Next Article in Journal
Modelling of Low-Temperature Sulphur Dioxide Removal Using Response Surface Methodology (RSM), Artificial Neural Network (ANN) and Adaptive Neuro-Fuzzy Inference System (ANFIS)
Previous Article in Journal
Fault Detection of Multi-Rate Two Phase Reactor Condenser System with Recycle Using Multiple Probabilistic Principal Component Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Fabrication of Aluminum-Based Hybrid Nanocomposite for Photocatalytic Degradation of Methylene Blue Dye: A Techno-Economic Approach †

1
Environmental Engineering Department, Egypt-Japan University of Science and Technology (E-JUST), Alexandria 21934, Egypt
2
Department of Chemical Science and Engineering, Tokyo Institute of Technology, Tokyo 152-8552, Japan
3
Sanitary Engineering Department, Faculty of Engineering, Alexandria University, Alexandria 21544, Egypt
*
Author to whom correspondence should be addressed.
Presented at the 2nd International Electronic Conference on Processes: Process Engineering—Current State and Future Trends (ECP 2023), 17–31 May 2023; Available online: https://ecp2023.sciforum.net/.
Eng. Proc. 2023, 37(1), 87; https://doi.org/10.3390/ECP2023-14638
Published: 17 May 2023

Abstract

:
Al2O3–MgO nanocomposite was synthesized using the co-precipitation method for photocatalytic degradation of methylene blue (MB) dye under UV–Vis light. Box–Behnken design (BBD) in response to surface methodology (RSM) was used for the optimization and modelling of the photocatalytic degradation of MB dye. An analysis of variance (ANOVA) revealed a quadratic model with an R2 of 0.9880. MB removal followed the first-order kinetic model (R2 = 0.9520, k1 = 0.007 min−1). Economic feasibility study at optimized conditions showed that the wastewater treatment cost is USD 16.50/m3 and the payback period is 3.17 years.

1. Introduction

Methylene blue (MB) is a highly toxic cationic dye widely used in textile, pharmaceutical and printing industries. Exposure to MB dye results in physiological disorders, leading to acute and chronic illnesses such as vomiting, eye irritation and jaundice [1,2]. The discharge of MB-laden wastewater has detrimental effects on aquatic biota. MB changes the aesthetics of aquatic systems, leading to low sunlight available for photosynthesis, resulting in stunted growth and the death of aquatic plants. MB dye forms complex products in aqueous systems, leading to low dissolved oxygen and the death of aquatic organisms, resulting in a loss of biodiversity in marine ecosystems [2].
MB dye can be removed by conventional wastewater treatment methods such as adsorption, coagulation and phytoremediation [1]. However, the major drawback of conventional systems is the formation of secondary pollutants. Additionally, there is a high cost for the regeneration of adsorbents and disposal of pollutant-laden sludge [3].
Currently, advanced oxidation processes (AOPs) have been adopted for the degradation of organic pollutants by utilizing highly reactive hydroxyl radicals [4]. AOPs such as photocatalysis and photo-Fenton reaction can degrade MB dye into smaller and non-toxic molecules [3]. Due to the high stability and efficiency of TiO2 and ZnO nano-semiconductors, they are widely used in the photodegradation of organic dyes [4].
Al2O3 has high thermal stability, a large surface area and well-defined nanopore structures [5]. It is widely used in sensors, energy storage and pharmaceutical applications [6]. Al2O3 has significant optical properties; hence, it can be used as a photocatalyst [5]. Al2O3 can be synthesized from abundant waste aluminum cans [7], resulting in lower economic costs in comparison to TiO2 and ZnO.
Al2O3 has a relatively high band gap energy of 5.97 eV [8], and hence a lower photocatalytic activity. Al2O3 can be combined with MgO to form a nanocomposite with higher photocatalytic efficiency. MgO has a large surface area and structural defects that allow for the generation of reactive radicals under UV radiation [9]. MgO can form a heterojunction with Al2O3 to lower the latter’s band gap energy. This leads to an improved photocatalytic degradation efficiency.
This study focuses on the synthesis of Al2O3–MgO nanocomposite using co-precipitation for the photodegradation of MB dye. The effect of operating parameters in MB photodegradation is investigated. An economic feasibility analysis is carried out to assess the costs of the photocatalytic degradation of MB using Al2O3–MgO.

2. Materials and Methods

2.1. Materials

Aluminum nitrate nonahydrate (Al(NO3)3.9H2O) was purchased from Qualikems Fine Chem Ltd., Vadodara, Gujarat, India. Magnesium chloride hexahydrate (MgCl2·6H2O) was purchased from Tekkim, Bursa, Turkey. Methylene blue (C16H18ClN3S, 99%) was bought from Alpha Chemical group, 6th of October, Egypt. Methanol (CH3OH, 99.8%) and sulfuric acid (H2SO4, 97%) were purchased from Piochem, 6th of October, Egypt. Sodium hydroxide pellets (NaOH, 97%) were bought from SD fine Chem Ltd., Mumbai, India.

2.2. Photocatalyst Synthesis

The photocatalyst was prepared via a simple co-precipitation technique [8] as shown in Figure S1. Briefly, 0.2 M Al(NO3)3·9H2O was mixed with 0.2 M MgCl2·6H2O, and 1 M NaOH was added dropwise until pH was 11. A white precipitate formed with constant stirring for 30 min. The precipitate was washed with distilled water, filtered, dried for 12 h at 105 °C and finally calcined at 800 °C for 3 h to obtain Al2O3–MgO nanocomposite.

2.3. Characterization

The optical band gap energy was measured using a Jasco V-570 UV-DRS spectrophotometer (Tokyo, Japan) in the range of 220–2000 nm. The point of zero charge was determined using the batch equilibrium method [4].

2.4. Experimental Setup

A stock solution of 50 ppm of methylene blue solution was prepared by dissolving 50 mg of methylene blue dye in 1L of distilled water. The required solutions were prepared using the stock solution. Photodegradation of MB dye was carried out in a photoreactor with a 400 W metal halide lamp. The pH of the MB dye solution was adjusted using 1 M NaOH or 1 M H2SO4. The experiment was initially run in the dark for 60 min to achieve adsorption–desorption equilibrium. The concentration of MB after adsorption (Co) was measured at 664 nm using a UV spectrophotometer (Jasco V-630 spectrophotometer (Tokyo, Japan). The light was turned on to initiate photocatalytic degradation. After photodegradation, samples of MB were withdrawn, quenched with methanol and centrifuged at 6000 rpm. The final concentration of MB (Cf) was measured. MB removal was calculated using Equation (1):
MB removal = [(Co − Cf)/Co] × 100
The experimental conditions in Table 1 are derived from BBD in RSM.

2.5. Statistical Analysis

Experimental results were analyzed by response surface methodology (RSM) and one-way ANOVA analysis in Design Expert 13.

3. Results and Discussion

3.1. Material Characterization

Al2O3–MgO was characterized using UV–DRS. Figure 1a shows the plot of absorbance and wavelength for Al2O3–MgO. The maximum absorbance of Al2O3–MgO was 290 nm.
The band gap of Al2O3–MgO was calculated from the Kubelka–Munk function, [F(R)], using the Tauc equation as shown in Equation (2) [10]:
[F(R)hv] (1/γ) = B(hv − Eg)
where h is Planck’s constant, v is the frequency of the photon, B is constant and Eg is the band gap energy of Al2O3–MgO; γ is equal to 2 for indirect band gap transition [10]. From the Kubelka–Munk plot in Figure 1a, the Eg for Al2O3–MgO is 3.50 eV. The Eg for pure Al2O3 is reported to be 5.97 eV [8]. Al2O3–MgO has a lower Eg than pure Al2O3 due to the presence of defects and heterojunction in the Al2O3–MgO nanocomposite [6,11].

3.2. Initial Photocatalyst Tests and Effect of Operational Parameters

Figure 2a shows the photodegradation tests of 11 ppm MB dye with pH, time and photocatalyst dosage as 7, 180 min and 500 mg/L, respectively. Pure Al2O3 has a photodegradation efficiency of 43.57%, whereas Al2O3–MgO has 72.72%. Al2O3–MgO has a lower band gap energy than Al2O3 and hence a higher photocatalytic activity.
The effect of pH is shown in Figure 2b. As pH increased from 3 to 11, MB removal increased from 0% to 57%. In acidic pH, MB exists as a neutral molecule below its pKa value of 3.8 [12]. The point of zero charge (pHzc) of Al2O3–MgO is 10.04, as shown in Figure S2. In acidic conditions, the surface of Al2O3–MgO is positively charged and MB molecules have low adsorption on the catalyst surface, leading to a low photodegradation efficiency. MB dye exists as cations above the pKa value. As the pH approaches the pHzc, the repulsive force against MB cations is reduced [13]. In alkaline conditions, above the pHzc, the surface of Al2O3–MgO is negatively charged and attracts the MB cations. This leads to an uptake of photogenerated holes and electrons, leading to higher photodegradation efficiency.
Figure 2c shows the effect of time on the photodegradation of MB dye. MB removal increased from 54% at 60 min to 58% at 180 min. At constant photocatalyst dosage, more electrons and holes are generated by Al2O3–MgO as irradiation time increases. This leads to the production of more reactive radical species, leading to an increment in the photocatalytic degradation efficiency [14].
The effect of photocatalyst dosage is shown in Figure 2d. MB removal increased from 48% at 200 mg/L to 58% at 1000 mg/L. An increasing Al2O3–MgO dosage leads to the generation of more active sites and subsequently leads to a higher photodegradation efficiency for MB degradation [11]. However, a high Al2O3–MgO dosage causes agglomeration of particles. This results in low photon uptake by Al2O3–MgO and a low generation of electrons and holes, and hence, a decrease in MB photodegradation efficiency [15].
The effect of the initial MB dye concentration is shown in Figure 2e. MB removal was reduced from 57% to 0% as the initial MB dye increased from 10 to 100 ppm. As the concentration of MB dye increases, more MB molecules are adsorbed on the surface of Al2O3–MgO. This limits the influx of photons on the photocatalyst surface, resulting in a low generation of electrons and holes [11]. At constant Al2O3–MgO dosage and light radiation, the number of reactive radicals generated is insufficient for MB photodegradation. Intermediates will also compete with the parent MB molecule for radicals, leading to a lower photodegradation efficiency.
Figure S3 shows the 3D contour plots for the interaction of parameters. Each parameter contributes to achieving MB photodegradation.

3.3. Optimization of Operational Factors3.4. Model Validation and Kinetics

The operating factors can be adjusted by the following equation:
(MB removal + 22.5)0.81 = 12.63 + 1.56A − 0.9753B + 0.4529C − 5.39D + 1.97AB + 0.3677AC − 10.15AD − 14.3BD − 1.01CD − 1.38A2 − 0.0321B2 − 0.0093C2 + 4.86D2

3.4. Model Validation and Kinetics

The optimized conditions of pH, time, photocatalyst dosage and initial MB concentration were 11, 115.6 min, 996.846 mg/L and 10 ppm, respectively, resulting in an MB dye removal of 57.82%, as shown in Figure S4a. The verification experiment at optimized conditions revealed that MB removal was 59.20%, as shown in Figure S4b, with an error of 1.68%. The photodegradation of MB dye with Al2O3–MgO followed first-order kinetics with an R2 of 0.9520 and a k1 of 0.007 min−1, as shown in Figure S5.

3.5. Suggested Removal Mechanism

The mechanism for MB degradation by Al2O3–MgO is shown in Figure S6. It is assumed that MB degradation is due to the attack of generated hydroxyl (OH•) and superoxide (O2−•) radicals. The radicals are generated when photons of light are incident on Al2O3–MgO, leading to the excitation of electrons from the valence band (VB) to the conductance band (CB). This leads to the generation of holes (h+ VB) and electrons (e CB). Hydroxyl ions from the water will then react with the holes to produce OH•. Oxygen present in the MB solution reacts with electrons to produce O2−•. The radicals will react with MB dye molecules, resulting in degradation products, carbon dioxide and water. Equations (4)–(8) summarize the degradation mechanism [10,12]:
Al2O3–MgO + hv (Vis region) → Al2O3–MgO (ecb + hvb+)
hvb+ + H2O(OH-) → OH• + H+
ecb + O2 → O2
O2• + MB → degradation products + CO2 + H2O
OH• + MB → degradation products + CO2 + H2O

3.6. Economic Evaluation

An economic evaluation for the synthesis of Al2O3–MgO and its application in the photodegradation of MB dye was conducted. The textile wastewater treatment plant was assumed to be treating 80 m3 of wastewater per day under optimized conditions. The annual operating time was 300 days. Table S2 summarizes the economic evaluation. The amortization cost (A) was calculated using Equation S5 [15] for a period of 25 years at an interest rate of 6% per annum and was found to be USD 365,765.78. The annual cost (AC) of treating textile wastewater cost per m3 was evaluated using Equation (S6). For 300 working days, the AC was USD 16.50/m3. The revenue generated from selling the Al2O3–MgO nanocomposite was calculated as USD 3.25 for treating 1 m3 of wastewater. The estimated cost after wastewater reuse and pollution reduction was USD 0.46/m3 and USD 0.16/m3, respectively. The payback period was 3.17 years, as shown in text S1.

3.7. Conclusions

An Al2O3–MgO nanocomposite was prepared using the co-precipitation method. UV–DRS analysis showed that the band gap energy was 3.50 eV. Al2O3–MgO nanocomposite was used in the photocatalytic degradation of MB dye. The effect of operating parameters was investigated using RSM. The RSM model had an R2 of 0.9880, indicating a good prediction of MB removal. MB photocatalytic degradation using Al2O3–MgO followed first-order kinetics. Economic estimation revealed that the wastewater treatment cost was USD 16.50/m3 with a payback period of 3.17 years.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ECP2023-14638/s1. Figure S1: Synthesis of Al2O3–MgO; Figure S2: pHzc of Al2O3–MgO; Figure S3: Three-dimensional contour plots for the interaction of parameters; Table S1: ANOVA analysis for quadratic model MB photodegradation; Figure S4: Optimized conditions and the result of validation experiment; Equations (S1)–(S4): zero-, half-, first- and second-order kinetic models; Figure S5: Kinetic models for (a) zero order, (b), half order, (c) first order, (d) second-order; Figure S6: Mechanism of photocatalytic degradation of MB dye by Al2O3–MgO; Table S2: CAPEX and OPEX; Equation (S5): Amortization cost; Equation (S6): Annual cost; Equation (S7): payback period

Author Contributions

Conceptualization, S.D. and A.A.; methodology, S.D.; software, A.A.; validation, S.D.; formal analysis, S.D.; investigation, S.D.; resources, M.N., A.A. and S.O.; data curation, S.D.; writing—original draft preparation, S.D.; writing—review and editing, A.A. and M.N.; visualization, S.D.; supervision, M.N., A.A. and S.O.; project administration, A.A. and M.N. 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 are available upon request from the corresponding author.

Acknowledgments

The authors wish to acknowledge TICAD 7 for the scholarship of the researcher as well as JICA and E-JUST for all equipment used throughout the research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khan, I.; Saeed, K.; Zekker, I.; Zhang, B.; Hendi, A.H.; Ahmad, A.; Ahmad, S.; Zada, N.; Ahmad, H.; Shah, L.A.; et al. Review on Methylene Blue: Its Properties, Uses, Toxicity and Photodegradation. Water 2022, 14, 242. [Google Scholar] [CrossRef]
  2. Hamdy, A.; Mostafa, M.; Nasr, M. Zero-valent iron nanoparticles for methylene blue removal from aqueous solutions and textile wastewater treatment, with cost estimation. Water Sci. Technol. 2018, 78, 367–378. [Google Scholar] [CrossRef] [PubMed]
  3. Safo, K.; Noby, H.; Matatoshi, M.; Naragino, H.; El-Shazly, A.H. Statistical optimization modeling of organic dye photodegradation process using slag nanocomposite. Res. Chem. Intermed. 2022, 48, 4183–4208. [Google Scholar] [CrossRef]
  4. Abdelhaleem, A.; Chu, W. Photodegradation of 4-chlorophenoxyacetic acid under visible LED activated N-doped TiO2 and the mechanism of stepwise rate increment of the reused catalyst. J. Hazard. Mater. 2017, 338, 491–501. [Google Scholar] [CrossRef] [PubMed]
  5. Ateş, S.; Baran, E.; Yazıcı, B. The nanoporous anodic alumina oxide formed by two-step anodization. Thin Solid Film. 2018, 648, 94–102. [Google Scholar] [CrossRef]
  6. Mebed, A.M.; Abd-Elnaiem, A.M.; Alshammari, A.H.; Taha, T.A.; Rashad, M.; Hamad, D. Controlling the Structural Properties and Optical Bandgap of PbO–Al2O3 Nanocomposites for Enhanced Photodegradation of Methylene Blue. Catalysts 2022, 12, 142. [Google Scholar] [CrossRef]
  7. Bin Mokaizh, A.A.; Shariffuddin, J.H.B.H. Manufacturing of Nanoalumina by Recycling of Aluminium Cans Waste. In Waste Recycling Technologies for Nanomaterials Manufacturing; Topics in Mining, Metallurgy and Materials Engineering; Makhlouf, A.S.H., Ali, G.A.M., Eds.; Springer International Publishing: Cham, Switzerland, 2021; pp. 851–870. [Google Scholar] [CrossRef]
  8. Bharthasaradhi, R.; Nehru, L. Structural and phase transition of α-Al2O3powders obtained by co-precipitation method. Phase Transitions 2016, 89, 77–83. [Google Scholar] [CrossRef]
  9. Mantilaka, M.P.G.; De Silva, R.T.; Ratnayake, S.P.; Amaratunga, G.; de Silva, K.N. Photocatalytic activity of electrospun MgO nanofibres: Synthesis, characterization and applications. Mater. Res. Bull. 2018, 99, 204–210. [Google Scholar] [CrossRef]
  10. Makuła, P.; Pacia, M.; Macyk, W. How To Correctly Determine the Band Gap Energy of Modified Semiconductor Photocatalysts Based on UV–Vis Spectra. J. Phys. Chem. Lett. 2018, 9, 6814–6817. [Google Scholar] [CrossRef] [PubMed]
  11. Hassena, H. Photocatalytic Degradation of Methylene Blue by Using Al2O3/Fe2O3 Nano Composite under Visible Light. Mod. Chem. Appl. 2016, 2016, 1000176. [Google Scholar] [CrossRef]
  12. Salazar-Rabago, J.J.; Leyva-Ramos, R.; Rivera-Utrilla, J.; Ocampo-Perez, R.; Cerino-Cordova, F.J. Biosorption mechanism of Methylene Blue from aqueous solution onto White Pine (Pinus durangensis) sawdust: Effect of operating conditions. Sustain. Environ. Res. 2017, 27, 32–40. [Google Scholar] [CrossRef]
  13. Allawi, F.; Juda, A.M.; Radhi, S.W. Photocatalytic degradation of methylene blue over MgO/α-Fe2O3 nano composite prepared by a hydrothermal method. AIP Conf. Proc. 2020, 2290, 030020. [Google Scholar] [CrossRef]
  14. Vinothkumar, P.; Manoharan, C.; Shanmugapriya, B.; Bououdina, M. Effect of reaction time on structural, morphological, optical and photocatalytic properties of copper oxide (CuO) nanostructures. J. Mater. Sci. Mater. Electron. 2019, 30, 6249–6262. [Google Scholar] [CrossRef]
  15. Mensah, K.; Samy, M.; Ezz, H.; Elkady, M.; Shokry, H. Utilization of iron waste from steel industries in persulfate activation for effective degradation of dye solutions. J. Environ. Manag. 2022, 314, 115108. [Google Scholar] [CrossRef] [PubMed]
Figure 1. UV–DRS: (a) absorbance and (b) Kubelka–Munk plots for Al2O3–MgO.
Figure 1. UV–DRS: (a) absorbance and (b) Kubelka–Munk plots for Al2O3–MgO.
Engproc 37 00087 g001
Figure 2. (a) Initial photocatalyst test; effect of (b) pH; (c) time; (d) photocatalyst dosage; (e) initial MB dye concentration on MB photodegradation.
Figure 2. (a) Initial photocatalyst test; effect of (b) pH; (c) time; (d) photocatalyst dosage; (e) initial MB dye concentration on MB photodegradation.
Engproc 37 00087 g002
Table 1. Experimental factors and limits in BBD.
Table 1. Experimental factors and limits in BBD.
FactorParameterUnitLevels
LowMedianHigh
ApH-3711
BTimemin60120180
CPhotocatalyst dosagemg/L2006001000
DInitial MB dye concentrationppm1055100
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

Demarema, S.; Abdelhaleem, A.; Ookawara, S.; Nasr, M. Fabrication of Aluminum-Based Hybrid Nanocomposite for Photocatalytic Degradation of Methylene Blue Dye: A Techno-Economic Approach. Eng. Proc. 2023, 37, 87. https://doi.org/10.3390/ECP2023-14638

AMA Style

Demarema S, Abdelhaleem A, Ookawara S, Nasr M. Fabrication of Aluminum-Based Hybrid Nanocomposite for Photocatalytic Degradation of Methylene Blue Dye: A Techno-Economic Approach. Engineering Proceedings. 2023; 37(1):87. https://doi.org/10.3390/ECP2023-14638

Chicago/Turabian Style

Demarema, Samuel, Amal Abdelhaleem, Shinichi Ookawara, and Mahmoud Nasr. 2023. "Fabrication of Aluminum-Based Hybrid Nanocomposite for Photocatalytic Degradation of Methylene Blue Dye: A Techno-Economic Approach" Engineering Proceedings 37, no. 1: 87. https://doi.org/10.3390/ECP2023-14638

Article Metrics

Back to TopTop