphology of β-MgMoO4 after the first regeneration, (D) the morphology of β-MgMoO4 after the second removal cycle of MB cationic dye, (E) the morphology of β-MgMoO4 after the second regeneration process, (F) the morphology of β-MgMoO4 after the third removal cycle of MB cationic dye, and (G) the morphology of β-MgMoO4 after the third regeneration process.
Figure 13. SEM micrographs of (A) pure magnesium molybdate (β-MgMoO4), (B) the morphology of β-MgMoO4 after MB dye removal, (C) the morphology of β-MgMoO4 after the first regeneration, (D) the morphology of β-MgMoO4 after the second removal cycle of MB cationic dye, (E) the morphology of β-MgMoO4 after the second regeneration process, (F) the morphology of β-MgMoO4 after the third removal cycle of MB cationic dye, and (G) the morphology of β-MgMoO4 after the third regeneration process.
Molecules 30 01606 g013
Figure 14. X-ray diffraction pattern of the prepared MgMoO4 nanoparticle powder before and after adsorption of MB.
Figure 14. X-ray diffraction pattern of the prepared MgMoO4 nanoparticle powder before and after adsorption of MB.
Molecules 30 01606 g014
Table 1. MB removal by β-MgMoO4: thermodynamic parameters.
Table 1. MB removal by β-MgMoO4: thermodynamic parameters.
AdsorbentAdsorbate∆H° (KJ·mol−1)∆S° (KJ·mol−1·K)∆G° (KJ·mol−1)
β-MgMoO4MB−32.463−0.017293 K313323 K333 K
−27.425−25.649−24.699−23.422
Table 2. Kinetic models’ equations.
Table 2. Kinetic models’ equations.
ModelEquationParameters
Pseudo-first-order (PFD) [58]. L n ( q e q t ) = L n q e + K 1 t (4)qt stands for the removal capacity at time t (mg/g);
qe stands for the removal capacity at equilibrium (mg/g);
K1 stands for the rate constant of pseudo-first-order adsorption (1/min).
Pseudo-second-order (PSD) [75]. 1 q t = ( 1 K 2 q e 2 ) 1 t + t q e
(The plot of 1/qt vs.1/t) (5)
qt stands for the removal capacity at time t (mg/g);
qe stands for the removal capacity at equilibrium (mg/g);
K2 stands for the pseudo-second-order rate constant (g·mg−1·min−1).
Intra-particle diffusion (IPD) [76]. q t = K I t 0.5 + l (6)I (mg/g) and KI (mg/(g·min0.5)) are the intra-particle diffusion constants;
qt stands for the removal capacity (mg/g) at time t;
t stands for the contact time (min).
Table 3. MB cationic dye removal by β-MgMoO4: kinetic parameters.
Table 3. MB cationic dye removal by β-MgMoO4: kinetic parameters.
Dye Cimg/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
1603191450.3410.9753192.200.97919041.170.958
1703181100.1460.7493051.640.98716543.710.810
1803301290.1840.9273161.650.99016549.040.884
2003601670.1620.9393411.240.99814959.630.904
Table 4. Adsorption isotherm models for MB cationic dye removal via β-MgMoO4.
Table 4. Adsorption isotherm models for MB cationic dye removal via β-MgMoO4.
ModelEquationParameters
Freundlich [78]. L n q e = L n q F + 1 n L n C e
(7)
qF stands for the Freundlich constant (mg(1−1/n)L1/ng−1),
n: stands for the heterogeneity factor (g/L);
qe stands for the amount of MB cationic dye adsorbed by β-MgMoO4 at equilibrium (mg/g);
Ce stands for MB the cationic dye concentration at equilibrium (ppm).
Langmuir [78]. C e q e = 1 q m K L + C e q m
(8)
qe stands for the amount of MB cationic dye adsorbed by β-MgMoO4 at equilibrium (mg/g);
Ce stands for the MB cationic dye concentration at equilibrium (ppm);
qm stands for the highest amount of MB cationic dye removed by β-MgMoO4 (mg/g);
KL stands for the Langmuir adsorption constant (L/mg).
R L = 1 1 + K L C i
(9)
Ci stands for the initial cationic MB dye concentration;
KL stands for the Langmuir constant;
RL stands for the values that indicate that the removal of MB cationic dye may be either linear (RL = 1), irreversible (RL = 0), favorable (0 < RL< 1), or unfavorable (RL > 1).
Dubinin–Radushkevich (D-R) [79]. L n q e = L n q m K ε 2 (10)
ε = R T L n ( 1 + 1 C e )
(11)
K stands for the sorption energy constant (mol2/kJ2);
ε stands for the Polanyi potential;
T stands for the temperature (K);
R stands for the universal gas constant (8.314 J·mol−1 K−1);
qm stands for the theoretical saturation capacity;
Ce stands for MB cationic dye concentration at equilibrium (ppm).
Temkin [80]. q e = B T L n A T + B T L n C e
(12)
bT stands for the Temkin constant associated with the heat of sorption (J/mol);
BT = RT/bT;
R stands for the gas constant (8.314 J/mol K);
AT stands for the Temkin isotherm constant (L/g);
T stands for the absolute temperature (K).
Table 5. MB removal by β-MgMoO4: isotherm parameters.
Table 5. MB removal by β-MgMoO4: isotherm parameters.
LangmuirFreundlichTemkinDubinin–Radushkevich
qm (mg/g)KL (L/mg)R2Range RLqF (mg(1−1/n)L1/ng−1)1/nR2AT (L/g)BTR2qm (mg/g)R2E (Kj/mol)
3561.630.9910.0031–0.00383150.0250.3433 × 10168.3060.3383430.21012.8
Table 6. Maximum adsorption capacity (qm) of different nanosorbents for the removal of MB cationic dye as reported previously in the literature.
Table 6. Maximum adsorption capacity (qm) of different nanosorbents for the removal of MB cationic dye as reported previously in the literature.
Nanosorbentqm (mg/g)Reference
Magnetic iron oxide nanosorbent25.54[26]
Fe3O4 magnetic nanoparticles modified with 3-glycidoxypropyltrimethoxysilane and glycine158.00[81]
Zinc molybdate nanoparticles217.86[66]
Calcined titanate nanotubes133.33[82]
Molybdenum trioxide nanorods and stacked nanoplates152.00[83]
Magnesium molybdate (β-MgMoO4)356.00This current work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mohmoud, A.; Rakass, S.; Oudghiri Hassani, H.; Popoola, S.A.; Kooli, F.; Assirey, E.; Abboudi, M. Magnesium Molybdate: An Efficient Nanosorbent for Methylene Blue Cationic Dye Removal from Aqueous Solutions. Molecules 2025, 30, 1606. https://doi.org/10.3390/molecules30071606

AMA Style

Mohmoud A, Rakass S, Oudghiri Hassani H, Popoola SA, Kooli F, Assirey E, Abboudi M. Magnesium Molybdate: An Efficient Nanosorbent for Methylene Blue Cationic Dye Removal from Aqueous Solutions. Molecules. 2025; 30(7):1606. https://doi.org/10.3390/molecules30071606

Chicago/Turabian Style

Mohmoud, Ahmed, Souad Rakass, Hicham Oudghiri Hassani, Saheed A. Popoola, Fethi Kooli, Eman Assirey, and Mostafa Abboudi. 2025. "Magnesium Molybdate: An Efficient Nanosorbent for Methylene Blue Cationic Dye Removal from Aqueous Solutions" Molecules 30, no. 7: 1606. https://doi.org/10.3390/molecules30071606

APA Style

Mohmoud, A., Rakass, S., Oudghiri Hassani, H., Popoola, S. A., Kooli, F., Assirey, E., & Abboudi, M. (2025). Magnesium Molybdate: An Efficient Nanosorbent for Methylene Blue Cationic Dye Removal from Aqueous Solutions. Molecules, 30(7), 1606. https://doi.org/10.3390/molecules30071606

Article Metrics

Back to TopTop