Next Article in Journal
Carbon Accounting of Weihe CSA Pilot Demonstration Area in Longjiang Forest Industry
Previous Article in Journal
The Application of Hyperspectral Images in the Classification of Fresh Leaves’ Maturity for Flue-Curing Tobacco
Previous Article in Special Issue
Magnetic Iron Nanoparticles: Synthesis, Surface Enhancements, and Biological Challenges
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Titanium(III) Oxide Doped with meta-Aminophenol Formaldehyde Magnetic Microspheres: Enhancing Dye Adsorption toward Methyl Violet

by
Suriyan Radha
1,*,
Paul Christygnanatheeba
1,
Karuppiah Nagaraj
2,
Saradh Prasad
3,
Mohamad Saleh AlSalhi
3,
Jeyaraj Vinoth Kumar
4,
Prabhakarn Arunachalam
5 and
Chelladurai Karuppiah
6,*
1
Department of Chemistry, Saiva Bhanu Kshatriya College, Aruppukkottai 626101, Tamil Nadu, India
2
SRICT-Institute of Science & Research, Department of Chemistry, UPL University of Sustainable Technology, Block No. 402, Valia Rd., Vataria, Ankleshwar 393135, Gujarat, India
3
Department of Physics and Astronomy, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
4
Nano Inspired Laboratory, School of Integrated Technology, Yonsei University (International Campus), Incheon 21983, Republic of Korea
5
Chemistry Department, College of Science, King Saud University, P.O. Box 2455, Riyadh 11451, Saudi Arabia
6
Battery Research Center of Green Energy, Ming Chi University of Technology, New Taipei City 24301, Taiwan
*
Authors to whom correspondence should be addressed.
Processes 2023, 11(4), 1250; https://doi.org/10.3390/pr11041250
Submission received: 27 January 2023 / Revised: 4 April 2023 / Accepted: 10 April 2023 / Published: 18 April 2023
(This article belongs to the Special Issue Magnetic Materials for Environmental and Biomedical Applications)

Abstract

:
The demand to synthesize economical detoxification adsorbents of organic pollutants has been a thriving solicitude for most environmental research aspirants. Here, we synthesized a titanium(III) oxide doped with spherical shaped meta-aminophenol formaldehyde magnetic microspheres (Ti2O3/mAPF MMSs) by the polymerization method of Ti2O3 nanoparticles with formaldehyde and m-aminophenol. SEM analysis confirmed the synthesized material as crystalline in nature and had ~400–450 nm sized particles. The physical characterization of the Ti2O3/mAPF MMSs were quantitatively revealed by FTIR spectrum and PXRD in elaboration. The carboxylate frequency and the characteristic apex of the titanium–oxygen bond was found in the FTIR spectrum for Ti2O3/mAPF derived from Ti2O3. The PXRD patterns proved that the synthesized magnetic microspheres contained Ti2O3 nanoparticles. The experimental methods of TGA and DTA confirmed the thermal stability and its composition of Ti2O3/mAPF MMSs. The kinetic adsorption study for methyl violet was confirmed as first-order kinetics. The present study was to investigate the dye desorption of methyl violet from simulated water samples by using a titanium(III) oxide doped with meta-aminophenol formaldehyde magnetic microspheres in an adsorption process.

1. Introduction

The progress of rapid industrialization may bring contaminants into aqueous environments that affect the ecosystem. Methyl violet (MV), cationic dye methylene blue (MB), and the anionic dye methyl orange (MO) are three typical organic dyes in wastewater that mainly come from the food and drug industries, and result in negative effects [1,2,3]. Therefore, it is urgent to find strategies for the fast and efficient removal of organic dyes. Several methods have been reported such as adsorbent materials and chemical catalysis that treat industrial wastewater, but there are some problems in the most common method such as a slow process, secondary pollutants, and regeneration costs, etc. [4]. To solve this problem, many novel photo catalyst methods of dye removal have been created, and they have played an important role in the removal of organic dyes [5,6]. In the recent past, the catalytic potency of coinage metals at the nanoscale has attracted much attention, and leads to weak catalysis and low cycling properties in cyclic performance due to strong aggregation. Hence, we developed a suitable method to prevent such aggregation and to increase the catalytic activity of coinage metal nanoparticles. Based on this requirement, titanium oxide nanoparticles and titanium oxide based magnetic microspheres on various solid carriers are becoming more popular due to advantages such as dye degradation without producing secondary pollutants, and the degraded products are eco-friendly with biological systems. Additionally, titanium oxide based magnetic microspheres and large numbers of reports such as MOFs and PET have been reported on their carriers to enhance their dispersion, catalytic activity, and cycling stability.
Titanium oxide based magnetic microspheres have been attracting much attention due to their dye adsorption property. Synthetic cationic and anionic dyes like methyl violet and EBT substances are usually colored and are difficult to biodegrade because of their mosaic and synthetic structures of an aromatic compound [7,8,9,10,11]. The production of a commercial dye is a major source of the large number of effluent deterioration due to the byproduct of industrial pollution [12,13]. Moreover, the defilement leads to mutagenic and carcinogenic effects because of both the surface and groundwater hazards, which are particularly used in anionic dyes [3,14]. Anionic dyes based on a metal complex and reactive dyes are problematic to degrade because they can resist heat and microbial attack [15]. Therefore, methyl violet dyes need to be eliminated from industrial effluent by various treatment technologies such as advanced oxidation processes (AOPs), solvent extraction, precipitation, and photo catalytic degradation [16,17,18,19,20]. Due to the use of industrial organic dyes to supply the constant growth in people’s living standards, they have become of high concern because their byproducts lead to serious human health issues because of their highly toxic products [21,22,23,24]. An adsorption technique is easy to handle; in particular, low cost resin-based phenolic materials and agricultural wastes, especially resin-based phenolic materials, have very good properties such as low toxicity, outstanding biocompatibility, and porosity [23,24,25,26,27,28,29]. They have a well-to-do array of patterns (e.g., graphitized flakes and nanofibers) [29,30,31,32,33]. In all of these, colloidal resin microspheres are gaining significant attention due to a variety of applications in particle templates, colloidal assemblies, biomedical fields, and optical sensing [9,28,29,30,31,32].
The interesting new catalytic, electronic, optical, and biocompatible properties of colloidal resin microsphere-based nanocomposites have moved forceful research over the past decades. In particular, amine group-attached magnetic particles have a wide variety of applications in drug delivery, catalysis, and enzyme immobilization [34,35,36,37]. In the last two decades, much attention has focused on various phenolic-formaldehyde resins such as 3-aminophenol-formaldehyde resin and resorcinol-formaldehyde resin because phenolic resins exhibit a strong stickiness due to their mussel-like properties [38,39,40,41,42]. For instance, the catalytic effect of resorcinol-formaldehyde-hybridized nanogold for MB [43,44,45,46] and the development of phenolic-formaldehyde resin nanomaterials are of great scientific significance and practical application value. Recently, magnetic core-shell composites have attracted a great deal of attention for their good magnetic responsibility and can be easily magnetized. These microspheres are an essential catalytic property in the advancement of photocatalysis, batteries, and gas storage, which tune the material properties. A microsphere magnetic particle [47,48,49,50] is a fascinating design to use as a microsphere material to enhance the nanoparticles’ stability, dispersion as well as protect microsphere materials and develop microsphere magnetic materials [51,52,53,54].
Over the past few years, the research has focused on phenolic-formaldehyde resin (PF resin)- and resorcinol-formaldehyde resin (RF resin)-based magnetic microspheres due to the progressive development of implications and the applications of magnetic microspheres. The synthesis of various types of PF resins such as phenoplasts has been progressively reported due to their mussel-like properties, [55,56]. For instance, Gavina et al. [57] studied the catalytic effect of RF resin-nano gold hybridized for methyl blue. The significant advancement of RF resins as a nanomaterial and added a value of practical application. Furthermore, mussel chemistry involving catechol groups can adhere to a variety of organic and inorganic composites with various structures that can be fabricated by combining organic and inorganic materials with phenolic-formaldehyde resins. Methyl violet is an eminent organic dye commonly used by industries concerned in the rubber, pharmaceutical, cosmetics, textile, plastics, leather, paper, pharmaceutical, and food industries. Colored wastewater lacking correct action can profusely cause trouble such as the chemical oxygen demand in the water body and expansion in toxicity [58,59,60]. Dyes have mainly been reported to be directly connected to human diseases such as hyperbilirubinemia, amyloidosis, and anemia [61,62,63]. Nowadays, the most extensively used methods for the elimination of colored dyes from dye-rich wastewater are physicochemical (e.g., photo catalytic degradation, ultra-filtration, and physical adsorption on activated carbon). In recent years, activated carbon prepared from agricultural wastes/byproducts have been used as adsorbents for the removal of dyes as an alternative to commercial activated carbon. These methods are attractive given their high effectiveness, but are difficult and exclusive [64,65,66,67,68,69,70,71,72]. A recent study investigated nanocomposites consisting of a magnetite core and an outer Ag-decorated anatase shell and their application for the visible-light photodegradation of cationic and anionic dyes [73,74,75,76]. As research develops, more high attainment MNP hybrids may stimulate various new fields of environmental monitoring, bio-catalysis, and bio-detection applications [77,78,79,80].
In the present study, we report on the preparation of titanium(III) oxide doped with spherical shaped meta-aminophenol formaldehyde magnetic microspheres (Ti2O3/mAPF MMSs) by in situ polymerization in an aqueous solution. We believe that the present results of Ti2O3/mAPF MMSs is an inexpensive and eco-friendly option and can be used as an effective adsorbent for the removal of methyl violet from an industrial effluent. These microspheres will efficiently take out methyl violet. The as prepared Ti2O3/mAPF MMSs had large clusters of NH2 groups over the surface, allowing for its unique potential and applicability.

2. Experimental Method

2.1. Materials

meta-Aminophenol (C6H7NO), methylene oxide (CH2O), titanium tetrabutoxide (Ti(C4H9O)4), polyethylene oxide 600 (H(OCH2CH2)nOH), aqueous ammonia (NH4OH), methyl violet (C24H28N3Cl), and ethanol (C2H5OH) (≥99.8%) were purchased from Sigma Aldrich, St. Louis, MO, USA.

2.2. Preparation of Titanium(III) Oxide Nanoparticles

Prior to synthesis, Ti(C4H9O)4 (6 mL) was poured with 1 mL of anhydrous C2H6O in a round bottom flask (RBF) and maintained in a water bath (thermostat-controlled) at 50 °C. Then, 2.5 mL of PEG-600 was added drop by drop in the direction of the RBF and the mixture was stirred for 3 h to obtain a brightness sol. Finally, the brightness sol was dried with the help of a vacuum oven at 120 °C for 24 h to obtain a gel (yellow color). The hybrid polyethylene oxide 600-based gel was in a porcelain boat, maintained at RT (high pure argon flow) for 30 min, and in a quartz tube in a controlled atmosphere furnace (heated at 5 °C min−1) at 1200 or 1500 °C for 6 h. Finally, reduced black oxides were ground into regular powders [81].

2.3. Synthesis of Ti2O3/mAPF MMSs

The meta-aminophenol-formaldehyde-based microspheres (mAPF) were prepared with a slight modification as per the previous literature [82]. In this synthetic procedure, 0.33 g of meta-aminophenol was added into 52 mL of C2H5OH and 20 mL of H2O, afterward, 224 μL of an aqueous solution (NH4OH,) of NH3 was added to form a consistent solution for 30 min at 40 °C after the addition of 180 μL of the CH2O solution. The obtained solution was stirred at 60 °C for 300 min and then heated at 120 °C for 24 h (Teflon-lined autoclave, Shilpa Enterprises, Bengaluru, India). The resin spheres were purified with double-distilled (DD) water and C2H5OH by centrifugation at a speed of 6000 rpm. The synthesized precipitates (mAPF microspheres) and Ti2O3 nanoparticles were ultrasonicated in 10 mL of DD H2O for 3 h.

3. Results and Discussion

Spherical shaped Ti2O3/mAPF MMSs were synthesized by a sol–gel polycondensation method in dihydrogen monoxide. The measured procedure of Ti2O3/mAPF MMSs and the mechanism for the formation of meta-aminophenol-formaldehyde microparticles is shown in Scheme 1.

3.1. Characterization

The as-synthesized mAPFR MMSs were examined by scanning electron microscopy (SEM). Figure 1A shows that the SEM images of the as synthesized MMSs (a) and Ti2O3/mAPF MMSs (b) contained a number of MMSs. The magnified SEM image of these MMSs demonstrates that the mAPFR MMSs were in the diameter of ca. ~400–450 nm. Additionally, Figure 1A(a) reveals that the MMSs had an even spherical shape and that particles did not accumulate. The SEM image of Ti2O3 decorated on the mAPFR MMSs is shown in Figure 1A(b).
Figure 1B shows the PXRD studies of the (a) mAPF MMS and (b) Ti2O3/mAPF MMS samples. Typically, the PXRD pattern for the mAPF MMSs shows a broad line parallel to the amorphous nature of mAPF, which appeared at a 2θ value of 20–30°, and the Ti2O3 diffraction patterns indexed to (2θ = 3.73 (012), 2.70 (104), 2.57 (110), 2.27 (006), 2.24 (113), 2.12 (202), 1.86 (024), 1.70 (116), 1.50 (122), 1.50 (214), 1.48 (125), 1.30 (220), 1.29 (036) and 1.24 (312) are practical in the 10–80° 2θ range. All the peaks were referenced according to the JCPDS cards (JCPDS # 01-071-0281). The occurrence of all peak intensities according to the JCPDS cards suggests a rhombohedral (R-3c space group) structure. The PXRD patterns proved that the synthesized magnetic microspheres contained a Ti2O3 stable phase and remained unchanged during the synthetic methodology. After doping the Ti2O3 nanoparticles with mAPF, the intensities of the PXRD diffraction lines of mAPF decreased [82]. Using Debye Scherrer’s equation D = Kλ/βhkl cos θ, the crystalline size of the Ti2O3 was around ~100 nm.
Figure 2 shows the FTIR spectrum of the pristine mAPF and Ti2O3 compared to the Ti2O3/mAPF MMSs. As seen from Figure 2a, the peak at 1289 cm−1 formed from the aromatic-oxygen–carbon stretching; at the same time, the broad band at 1379 cm−1 can be associated with the carbon–nitrogen from the NH2 group [83]. The arrival of a low and broad peak at 3372 cm−1 and a strong and narrow peak at 1613 cm−1 were accredited to the occurrence of the –OH and NH2 groups, respectively. The appearance of a peak around 1500 cm−1 was attributed to the carbon–hydrogen stretching, considering that the absorption band around 2900 cm−1 was because of the methylene asymmetric stretching vibration. For the Ti2O3 magnetic nanoparticles (Figure 2b), the peak that appeared at 526 cm−1 was responsible for the titanium–oxygen bonds. The symmetric (1578 cm−1) and asymmetric vibrations (1634 cm−1) of the carboxylate groups derived from ligands on the surface Ti2O3. Figure 2c reveals the spectra of the Ti2O3/mAPF MMS composites. The carboxylate groups appeared at 1500 cm−1, and the respective apex of the titanium–oxygen bond that appeared around 476, 534 cm−1 in the FTIR of Ti2O3/mAPF was adopted from Ti2O3. The rest of the bands originated from mAPF along with the stretching vibrations at 3252 cm−1 (oxygen–hydrogen or –OH…nitrogen), 3366 cm−1 (N-H or NH…O), Ar–H vibration (1212 cm−1), and the respective aromatic rings bands around 1517 cm−1 [84,85,86,87,88].
Thermogravimetric analysis (TGA) of the meta-aminophenol-formaldehyde-based microspheres was carried out using a Mettler Toledo, TGA/DSC 1 series apparatus and conducted at a heating rate of 10 °C min−1 under a constant flow of nitrogen to determine the calcination temperature. The meta-aminophenol-formaldehyde-based microspheres (mAPF) were investigated using an X-ray diffractometer (Desktop X-ray) and recorded using CuKα (λ = 1.54056 Å) radiation operated at 40 kV and 30 mA. The mAPF was characterized and recorded using Fourier transform infrared spectroscopy (FTIR, Bruker IFS28) to analyze and detect the functional group at the scanning region of 4000–400 cm−1. Figure 3 shows the composition and thermal stability of the Ti2O3/mAPF MMSs supplementally described by a thermogravimetric analyzer. The sample was heated from room temperature to 900 °C. The sample weights of the Ti2O3 and Ti2O3/mAPF MMSs were 13.6 and 15.2 mg, respectively. The Ti2O3/mAPF MMS decayed in two stages (Figure 3, blue line). The initial weight loss to 100 °C was equivalent to the evaporation of the solvent and dihydrogen monoxide. Second, 39% weight loss occurred upon increasing the heat from 245 to 577 °C. During this time, the mAPF deliberately abstracted due to the eradication of sundry compounds along with water, H2, CH4, NH3, carbon dioxide, carbon monoxide, ethane, and some truncated hydrocarbons.
The obtained Ti2O3/mAPF MMSs were predicted to absorb organic dyes displaying NH2 and OH groups, since they have distinct functional groups with similar surface polarity (e.g., NH2, OH, OH-CH3, etc.). In the present report, methyl violet was removed to analyze the adsorption potential of Ti2O3/mAPF MMSs. As shown in Figure 4A, the color of methyl violet disappeared after the integration of Ti2O3/mAPF MMSs because of the adsorption of methyl violet onto the Ti2O3/mAPF MMSs. Figure 4A shows the UV–Vis spectra of the methyl violet adsorption by the Ti2O3/mAPF MMSs at various time (0, 6, 12, 18, 24, 30, 36, 42, 48, and 60 min) intervals. Methyl violet was adsorbed on the surface of the Ti2O3/mAPF MMSs via conjugated л bonding and hydrogen bonding.

3.2. Adsorption Isotherm

Adsorption isotherms are of extensive research importance to primary methods of adsorption techniques that are used to predict the adsorption capacity of a given material (e.g., the Langmuir and Freundlich isotherms). Adsorption isotherms can be employed to resolve the greatest use of adsorbents and to design a system of adsorption for the removal of a dye from its solution, in order to provide an appropriate correlation for the equilibrium curve [89]. The interrelationship of Langmuir and Freundlich isotherms are shown in Table 1. The value of RL indicates the shape of the isotherms to be either unfavorable (RL > 1), linear (RL = 1), favorable (0 < RL < 1), or irreversible (RL = 0) [90]. In these cases, an RL value less than 1 indicates that adsorption is favorable. Another Langmuir constant, Q0, indicates the adsorption capacity. During this study, the value of RL was found to be 0.007 (Ti2O3) and 0.003 (Ti2O3/mAPF MMS), suggesting that methyl violet adsorption on both was favorable for Ti2O3 and Ti2O3/mAPF MMSs. Furthermore, the value of the Freundlich isotherm slope was found to be 0.506 (Ti2O3) and 0.096 (Ti2O3/mAPF MMS), suggesting that a large amount of methyl violet was adsorbed by Ti2O3 and Ti2O3/mAPF MMSs. A comparison of the previous studies of different dyes adsorbed on various bio adsorbents and methyl violet/EBT adsorbed on different adsorbents is given in Table 2 and Table 3 [82,83,84,85,86,87,88,89,90]. As seen from Table 2 and Table 3, the Ti2O3/mAPF MMSs had a moderate catalytic performance with the degradation of cationic dye, suggesting that Ti2O3/mAPF MMSs have an excellent application compared to other adsorbents.

3.3. Kinetics of Adsorption Study

Kinetics of adsorption was used to understand the pathway of adsorption of methyl violet upon the Ti2O3/mAPF MMSs. It was noted that the organic dye degradation primarily followed first-order reaction kinetics [91,92,93,94,95]. The experimental data results were fitted into a first-order model for the adsorption of methyl violet onto Ti2O3/mAPF MMSs. The results of the kinetics of the adsorption of methyl violet on Ti2O3/mAPF MMSs are shown in Table 4. It can be seen that the first-order model fit better than the pseudo-first-order model, with a value of R2 = 0.999 for the Ti2O3/mAPF MMSs. Similarly, the calculated qe value fit very well with the experimental data. The results suggest that the degradation of methyl violet over Ti2O3/mAPF MMSs can be described by the first-order reaction of kinetic ln (Ci/Ct) = kt, where k is the rate constant (h−1), Ci is the initial concentration, and Ct is the dye concentration at time. Linear correlations are important, as evidenced from the value of ‘r’. This indicates the applicability of these kinetic equations and the first-order of the adsorption process.

3.4. Activation Parameters

The Van’t Hoff equation [96,97] gives the relationship between the standard Gibbs free energy change and the equilibrium constant. It is expressed as follows:
Gibbs free energy change = −RT ln equilibrium constant reaction for adsorption
Gibbs free energy change = enthalpy of activation − T × entropy of activation
The activation parameters affecting the methyl violet adsorption on Ti2O3/mAPF MMSs and Ti2O3 are shown in Table 5.
The Gibbs free energy change was calculated using Equation (1) where the entropy and enthalpy of activation was calculated from the slope and intercept of lnK vs. 1/T. As seen from Table 5, at all temperatures (20 to 45 °C) for both Ti2O3/mAPF MMSs and Ti2O3, the Gibbs free energy change was negative, suggesting a spontaneous and favorable adsorption process that also decreased with an increase in temperature, which shows the endothermic reaction of adsorption [98,99]. In addition, the adsorption of methyl violet is a physically controlled process that indicates that ΔG° fell in the region of −20 to 0 KJ mol−1 [100]. Hence, based on the Mattson criteria [40], we can safely assume that the adsorption of methyl violet onto Ti2O3/mAPF MMSs and Ti2O3 is a physical process, and that it may involve physical forces such dipole and/or hydrogen bonding forces. The positive enthalpy of activation (2.383 KJ mol−1 (Ti2O3/mAPF MMS) and 2.168 KJ mol−1 (Ti2O3) suggests an endothermic adsorption process in both cases. The positive entropy of activation (15.39 JK−1 mol−1 (Ti2O3/mAPF MMS) and 13.55 JK−1 mol−1 (Ti2O3)) suggests an increase in temperature assisted in dye adsorption by the removal of H2O molecules from the surface of the Ti2O3/mAPF MMSs [88,101,102].

3.5. Desorption Studies

In the present desorption study, we used various solutions of pH 9, 10, 11, and 12 to desorb methyl violet adsorbed on the Ti2O3/mAPF MMSs. Ours result suggest a lower adsorption at higher pH. When increasing the pH from pH 9 to pH 12, the desorption of methyl violet increased from 21.95 to 43.13%, indicating an increased repulsion between the adsorption sites and methyl violet molecules. The moderate desorption efficiency of the Ti2O3/mAPF MMSs suggests that the interactions between the methyl violet molecules and the Ti2O3/mAPF MMSs were via conjugated л bonding and hydrogen bonding.

4. Conclusions

The present investigation dealt with the removal of the organic dye methyl violet on Ti2O3/mAPF MMSs. Adsorption isotherms and kinetic studies of the Ti2O3/mAPF MMSs are shown first order of adsorption process. The desorption of methyl violet increased with a decrease in the concentration and particle size. With an increase in the contact time and amount of the adsorbent, the percentage removal increased. The Ti2O3/mAPF MMSs improved the actual adsorption in the elimination of crystal violet; successful dye decolorization was attained within 60 min. The experimental data confirm the first-order kinetics for the Ti2O3/mAPF MMS samples. Ti2O3/mAPF MMSs showed a good catalytic performance with the degradation of methyl violet, indicating that Ti2O3/mAPF MMSs have an excellent application in environmental protection. It can be concluded from these results that Ti2O3/mAPF MMSs are the best adsorption method for the removal of methyl violet from an industrial effluent.

Author Contributions

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

Funding

The authors express their sincere appreciation to the Researchers Supporting Project number (RSPD2023R723) at King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Gupta, N.; Singh, H.P.; Sharma, R.K. Metal nanoparticles with high catalytic activity in degradation of methyl orange: An electron relay effect. J. Mol. Catal. A Chem. 2011, 335, 248–252. [Google Scholar] [CrossRef]
  2. Dong, Y.; Liu, T.; Sun, S.; Chang, X.; Guo, N. Preparation and characterization of SiO2/polydopamine/Ag nanocomposites with long-term antibacterial activity. Ceram. Int. 2014, 40, 5605–5609. [Google Scholar] [CrossRef]
  3. Yang, Y.; Ji, H.; Duan, H.; Fu, Y.; Xia, S.; Lu, C. Controllable synthesis of mussel-inspired catechol-formaldehyde resin microspheres and their silver-based nanohybrids for catalytic and antibacterial applications. Polym. Chem. 2019, 10, 4537–4550. [Google Scholar] [CrossRef]
  4. Greczynski, G.; Hultman, L. Reliable determination of chemical state in X-ray photoelectron spectroscopy based on sample-work-function referencing to adventitious carbon: Resolving the myth of apparent constant binding energy of the C 1s peak. Appl. Surf. Sci. 2018, 451, 99–103. [Google Scholar] [CrossRef]
  5. Gong, C.; Li, Q.; Zhou, H.; Liu, R. Tiny Au satellites decorated Fe3O4@3-aminophenol-formaldehyde core-shell nanoparticles: Easy synthesis and comparison in catalytic reduction for cationic and anionic dyes. Colloids. Surf. A Physicochem. Eng. Asp. 2018, 540, 67–72. [Google Scholar] [CrossRef]
  6. Gong, C.; Zhou, Z.; Li, J.; Zhou, H.; Liu, R. Facile synthesis of ultra stable Fe3O4@Carbon core-shell nanoparticles entrapped satellite au catalysts with enhanced 4-nitrophenol reduction property. J. Taiwan. Inst. Chem. Eng. 2018, 84, 229–235. [Google Scholar] [CrossRef]
  7. Alprol, A.; Heneash, A.; Ashour, M.; Abualnaja, K.; Alhashmialameer, D.; Mansour, A.; Sharawy, Z.; Abu-Saied, M.; Abomohra, A. Potential Applications of Arthrospira platensis Lipid-Free Biomass in Bioremediation of Organic Dye from Industrial Textile Effluents and Its Influence on Marine Rotifer (Brachionus plicatilis). Materials 2021, 14, 4446. [Google Scholar] [CrossRef]
  8. Gong, R.; Li, M.; Yang, C.; Sun, Y.; Chen, J. Removal of cationic dyes from aqueous solution by adsorption on peanut hull. J. Hazard. Mater. 2005, 121, 247–250. [Google Scholar] [CrossRef]
  9. Barka, N.; Abdennouri, M.; EL Makhfouk, M. Removal of Methylene Blue and Eriochrome Black T from aqueous solutions by biosorption on Scolymus hispanicus L.: Kinetics, equilibrium and thermodynamics. J. Taiwan Inst. Chem. Eng. 2011, 42, 320–326. [Google Scholar] [CrossRef]
  10. Islam, A.; Ahmad, A.; Laskar, M.A. Characterization of a Chelating Resin Functionalized via Azo Spacer and Its Analytical Applicability for the Determination of Trace Metal Ions in Real Matrices. J. Appl. Polym. Sci. 2011, 123, 3448–3458. [Google Scholar] [CrossRef]
  11. Ejhieh, N.; Khorsandi, M. Photodecolorization of Eriochrome Black T using NiS-P zeolite as a heterogeneous catalyst. J. Hazard. Mater. 2010, 176, 629–637. [Google Scholar] [CrossRef]
  12. Yaqoob, A.; Noor, N.H.M.; Serrà, A.; Ibrahim, M.N.M. Advances and challenges in developing efficient graphene oxide-based ZnO photocatalysts for dye photo-oxidation. Nanomaterials 2020, 10, 932. [Google Scholar] [CrossRef]
  13. Badawi, K.; Ismail, B.; Baaloudj, O.; Abdalla, K.Z. Advanced wastewater treatment process using algal photo-bioreactor associated with dissolved-air flotation system: A pilot-scale demonstration. J. Water Process. Eng. 2022, 46, 102565. [Google Scholar] [CrossRef]
  14. Ahmad, A.; Hameed, B. Fixed-bed Adsorption of Reactive Azo Dye onto Granular Activated Carbon Prepared from Waste. J. Hazard. Mater. 2009, 175, 298–303. [Google Scholar] [CrossRef]
  15. Idris, M.O.; Kim, H.C. Exploring the effectiveness of microbial fuel cell for the degradation of organic pollutants coupled with bio-energy generation. Sustain. Energy Technol. Assess. 2022, 52, 102183. [Google Scholar]
  16. Crini, G. Non-conventional low-cost adsorbents for dye removal: A review. Bioresour. Technol. 2006, 97, 1061–1085. [Google Scholar] [CrossRef]
  17. Yaqoob, A.A.; Parveen, T.; Umar, K.; Ibrahim, M.N.M. Role of Nanomaterials in the Treatment of Wastewater: A Review. Water 2020, 12, 495. [Google Scholar] [CrossRef]
  18. Yaqoob, A.; Noor, N.H.M.; Umar, K.; Adnan, R.; Rashid, M. Graphene oxide–ZnO nanocomposite: An efficient visible light photocatalyst for degradation of rhodamine B. Appl. Nanosci. 2021, 11, 1291–1302. [Google Scholar] [CrossRef]
  19. García-Montaño, J.; Ruiz, N.; Muñoz, I.; Domènech, X.; García-Hortal, J.A.; Torrades, F.; Peral, J. Environmental assessment of different photo-Fenton approaches for commercial reactive dye removal. J. Hazard. Mater. 2006, 138, 218–225. [Google Scholar] [CrossRef]
  20. Hussein, F.H. Comparison between solar and artificial photocatalytic decolorization of textile industrial wastewater. Int. J. Photoenergy 2012, 2012, 793648. [Google Scholar] [CrossRef]
  21. Chong, M.N.; Jin, B.; Chow, C.W.K.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef] [PubMed]
  22. Lachheb, H.; Puzenat, E.; Houasetal, A. Photocatalyticdegradation of various types of dyes (Alizarin S, Crocein Orange G, Methyl Red, Congo Red, Methylene Blue) in water by UV irradiated titania. Appl. Catal. B Environ. 2002, 39, 75–90. [Google Scholar] [CrossRef]
  23. Zhang, Q.; Jing, Y.H.; Shiue, A.; Chang, C.T.; Chen, B.Y.; Hsueh, C.C. Deciphering effects of chemical structure on azo dye decolorization/degradation characteristics: Bacterial vs. photocatalytic method. J. Taiwan Inst. Chem. Eng. 2012, 43, 760–766. [Google Scholar] [CrossRef]
  24. Hsueh, C.; Chen, B.Y. Comparative study on reaction selectivity of azo dye decolorization by Pseudomonas luteola. J. Hazard. Mater. 2007, 141, 842–849. [Google Scholar] [CrossRef] [PubMed]
  25. McCullagh, C.; Skillen, N.; Adams, M.; Robertson, P.K. Photocatalytic reactors for environmental remediation: A review. J. Chem. Technol. Biotechnol. 2011, 86, 1002–1017. [Google Scholar] [CrossRef]
  26. Chan, S.H.S.; Wu, T.Y.; Juan, J.C.; Teh, C.Y. Recent developments of metal oxide semiconductors as photocatalysts in advanced oxidation processes (AOPs) for treatment of dye waste-water. J. Chem. Technol. Biotechnol. 2011, 86, 1130–1158. [Google Scholar] [CrossRef]
  27. Guo, M.Y.; Ng, A.M.C.; Liu, F.; Djurisic, A.B.; Chan, W.K. Photocatalytic activity of metal oxides-the role of holes and OH radicals. Appl. Catal. B Environ. 2011, 107, 150–157. [Google Scholar] [CrossRef]
  28. Fenoll, J.; Hellin, P.; Mart, C.M.; Flores, P.; Navarro, S. Semiconductor oxides-sensitized photodegradation of fenamiphos in leaching water under natural sun light. Appl. Catal. B Environ. 2012, 115–116, 31–37. [Google Scholar] [CrossRef]
  29. Abualnaja, K.M.; Alprol, A.E.; Abu-Saied, M.A.; Ashour, M.; Mansour, A.T. Removing of anionic dye from aqueous solutions by adsorption using of multiwalled carbon nanotubes and poly (Acrylonitrile-styrene) impregnated with activated carbon. Sustainability 2021, 13, 7077. [Google Scholar] [CrossRef]
  30. Krifka, S.; Spagnuolo, G.; Schmalz, G.; Schweikl, H. A review of adaptive mechanisms in cell responses towards oxidative stress caused by dental resin monomers. Biomaterials 2013, 34, 4555–4563. [Google Scholar] [CrossRef]
  31. Liu, J.; Qiao, S.; Liu, H.; Chen, J.; Orpe, A.; Zhao, D.; Lu, G. Extension of the Stober Method to the Preparation of Monodisperse Resorcinol-Formaldehyde Resin Polymer and Carbon Spheres. Angew. Chem. Int. Ed. 2011, 50, 5947–5951. [Google Scholar] [CrossRef]
  32. Yang, P.; Xu, Q.; Jin, S.; Zhao, Y.; Lu, Y.; Xu, X.; Yu, S. Synthesis of Multifunctional Ag@Au@Phenol Formaldehyde Resin Particles Loaded with Folic Acids for Photothermal Therapy. Chem. Eur. J. 2012, 18, 9294–9299. [Google Scholar] [CrossRef]
  33. Deng, S.; Lei, J.; Yao, X.; Huang, Y.; Lin, D.; Ju, H. Electrocatalytic reduction of coreactant by highly loaded dendrimer-encapsulated palladium nanoparticles for sensitive electrochemiluminescent immunoassay. Chem. Commun. 2012, 48, 9159–9161. [Google Scholar] [CrossRef]
  34. Ma, C.; Song, Y.; Shi, J.; Zhang, D.; Guo, Q.; Liu, L. Preparation and electrochemical performance of heteroatom-enriched electrospun carbon nanofibers from melamine formaldehyde resin. J. Colloid Interface Sci. 2013, 395, 217–223. [Google Scholar] [CrossRef]
  35. Muylaert, I.; Borgers, M.; Bruneel, E.; Schaubroeck, J.; Verpoort, F.; Voort, P.V.D. Ultra stable ordered mesoporous phenol/formaldehyde polymers as a heterogeneous support for vanadium oxide. Chem. Commun. 2008, 37, 4475–4477. [Google Scholar] [CrossRef]
  36. Zhao, M.; Song, H. Catalytic Graphitization of Phenolic Resin. J. Mater. Sci. Technol. 2011, 27, 266–270. [Google Scholar] [CrossRef]
  37. Zhao, J.; Niu, W.; Zhang, L.; Cai, H.; Han, M.; Yuan, Y.; Majeed, S.; Anjum, S.; Xu, G. A Template-Free and Surfactant-Free Method for High-Yield Synthesis of Highly Monodisperse 3-Aminophenol-Formaldehyde Resin and Carbon Nano/Microspheres. Macromolecules 2013, 46, 140–145. [Google Scholar] [CrossRef]
  38. Mori, K.; Dojo, M.; Yamashita, H. Pd and Pd-Ag Nanoparticles within a Macroreticular Basic Resin: An Efficient Catalyst for Hydrogen Production from Formic Acid Decomposition. ACS Catal. 2013, 3, 1114–1119. [Google Scholar] [CrossRef]
  39. Wu, Y.; Li, Y.; Qin, L.; Yang, F.; Wu, D. Monodispersed or narrow-dispersed melamine–formaldehyde resin polymer colloidal spheres: Preparation, size-control, modification, bioconjugation and particle formation mechanism. J. Mater. Chem. B 2013, 1, 204–212. [Google Scholar] [CrossRef]
  40. He, X.H.; Wu, X.M.; Cai, X.; Lin, S.L.; Xie, M.R.; Zhu, X.Y.; Yan, D.Y. Functionalization of Magnetic Nanoparticles with Dendritic–Linear–Brush-Like Triblock Copolymers and Their Drug Release Properties. Langmuir 2012, 28, 11929–11938. [Google Scholar] [CrossRef]
  41. Wang, Y.Q.; Zou, B.F.; Gao, T.; Wu, X.P.; Lou, S.Y.; Zhou, S.M. Constructing carbon-coated Fe3O4 hierarchical microstructures with a porous structure and their excellent Cr(VI) ion removal properties. J. Mater. Chem. 2012, 22, 9034–9040. [Google Scholar] [CrossRef]
  42. Matsuno, R.; Yamamoto, K.; Otsuka, H.; Takahara, A. Polystyrene-Grafted Magnetite Nanoparticles Prepared through Surface-Initiated Nitroxyl-Mediated Radical Polymerization. Chem. Mater. 2003, 15, 3–5. [Google Scholar] [CrossRef]
  43. Zhang, H.; Zhong, X.; Xu, J.J.; Chen, H.Y. Fe3O4/Polypyrrole/Au Nanocomposites with Core/Shell/Shell Structure: Synthesis, Characterization, and Their Electrochemical Properties. Langmuir 2008, 24, 13748–13752. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, C.; Wang, Y.; Zhang, X.; Guo, H. Fe78Si9B13 amorphous alloys for the decolorization process of azo dye aqueous solutions with different initial concentrations at different reaction temperatures. Vacuum 2020, 176, 109301. [Google Scholar] [CrossRef]
  45. Qiao, J.-H.; Chen, K.; Li, S.-J.; Liu, Y.-C.; Cao, H.-W.; Wei, G.; Kong, L.-P.; Zhang, X.; Liu, H.-T. Plasma spray-chemical vapor deposition of nanotextured film with α/β Bi2O3 heterostructure and photocatalytic degradation performance. Vacuum 2021, 188, 110206. [Google Scholar] [CrossRef]
  46. Parida, D.; Moreau, E.; Nazir, R.; Salmeia, K.A.; Friston, R.; Zhao, R.; Lehner, S.; Jovic, M.; Gann, S. Smart hydrogel-microsphere embedded silver nanoparticle catalyst with high activity and selectivity for the reduction of 4-nitrophenol and azo dyes. J. Hazard. Mat. 2021, 416, 126237. [Google Scholar] [CrossRef]
  47. Kong, L.R.; Lu, X.F.; Jin, E.; Jiang, S.; Bian, X.J.; Zhang, W.J.; Wang, C. Accurately Tuning the Dispersity and Size of Palladium Particles on Carbon Spheres and Using Carbon Spheres/Palladium Composite as Support for Polyaniline in H2O2 Electrochemical Sensing. Langmuir 2010, 26, 5985–5990. [Google Scholar] [CrossRef]
  48. Xuan, S.H.; Wang, Y.X.J.; Leung, K.C.F.; Shu, K.Y. Synthesis of Fe3O4@Polyaniline Core/Shell Microspheres with Well-Defined Blackberry-Like Morphology. J. Phys. Chem. C 2008, 112, 18804–18889. [Google Scholar] [CrossRef]
  49. Milyutin, V.V.; Mikheev, S.V.; Gelis, V.M.; Kozlitin, E.A. Sorption of Cesium on Ferrocyanide Sorbents from Highly Saline Solutions. Radiochemistry 2009, 51, 298–300. [Google Scholar] [CrossRef]
  50. Zicman, L.R.; Neacsu, E.; Done, L.; Tugulan, L.; Dragolici, F.; Obreja, B.T.; Dobre, T. Removal of 137Cs ions from aqueous radioactive waste using nickel ferrocyanide, precipitated on silica gel. Bull. Rom. Chem. Eng. Soc. 2015, 2, 84–99. [Google Scholar]
  51. Nilchi, A.; Atashi, H.; Javid, A.H.; Saberi, R. Preparations of PAN-Based Adsorbers for Separation of Cesium and Cobalt from Radioactive Wastes. Appl. Radiat. Isot. 2007, 65, 482–487. [Google Scholar] [CrossRef] [PubMed]
  52. Kozlov, P.V.; Kazadaev, A.A.; Makarovskii, R.A.; Remizov, M.B.; Verbitskii, K.V.; Logunov, M.V. Development of a Process for Cesium Recovery from the Clarified Phase of High-Level Waste Storage Tanks of the Mayak Production Association with a Ferrocyanide Sorbent. Radiochemistry 2016, 58, 295–301. [Google Scholar] [CrossRef]
  53. Milyutin, V.V.; Gelis, V.M.; Ershov, B.G.; Seliverstov, A.F. Effect of Complexing Agents and Surfactants on Coprecipitation of Cesium Radionuclides with Nickel Ferrocyanide. Radiochemistry 2011, 50, 67–69. [Google Scholar] [CrossRef]
  54. Milyutin, V.V.; Zelenin, P.G.; Kozlov, P.V.; Remizov, M.B.; Kondrutskii, D.A. Sorption of Cesium from Alkaline Solutions onto Resorcinol-Formaldehyde Sorbents. Radiochemistry 2019, 61, 714–718. [Google Scholar] [CrossRef]
  55. Brown, G.N.; Russell, R.L.; Peterson, R.A. Small-Column Cesium Ion Exchange Elution Testing of Spherical Resorcinol-Formaldehyde; Pacific Northwest National Laboratory: Richland, WA, USA, 2011. [Google Scholar]
  56. Hubler, T.L.; Franz, J.A.; Shaw, W.J.; Bryan, S.; Hallen, R.; Brown, G.N.; Bray, L.A.; Linehan, J.C. Synthesis, Structural Characterization, and Performance Evaluation of Resorcinol-Formaldehyde (RF) Ion-Exchange Resin; Pacific Northwest Laboratory: Richland, WA, USA, 1995. [Google Scholar]
  57. Hubler, T.L.; Shaw, W.J.; Brown, G.N.; Linehan, J.C.; Franz, J.A.; Hart, T.R.; Hogan, M.O. Chemical Derivation to Enhance the Chemical/Oxidative Stability of Resorcinol-Formaldehyde (R-F) Resin; Pacific Northwest National Lab.: Richland, WA, USA, 1996. [Google Scholar]
  58. Taguchi, S.; Nakatani, T.; Saeki, H.; Tayakout-Fayolle, M.; Itoh, K.; Yamamoto, T. Characterization of Resorcinol–Formaldehyde Hydrogel as Adsorbent for Cesium Ion. Adsorption 2021, 27, 81–90. [Google Scholar] [CrossRef]
  59. Aquino, C.M.; Costero, A.M.; Gil, S.; Gavina, P. Resorcinol Functionalized Gold Nanoparticles for Formaldehyde Colorimetric Detection. Nanomaterials 2019, 9, 302. [Google Scholar] [CrossRef]
  60. Gupta, V.K. Suhas Application of low-cost adsorbents for dye removal—A review. J. Environ. Manag. 2009, 90, 2313–2342. [Google Scholar] [CrossRef]
  61. Alkan, M.; Demirbas, O.; Celikcapa, S.; Dogan, M. Sorption of Acid Red 57 from Aqueous Solutions onto Sepiolite. J. Hazard. Mater. 2004, 116, 135–145. [Google Scholar] [CrossRef]
  62. Turhan, K.; Ozturkcan, S.A. Decolorization and Degradation of Reactive Dye in Aqueous Solution by Ozonation in a Semi-Batch Bubble Column Reactor. Water Air Soil Pollut. 2012, 224, 1353. [Google Scholar] [CrossRef]
  63. Abdallah, R.; Taha, S. Biosorption of methylene blue from aqueous solution by nonviable Aspergillus fumigates. Chem. Eng. J. 2012, 195, 69–76. [Google Scholar] [CrossRef]
  64. Wang, B.E.; Hu, Y.Y. Comparison of four supports for adsorption of reactive dyes by immobilized Aspergillus fumigatus beads. J. Environ. Sci. 2007, 19, 451–457. [Google Scholar] [CrossRef]
  65. Wainwright, M.; Crossley, K.B. Methylene Blue-a therapeutic dye for all seasons? J. Chemother. 2002, 14, 431–443. [Google Scholar] [CrossRef]
  66. Frankenburg, F.R.; Baldessarini, R.J. Neurosyphilis, Malaria, and the Discovery of Antipsychotic Agents. Harv. Rev. Psychiatry 2008, 16, 299–307. [Google Scholar] [CrossRef]
  67. Chung, K.T.; Stevens, S.E. Degradation azo dyes by environmental microorganisms and helminths. Environ. Toxicol. Chem. 1993, 12, 2121–2132. [Google Scholar] [CrossRef]
  68. Aksu, Z.; Tezer, S. Equilibrium and Kinetic Modelling of Biosorption of Remazol Black B by Rhizopus arrhizus in a Batch System: Effect of Temperature. Process Biochem. 2000, 36, 431–439. [Google Scholar] [CrossRef]
  69. Lucas, M.S.; Peres, J.A. Decolorization of the Azo Dye Reactive Black 5 by Fenton and Photo Fenton Oxidatio. Dyes Pigm. 2006, 71, 236–244. [Google Scholar] [CrossRef]
  70. Albert, M.; Lessin, M.S.; Gilchrist, B.F. Methylene blue: Dangerous dye for neonates. J. Pediatr. Surg. 2003, 38, 1244–1245. [Google Scholar] [CrossRef]
  71. Srinivasan, M.; White, T. Degradation of Methylene Blue by Three-Dimensionally Ordered Macroporous Titania. Environ. Sci. Technol. 2007, 41, 4405–4409. [Google Scholar] [CrossRef]
  72. Zhang, Y.G.; Ma, L.L.; Li, J.L.; Yu, Y. In Situ Fenton Reagent Generated from TiO2/Cu2O Composite Film: A New Way to Utilize TiO2 under Visible Light Irradiation. Environ. Sci. Technol. 2007, 41, 6264–6269. [Google Scholar] [CrossRef]
  73. Bielska, M.; Szymanowski, J. Removal of methylene blue from waste water using micellar enhanced ultrafiltration. Water Res. 2006, 40, 1027–1033. [Google Scholar] [CrossRef]
  74. Hameed, B.H.; Ahmad, A.L.; Latiff, K.N.A. Adsorption of basic dye (methylene blue) onto activated carbon prepared from rattan sawdust. Dyes Pigm. 2007, 75, 143–149. [Google Scholar] [CrossRef]
  75. Zhang, S.; Fan, Q.; Gao, H.; Huang, Y.; Liu, X.; Li, J.; Xu, X.; Wang, X. Formation of Fe3O4@MnO2 ballin-ball hollow spheres as a high performance catalyst with enhanced catalytic performances. J. Mater. Chem. A 2016, 4, 1414–1422. [Google Scholar] [CrossRef]
  76. Wang, Y.; Pan, F.; Dong, W.; Xu, L.; Wu, K.; Xu, G.; Chen, W. Recyclable silver-decorated magnetic titania nanocomposite with enhanced visible-light photocatalytic activity. Appl. Catal. B Environ. 2016, 189, 192–198. [Google Scholar] [CrossRef]
  77. Gawande, M.B.; Branco, P.S.; Varma, R.S. Nano-magnetite (Fe3O4) as support for recyclable nanocatalysts in the development of sustainable methodologies. Chem. Soc. Rev. 2013, 42, 3371–3393. [Google Scholar] [CrossRef]
  78. Varma, R.S. Journey on greener pathways: From the use of alternate energy inputs and benign reaction media to sustainable applications of nano-catalysts in synthesis and environmental remediation. Green Chem. 2014, 16, 2027–2041. [Google Scholar] [CrossRef]
  79. Sharma, V.K. Festschrift in Honor of Rajender S. Varma. ACS Sustain. Chem. Eng. 2016, 4, 640–642. [Google Scholar] [CrossRef]
  80. Baig, R.B.N.; Varma, R.S. Organic synthesis via magnetic attraction: Benign and sustainable protocols using magnetic nanoferrites. Green Chem. 2013, 15, 398–417. [Google Scholar] [CrossRef]
  81. Atarod, M.; Nasrollahzadeh, M.; Sajadi, S.M. Green synthesis of Pd/RGO/Fe3O4 nanocomposite using Withania coagulans leaf extract and its application as magnetically separable and reusable catalyst for the reduction of 4-nitrophenol. J. Colloid Interface Sci. 2016, 465, 249–258. [Google Scholar] [CrossRef]
  82. Sajadi, S.M.; Nasrollahzadeh, M.; Maham, M. Aqueous extract from seeds of Silybum marianum L. as a green material for preparation of the Cu/Fe3O4 nanoparticles: A magnetically recoverable and reusable catalyst for the reduction of nitroarenes. J. Colloid Interface Sci. 2016, 469, 93–98. [Google Scholar] [CrossRef]
  83. Veremchuk, I.; Antonyshyn, I.; Candolfi, C.; Feng, X.; Burkhardt, U.; Baitinger, M.; Zhao, J.-T.; Grin, Y. Diffusion-Controlled Formation of Ti2O3 during Spark-Plasma Synthesis. Inorg. Chem. 2013, 52, 4458–4463. [Google Scholar]
  84. Yang, K.; Peng, H.B.; Wen, Y.H.; Li, N. Re-examination of characteristic FTIR spectrum of secondary layer in bilayer oleic acid-coated Fe3O4 nanoparticles. Appl. Surf. Sci. 2010, 256, 3093–3097. [Google Scholar] [CrossRef]
  85. Lasperas, M.; Llorett, T.; Chaves, L.; Rodriguez, I.; Cauvel, A.; Brunel, D. Amine functions linked to MCM-41-type silicas as a new class of solid base catalysts for condensation reactions. Stud. Surf. Sci. Catal. 1997, 108, 75–82. [Google Scholar] [CrossRef]
  86. Allen, D.J.; Ishida, H. Synthesis and properties of polybenzoxazines containing pyridyl group. Polymer 2007, 48, 6763–6772. [Google Scholar] [CrossRef]
  87. Chernykh, I.; Liu, J.P.; Ishida, H. Synthesis and properties of a new crosslinkable polymer containing benzoxazine moiety in the main chain. Polymer 2006, 47, 7664–7669. [Google Scholar] [CrossRef]
  88. Ma, Z.Y.; Guan, Y.P.; Liu, H.Z. synthesis and characterization of micron-sized monodisperse superparamagnetic polymer particles with amino groups. J. Polym. Sci. Part A Polym. Chem. 2005, 43, 3433–3439. [Google Scholar] [CrossRef]
  89. Ramanathan, T.; Fisher, F.T.; Ruoff, R.S.; Brinson, L.C. Amino-functionalized carbon nanotubes for binding to polymers and biological systems. Chem. Mater. 2005, 17, 1290–1295. [Google Scholar] [CrossRef]
  90. Morlieras, J.; Chezal, J.-M.; Miot-Noirault, E.; Roux, A.; Heinrich-Balard, L.; Cohen, R.; Tarrit, S. Development of gadolinium based nanoparticles having an affinity towards melanin. Nanoscale 2013, 5, 1603–1615. [Google Scholar] [CrossRef]
  91. Jahagirdar, S.S.; Shrihari, S.; Manu, B. Reuse of incinerated textile mill sludge as adsorbent for dye removal. KSCE J. Civ. Eng. 2015, 19, 1982–1986. [Google Scholar] [CrossRef]
  92. Rehman, M.S.U.; Kim, I.; Rashid, N.; Umer, M.A.; Sajid, M.; Han, J.I. Adsorption of Brilliant Green Dye on Biochar Prepared From Lignocellulosic Bioethanol Plant Waste. Clean Soil Air Water 2016, 44, 55–62. [Google Scholar] [CrossRef]
  93. Laskar, N.; Kumar, U. Removal of Brilliant Green dye from water by modified Bambusa Tulda: Adsorption isotherm, kinetics and thermodynamics study. Int. J. Environ. Sci. Technol. 2018, 16, 1649–1662. [Google Scholar] [CrossRef]
  94. Moeinpour, F.; Alimoradi, A.; Kazemi, M. Efficient removal of Eriochrome black-T from aqueous solution using NiFe2O4 magnetic nanoparticles. J. Environ. Health Sci. Eng. 2014, 12, 112. [Google Scholar] [CrossRef] [PubMed]
  95. Khalid, A.; Zubair, M.; Ihsanullah, A. A Comparative Study on the Adsorption of Eriochrome Black T Dye from Aqueous Solution on Graphene and Acid-Modified Graphene. Arab. J. Sci. Eng. 2018, 43, 2167–2179. [Google Scholar] [CrossRef]
  96. Fayoud, N.; Tahiri, S.; Younssi, S.A.; Albizane, A.; Gallart-Mateu, D.; Cervera, M.L.; De la Guardia, M. Kinetic, isotherm and thermodynamic studies of the adsorption of methylene blue dye onto agro-based cellulosic materials. Desalination Water Treat. 2016, 57, 16611–16625. [Google Scholar] [CrossRef]
  97. Baidya, K.S.; Kumar, U. Adsorption of brilliant green dye from aqueous solution onto chemically modified areca nut husk. Afr. J. Chem. Eng. 2021, 35, 33–43. [Google Scholar] [CrossRef]
  98. Salman, S.M.; Ali, A.; Khan, B.; Iqbal, M.; Alamzeb, M. Thermodynamic and kinetic insights into plant-mediated detoxification of lead, cadmium, and chromium from aqueous solutions by chemically modified Salvia moorcroftiana leaves. Environ. Sci. Pollut. Res. Int. 2019, 26, 14339–14349. [Google Scholar] [CrossRef]
  99. Mane, V.S.; Babu, P.V.V. Studies on the adsorption of Brilliant Green dye from aqueous solution onto low-cost NaOH treated saw dust. Desalination 2011, 273, 321–329. [Google Scholar] [CrossRef]
  100. Auta, M.; Hameed, B.H. Modified mesoporous clay adsorbent for adsorption isotherm and kinetics of methylene blue. Chem. Eng. J. 2012, 198–199, 219–227. [Google Scholar] [CrossRef]
  101. Mattson, J.S.M.H.B. Activated Carbon: Surface Chemistry and Adsorption from Solution; M. Dekker: New York, NY, USA, 1971. [Google Scholar]
  102. Rashidi, R.; Khaniabadi, Y.O.; Ghaderpoori, M. Adsorption of Eriochrome black-T from aqueous environment by raw Montmorillonite. Int. J. Environ. Anal. Chem. 2021, 1–15. [Google Scholar] [CrossRef]
Scheme 1. Plausible mechanism for the formation of Ti2O3/mAPF MMSs.
Scheme 1. Plausible mechanism for the formation of Ti2O3/mAPF MMSs.
Processes 11 01250 sch001
Figure 1. (A) SEM images for (a) mAPF and (b) Ti2O3/mAPF MMSs. (B) PXRD of (a) mAPF MMSs, (b) Ti2O3/mAPF MMSs.
Figure 1. (A) SEM images for (a) mAPF and (b) Ti2O3/mAPF MMSs. (B) PXRD of (a) mAPF MMSs, (b) Ti2O3/mAPF MMSs.
Processes 11 01250 g001
Figure 2. FTIR spectra of (a) mAPF, (b) Ti2O3 and (c) Ti2O3/mAPF MMS.
Figure 2. FTIR spectra of (a) mAPF, (b) Ti2O3 and (c) Ti2O3/mAPF MMS.
Processes 11 01250 g002
Figure 3. TGA curve of Ti2O3/mAPF (blue line—A) and the DTA curve of Ti2O3/mAPF (red line—B).
Figure 3. TGA curve of Ti2O3/mAPF (blue line—A) and the DTA curve of Ti2O3/mAPF (red line—B).
Processes 11 01250 g003
Figure 4. (A) UV−Vis spectra of methyl violet adsorption by the Ti2O3/mAPF MMSs at various time intervals. (B) The corresponding photographic image of dye degradation.
Figure 4. (A) UV−Vis spectra of methyl violet adsorption by the Ti2O3/mAPF MMSs at various time intervals. (B) The corresponding photographic image of dye degradation.
Processes 11 01250 g004
Table 1. The results of various adsorption isotherms.
Table 1. The results of various adsorption isotherms.
IsothermTi2O3/mAPF MMS Ti2O3
−x/m = k.P1/n
Slope (1/n)0.0960.506
Intercept (log k)1.9730.350
Correlation coefficient (r)0.9390.990
Langmuir (xm = aP1 + bP)
Intercept (1/ Q0b)0.0060.613
Correlation coefficient (r)0.9990.994
Q0 (mg g−1)132.6213.94
b (L mg−1)0.7958.571
RL0.0030.007
Table 2. Different types of adsorbents used for the decontamination of different dyes.
Table 2. Different types of adsorbents used for the decontamination of different dyes.
S. NoAdsorbentDyesQ0 (mg g−1)
1Sodium carbonate—Treated B VulgarisBG40.12
2Hydrochloric acid—Treated BambusaBG32.15
3Dialectical Behavior Therapy—Treated B VulgarisBG31.02
4Coconut fiberCongo Red2.80
5Musa sapientumBB919.90
6orange zestAcid Violet19.50
7stalks of grassesBB919.80
8Sugarcane megassAcid Orange 105.97
9Phoenix dactylifera (kernels)BB916.80
10Coconut fiberDirect red 286.70
11Citrullus lanatusCrystal Violet11.90
12Coal-based adsorbentDirect brown 15.90
13Rice hullBB919.50
14Cobnuts or filbertsBB98.80
15Live oaks BG2.08
16Rice hull ashBG24.13
17Saraca asoca BG123.0
18Rice straw BG113.10
19SCBABG118.17
20Native Allium sativumEBT100.22
21Washed Allium sativumEBT89.40
22Ti2O3/mAPF MMSs (this Study)Methyl Violet132.62
EBT = Eriochrome Black T; BG = Brilliant Green; BB9 = Basic Blue 9.
Table 3. Comparative study of the Ti2O3/mAPF MMSs vs. different adsorbents.
Table 3. Comparative study of the Ti2O3/mAPF MMSs vs. different adsorbents.
S. NoAdsorbent DetailQ0 (mg g−1)
1Graphite101.02
2Acid-modified Graphite69.39
3Native pyrena6.8
4Cold plasma treated-pyrena shells19.08
5Microwave treated-pyrena shells 30.38
5Magnetite/silica/pectinNPs66.28
6Ponded ash95.87
7Phosphoric acid -modified berry cultivation132.04
8Nickel ferrite@NPs82.41
9Nickel ferrite magnetic NPs 48.2
10C12H20N2O5 (Hydrophobic cross-linked)16.3
11β-CD21.4
12Nickel–iron alloy-hydroxides132.4
13Sweet lime-activated carbon47.43
14Smectite clay composite3.50
15China clay/polymer0.56
16Steatite or soapstone2.15
17Hydrochloric acid modified clay17.18
18Sulfuric acid modified clay17.48
19Al2H2O12Si4 (Gelwhite L. Bentolite.)100.5
20Stringybark53.25
21Native Allium sativum99.52
22Washed Allium sativum89.40
23Ti2O3/mAPF MMS (This Study)132.62
Table 4. Experimental evidence for the studies of the kinetics of the adsorption of methyl violet on Ti2O3/mAPF MMSs.
Table 4. Experimental evidence for the studies of the kinetics of the adsorption of methyl violet on Ti2O3/mAPF MMSs.
Time (min)Log (Ci/Ct)5 + Log [1 − U (t)]2 + Log (qe − qt)
100.614.322.29
200.644.272.24
300.834.001.98
400.853.951.96
501.023.691.66
601.073.271.23
Table 5. The thermodynamic activation parameters for the methyl violet adsorption.
Table 5. The thermodynamic activation parameters for the methyl violet adsorption.
AdsorbentTemp (K)KLΔG°
KJ mol−1
ΔG°
KJ mol−1
ΔS°
JK−1 mol−1
R2
Ti2O3/mAPF MMS2932.380−2.1122.43315.4800.989
2982.412−2.180
3032.445−2.247
3082.483−2.326
3132.536−2.415
3182.565−2.485
Ti2O32932.031−1.7162.27213.6650.944
2982.052−1.780
3032.130−1.904
3082.155−1.936
3132.187−1.998
3182.218−2.068
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

Radha, S.; Christygnanatheeba, P.; Nagaraj, K.; Prasad, S.; AlSalhi, M.S.; Kumar, J.V.; Arunachalam, P.; Karuppiah, C. Titanium(III) Oxide Doped with meta-Aminophenol Formaldehyde Magnetic Microspheres: Enhancing Dye Adsorption toward Methyl Violet. Processes 2023, 11, 1250. https://doi.org/10.3390/pr11041250

AMA Style

Radha S, Christygnanatheeba P, Nagaraj K, Prasad S, AlSalhi MS, Kumar JV, Arunachalam P, Karuppiah C. Titanium(III) Oxide Doped with meta-Aminophenol Formaldehyde Magnetic Microspheres: Enhancing Dye Adsorption toward Methyl Violet. Processes. 2023; 11(4):1250. https://doi.org/10.3390/pr11041250

Chicago/Turabian Style

Radha, Suriyan, Paul Christygnanatheeba, Karuppiah Nagaraj, Saradh Prasad, Mohamad Saleh AlSalhi, Jeyaraj Vinoth Kumar, Prabhakarn Arunachalam, and Chelladurai Karuppiah. 2023. "Titanium(III) Oxide Doped with meta-Aminophenol Formaldehyde Magnetic Microspheres: Enhancing Dye Adsorption toward Methyl Violet" Processes 11, no. 4: 1250. https://doi.org/10.3390/pr11041250

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop