Open access peer-reviewed chapter

Efficacy of Cobalt-Incorporated Mesoporous Silica towards Photodegradation of Azodyes and Its Kinetic Study for Advanced Application

Written By

Krishna Vaddadi, Nookaraju Muralasetti and Naginami Naidu

Reviewed: 04 April 2023 Published: 29 August 2023

DOI: 10.5772/intechopen.111520

From the Edited Volume

Toxicity of Nanoparticles - Recent Advances and New Perspectives

Edited by Mohammed Muzibur Rahman, Jamal Uddin, Abdullah Mohamed Asiri and Md Rezaur Rahman

Chapter metrics overview

47 Chapter Downloads

View Full Metrics

Abstract

Along with MCM-41, cobalt-incorporated mesoporous silica (Co-MCM-41) has been created. Powder X-ray diffraction, scanning electron microscopy, and nitrogen adsorption-desorption studies were used to describe the materials. It has been discovered that the Co-MCM-41 has less surface area (SBET, m2 g−1), pore volume (cc·g−1), and pore size (Å) than the MCM-41. The SEM-EDAX analysis has also unmistakably demonstrated the existence of the appropriate elements in the materials. The photoactivity was significantly impacted by the extremely distributed Co3+ species present on the MCM-41 structure. A theoretical loading of 3.5 wt% permitted an AO7 degradation percentage of about 70% for the samples that were simply treated with Co. Increased Co3+ inactive species, such as clusters or −Co2O3 nanoparticles, are present at higher loadings, but the photoactivity is not noticeably increased. By using the Kubelka-Munk function to the UV-Vis DRS results, it was discovered that the band gap (eV) in the Co-MCM-41 was also substantially smaller than in its parent template. The Alizarin Red S dye was successfully photodegraded employing the materials as photocatalysts, and pseudo-first order kinetics was carried out using the Langmuir-Hinshelwood kinetic model. The necessary experimental setups were all optimised.

Keywords

  • mesoporou silica
  • MCM-41
  • heterogeneous catalyst
  • photocatalyst
  • Alizarin Red S

1. Introduction

A growing focus on heterogeneous catalysts has emerged in recent years due to economic and environmental factors. These catalysts typically have low cost, high reactivity, environmental friendliness, high selectivity, easy setup, and recoverability of catalysts [1]. Mesoporous MCM-41 materials have come to light as highly stable compounds with a significant surface area. They have been widely used as catalysts or catalyst supports in a variety of processes. This was explained by the reaction mixture’s higher reusability and straightforward recoverability [2]. Due to their increased oxidative or acidic character, the introduction of metal ions including Ti4+, Al3+, Co3+, and Fe3+ into the framework has demonstrated improved catalytic activity [3].

Due to contaminated ground water and dangerous industrial effluents, the entire world is currently experiencing environmental issues [4, 5]. These highly coloured effluents harm the environment when they are dumped into water systems because they impede light penetration and hinder aquatic life’s ability to photosynthesize [6, 7]. Intense colour is imparted by the presence of dyes in the effluent at very low concentrations (1 mg L−1) but it is discovered that they are harmful to the environment [8, 9]. In order to effectively remove colour from waste fluids, physical or chemical procedures must be used [10]. The majority of dyes used on an industrial basis are derivatives of azo, anthraquinone, indigo, triphenylmethane, xanthene, and others [11, 12]. Due to their advantageous qualities—bright colour, easy application, and low energy consumption—these dyes are widely utilised in the textile industry. They are, however, typically the most poisonous and mutagenic substances found in nature [13, 14].

One of the anthraquinone class of dyes, Alizarin Red S (ARS), is widely used in the textile, woven fabric, wool, and cotton industries [15, 16, 17]. The paint, plastics, leather, and cosmetics sectors all utilise anionic dye extensively [18]. However, when industrial effluents are released into aquatic environments at amounts exceeding what is permitted, the aquatic life is negatively impacted [19, 20]. It was shown that traditional aerobic digestion techniques were ineffective at degrading these resistant compounds [21]. Photo catalysis has become a green technique for gathering solar energy and degrading organic pollutants due to the global energy crisis and environmental challenges [22, 23, 24, 25]. ARS was selected as the test molecule to undergo photodegradation in the presence of visible light as a result.

MCM-41 and its metal-included derivatives have a wide range of photocatalytic uses, so cobalt metal ion (Co+3) was successfully inserted into the structure of MCM-41. In order to research the photodegradation of ARS under ideal experimental settings, such as effect of photocatalyst, effect of photocatalyst dose, effect of dye concentration, and effect of pH, the materials were characterised and used as photocatalysts. To identify the active species participating in the photodegradation phenomenon, a scavenger experiment was done in addition to these research. To assess how well the outcomes matched each other, the kinetic research was carried out.

Advertisement

2. Results and discussion

2.1 Powder XRD studies

Figure 1 displays the XRD patterns (2°≤2θ ≤ 10°) of the MCM-41 and Co-MCM-41 samples. The patterns only display one low-angle peak for the d100 plane, which corresponds to the mesophase at a value of 2 approximately 2.2o (d-spacing: 32.54623 A, wall thickness: 2.712 A). This is typical of MCM-41’s long-range hexagonal structure [26]. Planes that mirror the characteristics of mesoporous nature as in MCM-41. Co-diffraction MCM-41’s pattern has a lesser intensity of the low angle peak than MCM-41, which suggests that the metal ions are obstructing the structure that directs the template’s action in the materials’ regular ordering.

Figure 1.

Powder XRD patterns of MCM-41 and CoMCM-41.

The diffraction planes d110 (d-spacing = 19.48761 A, wall thickness = 4.531 A), and d200 (16.95948 A, wall thickness = 5.207 A), which reflect the hexagonal array of MCM-41, are responsible for the less intense and broader peaks in the 2 of 4.0°–5.5°. Three diffraction peaks suggest that the mesopores are ordered crystallographically. The reason for the low value of 2 is primarily the template’s long carbon chain, which was employed to synthesise MCM-41 [27]. Co-MCM-41 materials, in contrast, exhibit one large peak at 2 = 2.5o, which corresponds to the mesoporous phase, and two succeeding, less intense peaks at (110) and (200) crystal planes, which mirror the mesoporous characteristic of MCM-41. Co-diffraction MCM-41’s pattern has a lesser intensity of the low angle peak than MCM-41, which suggests that the metal ions are obstructing the structure that directs the template’s action in the materials’ regular ordering.

2.2 Nitrogen adsorption-desorption studies

It was discovered that the synthetic materials followed a standard type-IV adsorption isotherm without hysteresis. This demonstrates how mesoporous these materials are [28, 29]. By using the BET (Brunauer, Emmet, and Teller) method to determine the specific surface area of the materials from adsorption isotherms, it can be demonstrated that the insertion of metal ions reduces the materials’ surface area. This might be explained by metal ions filling part of the pores. By using the BJH (Barrett-Joyner-Halenda) method, the pore size and pore volume of the materials are assessed. Table 1 lists the textural characteristics of MCM-41 and Co-MCM-41 materials.

MaterialSBET (m2g−1)Pore size (Å)Pore volume (cc g−1)
MCM-411023.5017.200.28
CoMCM-41698.2216.900.22

Table 1.

Textural characteristics of the materials.

2.3 SEM-EDAX studies

Figures 2 and 3 show the SEM-EDAX micrographs of MCM-41 and Co-MCM-41, respectively. All of the materials have spherical morphologies similar to those of MCM-41, according to SEM micrographs of the materials. The functionalization of materials with metal ions (Co+3) was also validated by EDAX analysis. It was discovered that adding metal ions to the framework has no effect on the morphology of the materials.

Figure 2.

SEM-EDAX micrographs of MCM-41.

Figure 3.

SEM-EDAX micrographs of Co-MCM-41.

2.4 UV-Vis diffuse reflectance spectra & Kubelka-Munk function curve

The UV-Vis diffuse reflectance spectra were taken in order to comprehend the coordination between the Cobalt and MCM-41. Pure MCM-41 was found to lack a distinctive absorption peak in the 200–800 nm range, which suggests that it was not sensitive in the UV-Visible range [30]. However, a strong absorption peak was seen in Co-MCM-41 in the 400–450 nm (430 nm) range, which is consistent with the octahedral geometry of the Co+3 crystal field transition.

5T2g(D)→5Eg(D) [(t2g4eg2)→(t2g3eg3)]. By using the Kubelka-Munk (KM) function, the produced mesoporous materials have also been described for their band gap values (in electron volts, eV). Figure 4 showed the findings and a plot of the band gap energy values (eV) vs. the modified Kubelka-Munk function [F(R)hv]2 [31]. In MCM-41 and Co-MCM-41 materials, the band gap was discovered to be 2.9 eV and 2.7 eV, respectively. It was clear from the analysis of these results that immobilising Cobalt caused a sizable reduction in the band gap as well as proper coordination of the Cobalt (+3) ion in the zeolite framework. A necessary component of the phenomena of photocatalysis, the photocatalytic activity in the visible light can be improved by the smaller band gap [31].

Figure 4.

KM function curve of MCM-41 and Co-MCM-41 materials.

Adsorption studies: Adsorption isotherm experiment was carried out in the dark to investigate the maximum ARS adsorption capacity on the surface of the MCM-41 and Co-MCM-41 materials. The following statement [32] inevitable factor for the phenomena of photocatalysis [31] was used to determine the amount of ARS absorbed per unit mass of the adsorbent at time, t (min).

qt=CoCtVm.E1

Co and Ct are the dye solution concentrations (mol/L) before and after adsorption, respectively. V is the dye solution’s volume (L) in the photoreactor, and m is the photocatalyst’s mass (g).

Figure 5 shows the amount of ARS adsorption as a function of time. In the photoreactor, it was seen that a fast adsorption developed after 15 min of contact. However, the adsorption propensity started out very mildly and only became saturated after 15 min. It demonstrates that the addition of Cobalt to the MCM-41 framework increased the dye’s tendency to be removed. Based on these findings, the equilibration time was optimised to be 15 min with both mesoporous materials in darkness, and the same was fixed as the equilibration duration for additional research.

Figure 5.

Adsorption studies of ARS with MCM-41 and Co-MCM-41 materials. Catalyst load = 10 mg, [ARS] = 2.92 × 10−5 M, ARS volume = 100 mL.

Effect of photocatalyst: The synthesised materials were used to investigate the impact of deterioration on the 2.92 × 10–5 M aqueous solutions of ARS. The results are depicted in Figures 6 and 7. Co-MCM-41 (10 mg) achieved about 98% degradation efficiency in 90 min, whereas MCM-41 only destroyed 68% of ARS during the same period (Figure 6). This might be because Co-MCM-41 material has a higher capacity for absorbing light due to its smaller pore size and volume, which could boost its activity for efficient photocatalytic degradation.

Figure 6.

Effect of photocatalyst on degradation of ARS with the photocatalysts (each 10 mg). (triangles) MCM-41; (stars) Co-MCM-41.

Figure 7.

Effect of Co-MCM-41 (10 mg) on absorbance of ARS at λ max = 597 nm with respect to time (min).

Studying the effect of Co-MCM-41 on ARS absorbance, it was found that the intensity of absorbance significantly decreased within 90 min, as shown in Figure 7. The dye’s efficient degradation may be to blame for the decrease in ARS absorption.

Effect of photocatalyst dose: Figure 8 shows the results of research on the effect of photocatalyst (Co-MCM-41) on ARS degradation. From 2 mg to 12 mg of catalyst were added to 100 mL of 2.92 × 10–5 M ARS solution. The findings indicated that increasing the catalyst dose increased the degradation efficiency, and that increasing the catalyst dose over 10 mg had no effect on the degradation. Because there are more active sites on the surface of the catalyst as catalyst concentration increases, the rate of dye adsorption also increases. High catalyst doses, however, cause opacity in aqueous solutions and limit visible light penetration while slowing the rate of breakdown. Therefore, a catalyst dose of 10 mg/100 mL of ARS solution was determined to be optimal for the tests.

Figure 8.

Effect of catalyst load on degradation of ARS with Co-MCM-41 (10 mg). (squares) 15 min; (pentagon) 30 min; (left in triangles) 45 min; (right in triangles) 60 min; (stars) 75 min; (diamonds) 90 min.

Effect of dye concentration: By adjusting the dye’s concentration from 1.16 × 10–5 M to 4 × 10–5 M, the impact of initial dye concentration on the photocatalytic degradation of ARS was calculated. According to Figure 9, the percentage degradation efficiency was high at 2.92 × 10–5 M concentration and subsequently declined as concentrations increased. The dye’s rate of adsorption on the active sites of the photocatalyst increases with increasing dye concentration while concurrently decreasing the dye’s adsorption inclination. Additionally, excessive dye concentrations reduce the number of photons and their adsorption on the catalyst’s surface, which ultimately reduces the degradation efficiency [33].

Figure 9.

Effect of dye concentration on degradation of ARS with Co-MCM-41. Catalyst load = 10 mg/100 mL of ARS. (a = 2.92 × 10−5 M, b = 2.33 × 10−5 M, c = 1.75 × 10−5 M, d = 3.5 × 10−5 M, e = 1.16 × 10−5 M, f = 4 × 10−5 M).

Effect of pH: At different pH levels, the effluents discharged by the textile industry include hazardous colours. As a result, the effect of the dye medium’s pH on its degradation was assessed by altering pH values between 2.0 and 12.0 for a constant dye concentration (2.92 × 10–5 M) and catalyst load (10 mg). The breakdown efficiency of ARS increased at pH 4.0 and peaked at pH 6.0, as shown in Figure 10. The effectiveness of the degradation has diminished when pH levels have increased to 12.0. Since it controls the charge on the particle surface, a solution’s pH is a crucial variable in photocatalytic reactions. Therefore, at pH levels below 6.8, the negatively charged dye molecules and surface of Co-MCM-41 may interact chemically strongly and adsorb. At a pH of between 4.0 and 6.0, this situation may have caused ARS dye to breakdown. Repulsion between the negatively charged dye and the catalyst surface may account for the extremely low degradation rate at higher pH levels.

Figure 10.

Effect of pH on degradation of ARS with Co-MCM-41. [Dye] = 2.92 × 10−5 M, catalyst load = 10 mg/100 mL of ARS.

Role of scavengers in the photocatalysis of ARS: Studies on the photocatalytic degradation of ARS were conducted in both the absence and presence of scavengers. Figure 11 illustrates how the dye’s C/Co varied as a function of time (in minutes).

Figure 11.

Variation of C/Co against time (min) in the scavenger experiment. (squares): Benzoquinone; (circles): Isopropyl alcohol; (upward triangles): No scavengers.

Because it prevents the release of a significant amount of superoxide radicals (O2•), benzoquinone limits the degradation of dye (78.5%). The elimination of ARS (91.2%) from its aqueous solution, which primarily regulates the activity of the hydroxyl radicals (OH•), is also slightly constrained by the presence of iso-propyl alcohol [34]. In contrast, when these scavengers were absent, the degradation efficiency was over 98.8%. Superoxide radical and hydroxyl radical were therefore the two main active species in the redox process of the current investigation.

Plausible mechanism for photocatalytic degradation of ARS: The electrons in the photocatalyst’s valence band would excite into the conduction band and generate electron-hole pairs when the dye solution and photocatalyst were exposed to visible light. The degradation efficiency would be decreased as a result of these charged pairs recombining either directly or through the catalyst surface. A good photocatalyst with increased visible light absorption (reduced band gap, Eg) might transport the electron to oxygen, generating the active superoxide radical ion, and therefore limit the tendency to recombine. According to the UV-Vis DRS investigation, Cobalt added MCM-41 has a lowered bang gap of 2.7 eV, which would support the visible light driven photocatalytic activity. The electron can be excited from the valence to conduction band by the photocatalyst when it interacts with visible light and proceeds in the manner described above. In subsequent steps, the superoxide radical ions might produce hydroxyl radicals, which ultimately efficiently degraded the ARS dye (Figure 12).

Figure 12.

Photocatalytic degradation mechanism of ARS using Co-MCM-41.

Kinetic analysis: The kinetic study was performed for the degradation of ARS dye using Langmuir-Hinshelwood kinetic model [35].

r=dC/dt=kKC/1+KCE2

On neglecting the value of KC in the denominator (KC < <1) and integrating with respect to time t, the above equation accords to the pseudo-first order equation.

lnCoC=kKt=kapptE3

where Co denotes the starting concentration and C denotes the current concentration of the ARS solution. K is the adsorption coefficient of the ARS dye onto the photocatalyst, and t. k is the rate constant.

For the rate of degradation using MCM-41 and Co-MCM-41, respectively, the rate constants (k) were estimated as 1.526 × 10–2 min−1 (2.543 × 10–4 s−1) and 9.20 × 10–2 min−1 (15.33 × 10–4 s−1). Compared to the MCM-41, the rate constant has grown six times with the Co-MCM-41. It demonstrates how adding Cobalt to MCM-41 increases the rate of reaction.

Advertisement

3. Experimental

Materials: Cetyltrimethylammonium bromide [CH3-(CH2)15-N(CH3)3Br, CTMAB], tetraethylorthosilicate [(C2H5O)4Si, TEOS], Cobalt acetate [Co(CH3COO)2.2H2O], Ethyl alcohol, were procured with AR grade quality (99% pure) from Sigma Aldrich and Merck. Alizarin Red S (Merck) with molecular formula, C14H7NaO7S (M.Wt = 342.26 g.mol−1, λ max = 597 nm). Scavenger experiments were carried out using benzoquinone and iso-propyl alcohol. None of the compounds were further purified before usage. Utilising double distilled water, all solutions needed for the experimental work were created.

Synthesis of MCM-41: MCM-41 A straightforward room-temperature co-precipitation technique was used to create the material [26]. In a typical synthesis, a clear, homogenous solution was created by dissolving 2.40 g of the surfactant CTMAB in 50.00 mL of distilled water and stirring continuously. This homogenous solution received 76.00 mL of ethyl alcohol and 13.00 mL of 25 weight percent aqueous ammonia while being stirred. The aforementioned mixture received a dropwise addition of 10.00 mL of TEOS. The solution turned milky and a gel was formed due to the hydrolysis of TEOS. The resultant mixture was stirred for about 2 h to completely hydrolyze TEOS. White precipitate thus formed was centrifuged, washed consecutively with distilled water and methanol. The product was dried overnight at 110°C. Solid product thus obtained was calcined at 550°C in air atmosphere for 5 h to remove the trapped surfactant.

Synthesis of Cobalt incorporated MCM-41 (Co-MCM-41): The in-situ approach was used to create MCM-41 with metal incorporation. These were made using the same procedure as MCM-41, with the exception that the addition of the appropriate amounts (Si/M ion ratio of 100:1, M = metal) of the corresponding metal precursors was made 20 min after the addition of TEOS. Using cobalt acetate as a precursor material, cobalt (Co) integrated MCM-41 materials were created.

Characterisation: Step scanning on the Rigaku D/MAX-2500 diffractometer (Rigaku Co., Japan) with Cu-K -radiation (k = 0.15406 nm), operated at 40 kV and 100 mA, was used to analyse the produced MCM-41 and its Cobalt doped derivative. A scanning electron microscope model Philips XL 30 ESEM was used to capture SEM images of the materials (FEI-Philips Company, Hillsboro). The surface areas of the catalysts were measured using a Quantachrome Nova 2000e surface area and pore size analyser by nitrogen adsorption-desorption in a liquid nitrogen atmosphere (77 K). With the help of the Single Monochromator UV-2600 (optional ISR-2600Plus, up to 1400 nm), UV-Vis diffuse reflectance spectra were captured.

Photocatalytic measurements: The efficiency of the photocatalysts in the photodegradation of ARS was examined using a UV-Vis Spectrophotometer (UV-2550, Shimadzu, wavelength range: 180–1100 nm). With a known weight (10 mg) of the photocatalyst per 100 mL of a variety of diluted ARS solutions, adsorption-desorption equilibrium tests were carried out. The chosen equilibration period was 15 min. The solutions were exposed to a 300 watt tungsten halogen lamp in a photocatalytic chamber with an electrical supply to conduct the photocatalytic study. A UV-Vis spectrophotometer was used to evaluate the translucent solution after centrifuging 5 mL aliquots at regular intervals of 5 min. Eq. (4) was used to get the percentage of dye degradation.

Photocatalytic degradation%=CoCt/Co×100E4

where C0 and Ct are the initial concentration and concentration of the dilute ARS solution at a time interval, t respectively.

Scavenger experiment: The purpose of the in-situ scavenger experiment was to look for active species produced by the efficient photocatalyst during the photocatalytic breakdown of ARS. Scavengers were added to the photocatalytic process to capture holes, or superoxide radicals (O2•) and hydroxyl radicals (OH•), such as benzoquinone (1 mmol/L) and iso-propyl alcohol (0.1 mmol/L).

Advertisement

4. Applications

There are many benefits to using this Co-MCM-41 heterogeneous catalyst in sophisticated oxidation processes [36]. Co-MCM-41 mesoporous materials have been synthesised, and their high specific surface area, well-ordered mesoporous structure, large surface area, and flexible framework make it easy to incorporate metal active sites in silica materials [37, 38]. These materials also make good dispersions for reaction and product molecule molecules. The host material’s adsorption characteristics are crucial for the purification of biomolecules. According to a publication, MCM-41 with Co inserted is a strong contender for use as a high surface area adsorbent for biological compounds like amino acids [39]. This has undergone testing as a method of regulated drug release for medications like ibuprofen [40]. These heterogeneous catalyst Co-MCM-41 based mesoporous materials have successfully immobilised or absorbed a number of proteins, including cytochrome C [41], lysozym [42], and trypsin [43], and their activities have been investigated [44]. Well-ordered mesoporous silica nanoparticles have recently been demonstrated to be useful as cell markers by Lin et al. [45].

Advertisement

5. Conclusions

The derivative of MCM-41 with cobalt was created and described. Under ideal circumstances, the materials have produced effective results in the photodegradation of the anthroquinone dye Alizarin Red S. The fact that Co-MCM-41 degrades more effectively than MCM-41 is proof of the metal’s influence on the microporous material’s framework when serving as a photocatalyst.

Advertisement

Acknowledgments

The author is thankful to Sophisticated analytical instrumentation facility (SAIF), India and National Institute of Technology, Warangal (NIT-W), India for providing the characterisation results. Also thankful to the Management of NS Raju Institute of Technology, Visakhapatnam, India for providing the funding and laboratory facilities.

References

  1. 1. Dodd J, Schwender CF, Moore JB Jr., Ritchie DM, Gray-Nunez Y, Loughney D, et al. Design and discovery of RWJ 22108--a novel bronchoselective calcium channel blocker. Drug Design Discovery. 1998;15:3
  2. 2. Vicini P, Geronikaki A, Incerti M, Busonera B, Poni G, Kabras CA, et al. Synthesis and Biological Evaluation of Benzo[d]isothiazole, Benzothiazole and Thiazole Schiff Bases. Bioorganic and Medicinal Chemistry. 2003;11:4785
  3. 3. Bergbreiter DE, Newcomb M. Asymmetric Synthesis. In: Morrison JD, editor. Vol. 2A. Academic Press; p. 243
  4. 4. Rubin AJ. Aqueous-Environmental Chemistry of Metals. Ann Arbor , MI, USA: Science Publishers; 1974
  5. 5. Mercier L, Pinnavaia TJ. Heavy Metal Ion Adsorbents Formed by the Grafting of a Thiol Functionality to Mesoporous Silica Molecular Sieves: Factors Affecting Hg(II) Uptake, Environmental Science & Technology. 1998;32:2749
  6. 6. Clarke EA, Anliker R. Organic dyes and pigments. In: Handbook of Environmental Chemistry, Anthropogenic Compounds. Vol. 3. New York: Springer-Verlag; 1980
  7. 7. Robinson T, Mcmullan G, Marchant R, Nigam P. Remediation of dyes in textile effluent: a critical review on current treatment technologies with a proposed alternative. Bioresource Technology. 2001;77:247
  8. 8. Nigam P, Armour G, Banat IM, Singh D, Marchant R. Physical removal of textile dyes from effluents and solid-state fermentation of dye-adsorbed agricultural residues. Bioresource Technology. 2000;72:219
  9. 9. Naeem A, Abdul H, Safia A. Physicochemical characterization and Bioremediation perspective of textile effluent, dyes and metals by indigenous Bacteria.Journal of Hazardous Materials. 2009;164:322
  10. 10. Sharma S, Kaur A. Various Methods for Removal of Dyes from Industrial Effluents - A Review. Indian Journal of Science and Technology. 2018;11(12):1
  11. 11. Fu Y, Viraraghavan T. Removal of Congo Red from an aqueous solution by fungus Aspergillus niger. Advanced Environmental Research. 2002;7:239
  12. 12. Baptista MS, Indig GL. Effect of BSA Binding on Photophysical and Photochemical Properties of Triarylmethane Dyes. The Journal of Physical Chemistry B. 1998;102:4678
  13. 13. Wu J, Eitman MA, Law SE. Evaluation of membrane filtration and ozonation processes for treatment of reactive-dye wastewater. Journal of Environmental Engineering. 1998;12:272
  14. 14. Ozdemir O, Armagan B, Turan M, Celik MS. Comparison of the adsorption characteristics of azo-reactive dyes on mesoporous minerals. Dyes and Pigments. 2004;62:49
  15. 15. Parra S, Sarria V, Moloto S, Periner P, Pulgarin C. Photocatalytic degradation of atrazine using suspended and supported TiO2. Applied Catalysis. B, Environmental. 2004;51:107
  16. 16. Talib TH, AlDamen MA, Alani RR. Modeling of Advanced Photo Oxidation of Alizarin Red-S Dye Using Tio2 as Photo Catalyst. American Chemical Science Journal. 2014;4(6):918
  17. 17. Kaur S, Sharma S. Sushil Kumar Kansal. Synthesis of ZnS/CQDs nanocomposite and its application as a photocatalyst for the degradation of an anionic dye, ARS. Superlattices and Microstructures. 2016;98:86
  18. 18. Navas CS, Reboredo MM, Granados DL. Comparative Study of Agroindustrial Wastes for their use in Polymer Matrix Composites. Procedia Materials Science. 2015;8:778
  19. 19. Schiavello M. Heterogeneous Photocatalysis. England: John Wiley & Sons; 1997
  20. 20. Kansal SK, Lamba R, Mehta SK, Umar A. Photocatalytic degradation of Alizarin Red S using simply synthesized ZnO nanoparticles. Materials Letters. 2013;106:385
  21. 21. Kansal SK, Kaur N, Singh S. Photocatalytic Degradation of Two Commercial Reactive Dyes in Aqueous Phase Using Nanophotocatalysts. Nanoscale Research Letters. 2009;4:709
  22. 22. Hwang SD, Kim SW, Lee WK, Kim EJ, Hahn SH. Visible Light Photocatalytic Activity of N-Ion Implanted TiO2 Thin Films Prepared by Oblique Incident Electron-Beam Evaporation Method. Journal of Nanoscience and Nanotechnology. 2013;13:7059
  23. 23. Corma A. From Microporous to Mesoporous Molecular Sieve Materials and Their Use in Catalysis. Chemical Reviews. 1997;97:2373
  24. 24. Ciesla U, Schuth F. Ordered mesoporous materials. Microporous and Mesoporous Materials. 1999;27:131
  25. 25. Kraushaar B, Hooff JHCV. A test reaction for titanium silicalite catalysts. Catalysis Letters. 1989;2:43
  26. 26. Rajini A, Nookaraju M, Reddy IAK, Venkatathri N. Vanadium dodecylamino phosphate: A novel efficient catalyst for synthesis of polyhydroquinolines. Chemical Papers. 2013;68(2):170
  27. 27. Rouquerol F, Rouquerol J, Sing K. Adsorption by Powders & Porous Solids. San Diego: Academic Press; 1999
  28. 28. Ferrini C, Kowenhoven HW. Modified Zeolites for Oxidation Reactions. Studies in Surface Science and Catalysis. 1990;55:53
  29. 29. Petrini G, Cesana A, De Alberti G, Geroni F, Leofanti G, Padovan M, et al. Selective oxidations with hydrogen peroxide and titanium silicalite catalyst. Studies in Surface Science and Catalysis. 1991;68:761
  30. 30. Shen S, Chen J, Koodali RT, Hu Y,Xiao Q, Zhou J, et al. Activation of MCM-41 mesoporous silica by transition-metal incorporation for photocatalytic hydrogen production. Applied Catalysis B. 2014;150:138
  31. 31. Addamo M, Augugliaro V, Paola AD,Garcia-Lopez E, Loddo V, Marci G, et al. Preparation and photoactivity of nanostructured TiO2 particles obtained by hydrolysis of TiCl4. Colloid Surface A. Physciochemical EngineeringAspects. 2005;265:23
  32. 32. Sekhar RS, Douglas SP. Graphene oxide-nano-titania composites for efficient photocatalytic degradation of indigo carmine. Journal of the Chinese Chemical Society. 2018;65:1423
  33. 33. Blasco T, Corma A, Navarro MT, Pariente JP. Synthesis, characterization, and catalytic activity of Ti-MCM-41 structures. Journal of Catalysis. 1995;156:65
  34. 34. Liu T, Wang L, Lu X, Fan J, Cai X, Gao B, et al. Comparative study of the photocatalytic performance for the degradation of different dyes by ZnIn2S4: adsorption, active species, and pathways. RSC Advances. 2017;22(7):12292
  35. 35. Sivakumar S, Ranga Rao V, Nageswara Rao G. Efficient Photocatalytic Degradation of Alizarin Red S by Silver-Impregnated Zinc Oxide. Proceedings of the National Academy Science. 2013;83:309
  36. 36. Anipsitakis GP, Stathatos E, Dionysiou DD. Heterogeneous activation of oxone using Co3O4. The Journal of Physical Chemistry. B. 2005;109:13052-13055
  37. 37. Pradhan AC, Paul A, Rao GR. Sol-gel-cum-hydrothermal synthesis of mesoporous Co-Fe@Al2O3−MCM-41 for methylene blue remediation. Chemical Science. 2017;129:381-395
  38. 38. Zhao Q, Zhou X, Ji M, Ding H, Jiang T, Li C, et al. Stability and textural properties of cobalt incorporated MCM-48 mesoporous molecular sieve. Applied Surface Science. 2011;257:2436-2442
  39. 39. O’Connor AJ, Hokura A, Kisler JM, Shimazu S, Stevens GW, Komatsu Y. Amino acid adsorption onto mesoporous silica molecular sieves. Separation and Purification Technology. 2006;48:197
  40. 40. Vallet-Regi M, Ramila A, Del Real RP, Perez-Pariente J. A new property of MCM-41: Drug delivery system. Chemistry of Materials. 2001;13:308
  41. 41. Muñoz B, Rámila A, Pérez-Pariente J, Dýáz I, Vallet-Regý M. MCM-41 organic modification as drug delivery rate regulator. Chemistry of Materials. 2003;15:500
  42. 42. Gimon-Kinsel ME, Jimenez VL, Washmon L. Mesoporous molecular sieve immobilized enzymes. Studies in Surface Science and Catalysis. 1998;117:373
  43. 43. Vinu A, Murugesan V, Hartmann M. Adsorption of lysozyme over mesoporous molecular sieves MCM-41 and SBA-15: Influence of pH and aluminum incorporation. The Journal of Physical Chemistry. B. 2004;108:7323
  44. 44. Fu F-M, DeOliveira DB, Trumble WR, Sarkar HK, Singh BR. Secondary structure estimation of proteins using the amide-III region of Fourier-transform infrared-spectroscopy – Application to analyze calcium binding-induced structural changes in casequestrin. Applied Spectroscopy. 1994;48(11):1432
  45. 45. Tope AM, Srivinas N, Kulkarni SJ, Jamil K. Mesoporous molecular sieve (MCM-41) as support material for microbial cell immobilization and transformation of 2,4,6-trinitrotoluene (TNT): A novel system for whole cell immobilization. Journal of Molecular Catalysis B. 2001;16:17

Written By

Krishna Vaddadi, Nookaraju Muralasetti and Naginami Naidu

Reviewed: 04 April 2023 Published: 29 August 2023