ISSN : 0970 - 020X, ONLINE ISSN : 2231-5039
     FacebookTwitterLinkedinMendeley

Unique Magnetically Separable MnFe2O4/HAP Nanocomposite Photocatalyst for the Removal of Dye Pollutants

Annette Jayam Somasundaram and I. Sharmila Lydia*

Department of Chemistry, Bishop Heber College (Autonomous), Affiliated to Bharathidasan University, Tiruchirappalli, 620017, Tamil Nadu, India.

Corresponding Author E-mail: sharmilalydia1966@gmail.com

DOI : http://dx.doi.org/10.13005/ojc/400405

Article Publishing History
Article Received on : 03 May 2024
Article Accepted on : 18 Jul 2024
Article Published : 25 Jul 2024
Article Metrics
Article Review Details
Reviewed by: Dr.T. Mary Vargheese
Second Review by: Dr. Suraj Pratap Singh
Final Approval by: Dr. Ayssar Nahle
ABSTRACT:

In this present work, co-precipitation approach was employed to produce manganese ferrite (MnFe2O4) nanoparticles and hydroxyapatite (HAP) obtained from microwave assisted co-precipitation method was loaded on the synthesized MnFe2O4 to produce MnFe2O4/HAP nanocomposite which is used for the application as photocatalyst. The appearance of sharp peaks in the XRD diffractogram for MnFe2O4 shows high degree of crystallinity. The presence of flaky spherical agglomerate morphology was confirmed through SEM analysis. The distinct and sharp peaks at 563 cm-1 and 411cm-1 in the FT-IR spectra confirm the presence of ferrite phase. Diffuse reflectance analysis confirms that the as prepared photocatalysts absorb light in the visible region. The energy gap values of as-synthesized MnFe2O4 nanoparticles and MnFe2O4/HAP nanocomposite calculated using Tauc plot were found to be 1.7 eV and 1.5eV respectively. Photocatalytic activity of MnFe2O4 and MnFe2O4/HAP nanocomposites was ascertained for BV3 dye degradation under the illumination of visible light in aqueous solution. The results revealed that the degradation efficiency of 85.35 % and 93.24 % was achieved with MnFe2O4 nanoparticles and MnFe2O4/HAP nanocomposite respectively owing to the higher crystallinity and simultaneous adsorption capacity of MnFe2O4/HAP nanocomposite photocatalyst compared to MnFe2O4 nanoparticles.

KEYWORDS:

Basic violet 3; Hydroxyapatite; MnFe2O4; Phytotoxicity

Download this article as: 

Copy the following to cite this article:

Somasundaram A. J, Lydia L. S. Unique Magnetically Separable MnFe2O4/HAP Nanocomposite Photocatalyst for the Removal of Dye Pollutants. Orient J Chem 2024;40(4).


Copy the following to cite this URL:

Somasundaram A. J, Lydia L. S. Unique Magnetically Separable MnFe2O4/HAP Nanocomposite Photocatalyst for the Removal of Dye Pollutants. Orient J Chem 2024;40(4). Available from: https://bit.ly/3LH8idA


Introduction

The escalated population combined with speedy development and modernization have worsened the accessibility of fresh water. The polluted water discharged exclusively from textile, printing, leather and dye industries pose severe threat to human and other forms of life. The main contaminants are colourants, and their removal is a paramount concern and a challenging task for researchers 1-2.  Approximately 14% of the total global production of these colourants are being constantly mixed with water bodies. The cationic dyes are more virulent than the other types of dyes ascribable to their harmfulness and the presence of plenty of non-localized pi electrons in the aromatic ring system 3. The removal of dye molecules from industrial wastewater prior to their release into waterbodies is crucial in the current scenario. The practical feasibility of remediation using the conventional techniques such as biological, chemical and physical methods seems to be unfruitful due to sludge formation, high operational cost and formation of secondary metabolites4. For complete and partial mineralization of dye molecules, advanced oxidation process namely photocatalytic degradation can be a wonderful process for the removal of colourants from the contaminated industrial effluent water 5. Hence research is being oriented to develop cost effective visible light responsive heterogeneous photocatalysts.

Among the plentiful metal-oxide based photocatalysts spinel ferrite magnetic nanoparticles are given much importance due to excellent catalytic, magnetic, optical and photocatalytic properties. In addition, ferrite based photocatalysts possess chemical stability and reusability 6. Su.et.al (2012) developed manganese ferrite via a redox method using H2O2 for the decolourisation of methyl violet dye under the radiance of visible light 7. More magnetic composite materials are functionalized with conducting polymers, carbon-based derivatives etc., where magnetic metal oxide material plays a part as core and the conducting material contributes as the shell, that can be implemented for the destruction of dye molecules. Extensive research is being carried out with core-shell magnetic composites coated with Ag, Au, Si and hydroxyapatite (HAP) to improve sturdiness, environmental friendliness and recyclability 8. The ferrite substances can be doped with inorganic entities namely hydroxyapatite with massively ameliorate photocatalytic activity of the entire composite as a whole. This doped material namely ferrites/HAP can also have numerous biomedical applications. Zarei et.al., (2016) prepared MnFe2O4/HAP as eco-friendly novel photocatalyst with good recycling ability 9. Tedsree et.al., (2017)prepared core shell MnFe2O4 / HAP and determined its photodegradation capability 10. Nguyen et al., (2014) have synthesized bifunctional core -shell nanocomposite MnFe2O4/HAP for photodegradation of methyl violet dye 11. Core shell structured MnFe2O4/HAP nanocomposites with enhanced photocatalytic performance by visible light illumination was  reported by Xia. Y., & Yin. L. (2013) 12.

Basic violet 3 (BV3) is an organic compound which is used as colourant in textile and paint industries to give deep violet colour to clothes, paints and ink. BV3 is highly harmful to flora and fauna with enduring effects 13-14. Owing to its toxicity, it has been sorted out to be a model dyestuff in this study. This study focuses on the synthesis of HAP supported manganese ferrite MnFe2O4/HAP nanocomposite using microwave assisted co-precipitation method without any surfactants, which may possess an immense ability to degrade the organic contaminants. The composite is magnetically separable with much recyclability that can be used for the degradation of model dye BV3 and dyes present in textile industry wastewater. Also, it has been planned to assess the toxicity of the treated water by assessing growth of Vigna radiata.

Materials and Methods

Materials

Manganese acetate, and calcium hydroxide were purchased from Nice Chemicals for the use as precursors in the synthesis of MnFe2O4/HAP and Iron (III) chloride, sodium hydroxide, ammonium dihydrogen phosphate, basic violet 3 were sourced from Merck. All the samples purchased were of analytical grade and used without further  purification.

Synthesis of MnFe2O4 nanoparticles

MnFe2O4 nanoparticles were prepared by co-precipitation method. Firstly, a clear solution was prepared by dissolving 8g of ferric chloride and 3g of manganese acetate in deionized water. To this solution, 2M of NaOH was added drop wise under vigorous stirring using magnetic stirrer by fixing the temperature at 70oC and the reaction was carried out for 1h. The resulting solution was cooled to 25oC. The brownish-black precipitate of MnFe2O4 formed was dried overnight in hot air oven at 80oC 15.

Synthesis of Hydroxyapatite (HAP)

Hydroxyapatite was synthesized using microwave assisted co-precipitation method. Exactly calcium hydroxide (1.4 M) and ammonium dihydrogen phosphate (0.84M) solutions were prepared using double distilled water. To the hot Ca(OH)2 solution, ammonium dihydrogen phosphate was added slowly with constant stirring using a magnetic stirrer. When the addition was complete the resulting solution was kept in a microwave oven for 5 minutes. Throughout the experiment, the reaction mixture was maintained at the temperature range of 60ºC-80ºC until the pH of the reaction mixture becomes basic in nature. The reaction mixture was allowed to run for 5 hours at this pH condition. The resulting milky white HAP precipitate was filtered using Whatman filter paper (Grade 40), dried using an oven at 80ºC for a day and then finally ground to fine powder 16.

Synthesis of MnFe2O4 / HAP nanocomposite

Exactly 4g of ferric chloride and 1.5g of manganese acetate were dispersed in 100 mL distilled water to obtain a homogenous mixture. With continuous swirling, lye solution(2M) was supplemented drop by drop to the mixture to form MnFe2O4 nanoparticles in basic medium (pH=12) and the precipitate was washed with surplus amount of water until the pH turns 7. Thereafter HAP was disseminated into the solution (1:1 mass ratio) with continual stirring to yield a homogenous suspension and it was microwaved for 5min. The composite formed was separated by centrifugation, followed by washing with distilled water. Then it was heated in an oven at 80oC for 60 min and cooled. Finally, MnFe2O4/HAP nanocomposite was obtained 17.

 Scheme 1: Synthesis of MnFe2O4/HAP nanocomposite

Click here to View Figure

Characterization

PXRD spectrum of the as-synthesized nanocomposite materials was documented using PANalytical /X’pert3 powder diffractometer with λ =0.1541 nm Cu-Kα radiation. The surface morphology of the synthesized material was analysed using scanning electron microscope coupled with EDX analyser-FESEM JOEL JSMT300, operated at 10 kV. TEM images were visualized using HitachiH-7500 transmission electron microscope at an accelerating voltage of 200 kV. UV-vis reflectance data and functional group identification were obtained using Shimadzu-RF 6000 spectrometer and Perkin Elmer-RES-10 spectrum 400 FT-IR spectrometers. PL spectrum was recorded on a fluorescence spectrophotometer   at λext = 325 nm. Magnetic property was analysed on a SQUID-VSM DC magnetometer 18.

Photodegradation Evaluation

The photocatalytic ability of the as-synthesized HAP, MnFe2O4 nanomaterials, and MnFe2O4/HAP nanocomposite was examined using BV3 dye solution under the gleam of 30 W LED flashlight. For the discolouration work, fifty millilitres of 1×10-4 M experimental solution of BV3 dye was processed with 2 drops of H2O2 with different catalyst dosages 19. After irradiation, 3mL of dye solution were taken from the suspension and the λmax of the solution was monitored at 587 nm which is the λ max of BV3 dye solution 20. The extensiveness of decolourisation (DD) of the sample was determined by using the relation DD% = (Ai-At) / Ai x 100 where Ai is the samples’ absorbance at zeroth time and At is the absorbance at regular time intervals t.

Results and Discussion

Determination of crystalline nature by XRD

Powder XRD patterns are used to verify the crystalline nature and average particle size. The XRD peaks of HAP, MnFe2O4, and MnFe2O4/HAP nanocomposite photocatalysts are depicted in Fig.1F1(a-c). The XRD pattern of HAP, with the major peaks observed at 2theta are 25.30, 31.84, 32.00, 32.99, 34.12, 39.66, and 49.37 which correspond to (002), (211), (112) (300), (202), (130), and (213) planes of hydroxyapatite respectively. (Fig.1. F1(a). These angles obtained for HAP agree as per JCPDS File No:24-0033 and therefore the hexagonal crystalline structure of HAP was confirmed. The diffraction pattern of MnFe2O4 is shown in Fig.1.F1(b) and it catalyst displayed seven characteristic peaks at 2 equal to 18.02, 29.72, 35.34, 42.68, 53.10, 56.25, and 61.88 which corresponds to the (111), (220), (311), (400), (422), (511) and (440) hexagonal planes of face cantered cubic close packing structure respectively with a lattice constant of 8.470oA which matches exactly with the JCPDS Card No:88-1965 21 . These patterns confirm the spinel structure of MnFe2O4 magnetic core. Fig.1.F1(c) represents the diffraction pattern of MnFe2O4/HAP nanocomposite which incorporates the peaks of MnFe2O4 and HAP. No peaks due to impurities were seen and all the diffraction peaks related to MnFe2O4 and HAP were also seen in the diffraction pattern of MnFe2O4/HAP nanocomposite. These results suggest the coexistence of MnFe2O4 and HAP without any impurities and the spinel structure of MnFe2O4 magnetic core remains unaltered after the incorporation of synthesized nanoparticles over MnFe2O4/HAP which confirmed the protective coating of HAP on MnFe2O4 surface. The average crystalline size of MnFe2O4/HAP and MnFe2O4 nanoparticles were assessed to be 27.2 nm to 18.0 nm using Scherrer formula, D = 0.9/βcos θ 22, where λ is the wavelength of X-ray, θ the diffraction angle and β is FWHM.

Figure 1: F1. XRD patterns of the synthesized a) HAP, b) MnFe2O4 and c) MnFe2O4/HAP nanocomposite and F2. FT-IR spectra of a) MnFe2O4 b) HAP c) MnFe2O4/HAP nanocomposite

Click here to View Figure

FT-IR Spectral Studies

Fig.1.F2(a-c) portrays the FT-IR spectra of MnFe2O4, HAP and MnFe2O4/HAP nanocomposite. The presence of vibrational peaks (Fig.1.F2a) at 563 cm-1 and 411 cm-1 confirm Fe-O and Mn-O bonds ascertaining the successful synthesis of MnFe2O4 nanoparticles with tetrahedral and octahedral sites 23. Fig.1.F2b shows the FT-IR spectrum of HAP. The H-O-H stretching and bending vibrations at 3418 cm-1 and 1626 cm-1 indicate the appearance of adsorbed water molecules. The presence of CO32- group can be confirmed by the peaks at wavenumbers around 870 cm-1 and 1402-1460 cm-1 24. The bending vibration of O-P-O is the cause of the distinctive phosphate group vibrations of HAP viewed at 567 cm-1 and 605 cm-1. The distinct bands at 962 cm-1 and 1036 cm-1 can be corroborated to symmetric and anti-symmetric stretching vibrations of phosphate group, while the symmetric bending mode of the phosphate group vibration is responsible for the band at 472 cm-1. These results confirmed the formation of HAP 25.     

Fig.1.F2c represents the FT-IR spectrum of MnFe2O4/HAP. The stretching vibrations seen due to the presence of adsorbed water molecules at 3418 cm-1 and 1626 cm-1 in pure HAP and MnFe2O4 also appear also in MnFe2O4/HAP nanocomposite. The sharp FT-IR bands at 2365 cm-1 and 2369 cm-1 in MnFe2O4/HAP nanocomposite and in pure HAP are due to the presence of CO2 which could have come from the air. The distinctive stretching vibrations seen from 600 cm-1 to 400 cm-1 in MnFe2O4/HAP nanocomposite indicate the presence of spinel ferrite structure even after, anchoring MnFe2O4 on HAP matrix 26. The typical vibrational peak at 565 cm-1 accounts for the intrinsic vibration of MnFe2O4, which overlaps with the stretching vibrations of the phosphate moiety. The vibrational peaks of HAP and MnFe2O4 could be seen in MnFe2O4/HAP and the peak shift to the lower frequencies indicate the formation of the MnFe2O4/HAP nanocomposite requires less energy of vibration. All the results confirm the successful formation ofMnFe2O4/HAP nanocomposite 27.

Morphology and Elemental mapping studies

The microscopic images of MnFe2O4, HAP and MnFe2O4/HAP nanocomposite are shown in Fig.2 with various enhancement. It can be ascertained from the images that MnFe2O4 nanoparticles exhibit flaky like morphology having a diameter of around 100 nm, whereas HAP nanoparticles show rod shaped morphology having a diameter of 200 nm. The SEM imagesof MnFe2O4/HAP nanocomposite show that MnFe2O4 nanoparticles are embedded on a flaky sheet structure of HAP. A clear boundary was observed between MnFe2O4 and HAP sheet and all the small crystallites of MnFe2O4 nanoparticles were arranged randomly on the surface of HAP. Thus, the decorated MnFe2O4/HAP nanocomposite demonstrates that MnFe2O4 nanoparticles are well dispersed 28.

Figure 2: SEM photographs of (a, b) MnFe2O4 (c, d) HAP (e, f) MnFe2O4/HAP nanocomposite

Click here to View Figure

The TEM images confirm the size and distribution of MnFe2O4 nanoparticles (Fig. 3a) and MnFe2O4/HAP nanocomposite (Fig. 3b,c) synthesized via co-precipitation and microwave assisted co-precipitation method respectively. The results point out that the particles in the photocatalysts exhibit uniform morphology and distribution. The smallest particle size found for MnFe2O4 nanoparticles and MnFe2O4/HAP nanocomposite was 20 nm and 50 nm respectively.  The particle size was increased due to the incorporation of HAP. This suggests that HAP fuse with MnFe2O to increase the particle size of the MnFe2O4/HAP nanocomposite.  The chemical composition of MnFe2O4/HAP nanocomposite was identified using the results of EDX investigation and are displayed in Fig.3d, which authenticate the MnFe2O4/HAP nanocomposite is constituted with the elements Ca, P, Mn, Fe, C and O atoms. The signal of carbon may be due to the absorption of any organic molecules and thus confirming the absence of any other elemental impurities 29.  

Figure 3: (a) TEM images of MnFe2O4 nanoparticles (b, c)  MnFe2O4/HAP nanocomposite and (d)EDX spectrum of MnFe2O4/HAP nanocomposite

Click here to view Figure

UV-Visible Diffuse Reflectance Spectral Study

The MnFe2O4 and MnFe2O4/HAP nanocomposites exhibit strong absorptions from 250 nm to 1200 nm. The DRS scan spectra and the Tauc plot of MnFe2O4 and MnFe2O4/HAP are shown in Fig.4 (a, b). The optical absorption near the band edge can be calculated using the formula αhν = C[hν-Eg]n, where energy of the photon (a standard value), and the band gap energy are represented as C and Eg. The variable n is a value obtained for the electronic shift takes place during absorption. The distinct absorption band from 350 nm to 1050 nm proves that the absorption of visible light can be attributable to the transition takes place in band gap and not caused by absorption due to impurities. Based on the Tauc plot, the Eg values were calculated to be 1.7eV and 1.5eV for MnFe2O4 and MnFe2O4/HAP nanocomposite respectively. Owing to the captivating power of ferrites to take up visible light, the band gap values was found to decline from 1.7eV to 1.5eV for the MnFe2O4 anchoredHAP nanocomposite 30.

Figure 4: Diffuse Reflectance spectra and Tauc graph of a) MnFe2O4 b) MnFe2O4/HAP nanocomposite

Click here to View Figure

Photoluminescence

Photoluminescence spectrum of HAP, MnFe2O4 and MnFe2O4/HAP nanocomposite was recorded to understand the separation of electron-hole pairs. Fig.5 presents the photoluminescence sweep of HAP, MnFe2O4 and MnFe2O4/HAP nanocomposite. All the three samples were excited at 325nm with Ar+– ion laser. Bare HAP powder sample showed a broad band centred around 464nm and 520nm due to the presence of thermal activation defects 31. Under these condition MnFe2O4 and MnFe2O4/HAP nanocomposites did not show luminescence spectra, demonstrating ferrite-based samples promote the suppression of PL emission of HAP. But when compared to MnFe2O4 sample, the MnFe2O4/HAP showed a remarkable lowering of intensity of photoluminescence peak which signifies an effective transfer of charges over MnFe2O4/HAP nanocomposite matrix and suggesting a significant reduction in recombination ratio of charge carrier. The decrease in the electron-hole recombination process will certainly aid the effective schlepping of electron hole pairs on the surface of the heterojunction, simultaneously elevating photodegradation efficiency of the magnetic MnFe2O4/HAP nanocomposite.

Figure 5: Photoluminescence spectra: HAP, MnFe2O4 and MnFe2O4 anchored HAP nanocomposite

Click here to View Figure

VSM analysis

The magnetic properties of MnFe2O4 and MnFe2O4/HAP nanocomposite were evaluated using vibrating sample magnetometer in the range of -15000 to + 15000 Oe at 250C. Fig.6 (a, b) shows the plot of saturation magnetization Vs   applied magnetic field depicting a small hysteresis loop which denotes that both the samples exhibits typical ferromagnetic or soft magnetic nature. The saturation magnetization (Ms) of MnFe2O4 and MnFe2O4/HAP nanocomposite were 8.0 and 1.56 emu/g respectively. The decrease in saturation magnetization value was mainly due to the addition of non- magnetic chitosan material. Though there was a decrease in saturation magnetization value owing to the soft magnetic nature of the MnFe2O4/HAP composite material, the catalyst could be effectively removed with the use of external magnet from the treated BV3 dye solution.

Figure 6: Magnetic hysteresis curves: a) MnFe2O4 and b) MnFe2O4 anchored HAP nanocomposite

Click here to View Figure

XPS analysis

XPS analysis was carried out to spot the chemical constituents on the surface and electronic environment of MnFe2O4/HAP nanocomposite. The XPS investigation and high definition images of MnFe2O4/HAP nanocomposite are shown in Fig.7 and Fig.7(a-f). The results authenticate that the MnFe2O4/HAP nanocomposite has a polymeric surface comprising Mn 2p (531.3 eV), Fe 2p (284.9 eV), O 1s (642.2 eV), C 1s (710.2 eV), Ca 2p (347.1 eV) and P 2p (133.0 eV) assigned to the sample composition, confirming that MnFe2O4/HAP nanocomposite has all the expected elements. The observed C 1s spectrum (Fig.7a) of carbon might be due to the precursor acetate used in the synthesis process or air contamination. For MnFe2O4/HAP nanocomposite the high-resolution O 1s scan portrayed in Fig.7b can be accredited to the oxygen atom bound with manganese, iron, calcium and phosphorous atoms owing to the existence of HAP polymeric matrix. Evidently the sharp peaks cantered at 531.4 eV is caused by Fe-O/C-O bonds 33.

Fig.7c presents the high definition XPS spectrum of Fe 2p with binding energy at 710.2 eV and 725.9 eV pertained to the energy levels of Fe 2p3/2 and Fe 2p1/2 respectively, which are attributed to Fe3+ 34. The representative binding energy at 654.3 eV and 642.2 eV are allocated to the predictable doublets of Mn 2p1/2 and Mn 2p3/2 as seen in Fig.7d. The swapping of Mn for Fe in Fe3O4 crystals, for the formation of MnFe2O4 crystals was further verified by the results of XPS spectra of Fe2p and Mn2p characteristic peaks. The high definition XPS scan of Ca 2p for MnFe2O4/HAP nanocomposite exhibit duplet peaks appertained to the binding energy values 350.6 eV and 347.1 eV of Calcium 2p1/2 and Calcium 2p3/2 (Fig.7e) respectively. The peak viewed around 347.1 eV validate the existence of Ca 2p3/2 which in turn corroborate that the calcium atoms are tethered to PO43- groups 35. Eventually, the P 2p XPS high resolution scan demonstrated duplet peaks, with the energy levels for P 2p3/2 at 133.8 eV and for P 2p1/2 at 133.0 eV which could be correlated to the Phosphorous – Oxygen bond in hydroxyapatite (Fig.7f) 36. Therefore, based on the XPS analysis MnFe2O4/HAP nanocomposite has been favourably fabricated via the synthetic procedure followed in this study.

Figure 7: XPS survey spectrum of MnFe2O4/HAP nanocomposite, (a-f) High magnification XPS spectra of MnFe2O4/HAP nanocomposite.

Click here to View Figure

Photocatalytic degradation of BV3

The photodegradation ability of the as-fabricated MnFe2O4 and MnFe2O4/HAP nanocomposite was tested by degrading BV3 dye aqueous solution by the illumination of visible light. The BV3 dye degradation in the dark was minimal, ie., 11.25%, as depicted in Fig.8. It is obvious that visible light irradiation is required for the BV3 dye solution to undergo photodegradation. A degradation efficiency of 14.64% for BV3 dye was observed using LED flashlight light illumination, which substantiate that only minimum degradation was achieved where the mineralization reaction was exposed to LED light without any photocatalyst. Additionally, a second set of degradation of BV3 dye solution was conducted using the as-synthesized MnFe2O4 and MnFe2O4/HAP nanocomposite. After 210 min of visible light illumination, BV3 dye degradation was assessed to be 85.35 % and 93.24 % for MnFe2O4 and MnFe2O4 /HAP nanocomposite respectively 37. It is worthy note to mention thatMnFe2O4/HAP nanocomposite exhibited higher photocatalytic activity owing to adsorption and photodegradation processes.

Figure 8: Photocatalytic degradation of BV3 dye aqueous solution under Dark, Light, MnFe2O4, MnFe2O4/ HAP. [BV3] = 1×10-4 M, pH=7, Temp=34 ± 2oC

Click here to View Figure

The UV-vis spectral changes for BV3 dye degradation during the photocatalytic treatment with MnFe2O4 and MnFe2O4/HAP nanocomposite under visible light illumination are highlighted in Fig.9(a,b). The characteristic absorption band of BV3 dye at 587 nm is monitored at regular intervals of visible light illumination for the entire irradiation time of 210 min. The distinctive absorption peak of BV3 monitored around 587 nm gradually decreases and finally a flattened curve was seen after 210 min, making MnFe2O4/HAP nanocomposite, an effective photocatalyst for the elimination of BV3 dye (Fig.9b) 38. As there were no additional peaks formed due to the mineralization of BV3 dye, it is understandable that the BV3 dye has been mineralized absolutely.

Figure 9: Changes in absorption scan for the BV3 dye degradation with the use of a) MnFe2O4 b) MnFe2O4/HAP nanocomposite 

Click here to View Figure

The concentration of the catalytic material holds an indispensable role in generating the active species to react with the oxidants, which results in the better photodegradation rates. The consequence of the photocatalyst concentration on the quickness of photocatalytic degradation was assessed by adding 10 mg to 50 mg of catalysts namely MnFe2O4 and MnFe2O4/HAP nanocomposites into BV3 dye aqueous solution. As seen in Fig.9(c, d) the degradation has occurred only after the addition of the photocatalysts and the speed of degradation efficiency has increased with the increase in the catalyst concentration to the extent of 30 mg/50 mL and thereafter a decrease in rate was noticed, for an increase of catalyst concentration to 50 mg/mL39. According to Fig.9(c, d), the degradation efficiency MnFe2O4 and MnFe2O4/HAP showed 85.35 % and 93.24 % dye removal respectively after 210 min of irradiation using 30mg catalyst. The decrease in percentage degradation might be related to the dissemination of light and shielding effect by the cluttering of the higher catalyst concentration, causing opacity acting as an obstacle for visible light absorption, resulting in decreased photodegradation ability 40. Thus, the optimal catalytic dosage of 30 mg/mL is employed in this work.

The effect of visible light photodegradation of BV3 dye under four different initial dye concentrations, viz   1x 10-4 M, 2×10-4 M, 3×10-4 M, 4×10-4 M, and 5×10-4 M was evaluated (Fig.10a) with MnFe2O4/HAP nanocomposites. It is noted from Fig.10a, that the rate of photodegradation is decreased with increased concentration of the BV3 dye solution for both the catalysts. The reason is that, greater extent of BV3 dye molecules are adhering on the exterior surface of MnFe2O4/HAP nanocomposites that can prevent light absorption, thereby lowering the production of photogenerated charge carriers (e and h+). Hence, the degradation efficiency is more at low concentration of BV3 dye 41.

Figure 10: Influence of BV3 dye initial concentration on BV3 dye molecules degradation using a)MnFe2O4/HAP nanocomposite, b) Reusability of MnFe2O4/HAP nanocomposite

Click here to View Figure

Reusability cycles of MnFe2O4/HAP for BV3 degradation

The stability of the MnFe2O4/HAP nanocomposite is examined by performing five cycles of BV3 dye degradation under optimal conditions. The as-fabricated photocatalytic material was retrieved after each cycle, washed thrice, dried, and used for further cycles. The degradation efficiencies of the first, second, third, fourth, and fifth cycles were 75 %, 67%, 63%, 31%, and 22% respectively which are shown in Fig.10b.  The findings demonstrated that the degradation efficiency was good across three successive cycles, demonstrating the synthesized catalyst MnFe2O4/HAP is robust and reusable 42.

Kinetic Analysis

The kinetic Langmuir- Hinshelwood expression,  ln C0/Ct =kt is used to predict the photodegradation rate of BV3 dye solution, where k denotes the rate constant for photodegradation of BV3 dye, Co stand for the concentration of BV3 prior to irradiation, Ct stand for the concentration of BV3 at the radiance time, t. Fig.10(c), depicts a straight line for the kinetic plot of ln Co/Ct Vs time (t) for HAP, MnFe2O4 and MnFe2O4/HAP nanocomposite indicating that the photo mineralization of BV3 tracks pseudo first order kinetics with the rate constant values 3.7×10-2,6.5×10-2 and 7.0×10-2 min-1 respectively43. The L-H model thus shows that the degradation rate is profoundly controlled by the as-prepared MnFe2O4/HAP nanocomposite.

COD removal

Figure 11: COD reduction of BV3 dye solution [BV3] = 1×10-4M using 30 mg/50mL of catalyst at various irradiation time intervals.

Click here to View Figure

Chemical oxygen demand was evaluated for BV3 dye degradation in aqueous solution using the MnFe2O4/HAP nanocomposite by exposing dye solution to illumination. Fig.11 depicts the percentage removal of COD. The percentage removal of COD reached up to 76% at the end of 210 min, which confirmed the mineralization of BV3 dye 44. The higher COD removal can be attributed to the simultaneous adsorption and photocatalytic degradation of BV3 dye molecules on the surface of MnFe2O4/HAP nanocomposite. This could have caused a synergistic effect on the COD removal caused by the oxidation of BV3 molecules and not because of localized chemical transformations.

Degradation mechanism

The photocatalytic degradation of MnFe2O4/HAP nanocomposite is believed to occur based on the schematic diagram depicted in Fig.12. When photons from the visible light are absorbed by MnFe2O4/HAP, electron and hole pairs (e-h+) can be generated within MnFe2O4/HAP nanocomposite. In the photocatalytic degradation process, the electrons in the valence band are ejected to the conduction band upon irradiation with visible light and the photogenerated holes react with hydroxyl ions leading to the inception of reactive oxygen species i.e., hydroxyl radical, which in turn mineralizes the BV3 dye molecules to harmless products. From O2 and H2O molecules, superoxide anion radical (O2) and hydroxyl radical (OH) were produced by the interaction of the positive charge of the mesoporous holes with the negative charge of the electrons. In order to break down the adsorbed BV3 dye molecules into smaller molecules of CO2 and H2O, these active species of the MnFe2O4/HAP nanocomposite have formed a strong bond with the positive holes (h+). Moreover, the rapid mineralization of BV3 dye molecules was enabled by the enhanced adsorption capability of MnFe2O4/HAP nanocomposite. Furthermore, research has indicated that the greater area specificity of the MnFe2O4/HAP structure may be the cause of its potent adsorption capabilities 45

Figure 12: Plausible photocatalytic pathway of MnFe2O4/HAP nanocomposite

Click here to View Figure

As a result, MnFe2O4/HAP would support a solid foundation for its use in hazardous organic dye mineralization. An innocuous and eco-friendly beneficial method of bioremediation for upcoming generations certainly involve the extensive manufacture of MnFe2O4/HAP with a splendid mesoporous structure using biotic substrates like fish shells, fish bones and egg shells. HAP also possesses strong photocatalytic activity for eliminating azo dye from contaminated waste water. The as-fabricated photocatalysts’ enhanced photocatalytic activity is confirmed by the photocatalytic activity data, indicating that the degradation mechanism indicated by a sequence of Eqs. (1-6) is directly caused by this semiconductor ferrite-based nanocomposite.

Phyto toxicity analysis

Vigna radiatais a common bio-indicator utilized primarily for virulency studies. This plant can be easily cultured in laboratory because of having good sensitivity to toxicants. The toxicity of the control solution, treated BV3 dye solution and untreated BV3 dye solution towards the harvest of Vigna radiatato explore the harmfulness of BV3 dye solution after photodegradation reactions 46. Therefore, the seeds Vigna radiataplant were chosen and cultivated in treated BV3 dye solution, untreated BV3 dye solution and control sample (Cauvery water). The MnFe2O4/HAP nanocomposite photocatalysts was able to degrade 93% of BV3 dye solution. The photocatalytic degradation of the mixture is still incomplete as their COD removal is only 76% and it is noteworthy to mention that the intermediates formed due to dye degradation could be deadliest than the BV3 dye molecules. Henceforth, assessing the toxicity of the treated BV3 dye solution is therefore crucial and essential.  Fig.13 portrays   Vigna radiataplant imagesgrown in control sample, treated BV3 dye solution and untreated BV3 dye solution. The Vigna radiataplant grown in control sample (Cauvery water) was 9.6 cm long, whereas those grown in treated BV3 dye solution and untreated dye solution were about 8.2 cm and 3.4 cm long at the end of 15 days. Results from the treated BV3 dye solution and control sample’s growth trend were satisfactory. In comparison to the control sample, it is believed that the treated BV3 dye solution has minimum or no harmful effects. Nevertheless, Vigna radiatacultivated in untreated BV3 dye solution showed no discernible growth, and gradually, 40% of the infant plants dried up, possibly as a result of the BV3 dye molecules’ negative effects.

Figure 13: Digital images of Vigna radiata grown in control sample, treated and untreated BV3 dye solution at 25° C at the end of 15 days.

Click here to View Figure

Table 1: Phototoxicity results depicting the growth of Vigna radiata

Sample

Petiole Length (cm)

2nd Day

 6th Day

 8th Day

 12th Day

15th Day

Control Sample

2.1

4.5

5.4

6.5

9.6

Treated BV3 dye solution

2.5

3.7

4.9

5.8

8.2

Untreated BV3 dye solution

2.1

2.6

2.9

3.1

3.4

                            Number of branches

 2nd Day

 6th Day

 8th Day

12th Day

15th Day

Control Sample

3

6

8

9

15

Treated BV3 dye solution

3

5

6

9

13

Untreated BV3 dye solution

1

2

2

3

3

These results imply that the untreated BV3 dye solution has persistent toxic effects on the growth and survival of Vigna radiataplant, thus inhibiting its normal growth rate. The treated dye BV3 solution was analysed and was found to contain the non-toxic degradation products such as NH3, H2O and CO2 which was responsible for growth of the plant under study. From the phytotoxicity result, it can be concluded that the toxicity of the treated dye solution using photodegradation process was minimized to a larger extent. Hence, the treated BV3 dye solution was proved to be nontoxic to Vigna radiataplant, and can be used without further treatment for domestic and agriculture purposes.

Conclusion

In the present study, magnetically separable MnFe2O4/HAP nanocomposite was carefully prepared by microwave assisted co-precipitation process. Compared with the pure MnFe2O4 nanoparticles, MnFe2O4 /HAP nanocomposites exhibited higher photoactivity to degrade BV3 dye solution studied under the influence of light illumination. The as- synthesized photocatalytic materials were predicted by XRD, FT-IR, SEM, XPS, UV-DRS, PL and VSM analyses attributing to its structure and morphology. The XRD images authenticate cubic close packing spinel crystalline material with high degree of crystallinity. The characteristic absorption bands in FT-IR spectra confirm the formation of single phase MnFe2O4. The room temperature VSM analysis ensures the ferromagnetic nature of MnFe2O4 nanoparticles and MnFe2O4/HAP nanocomposite material. The photocatalytic competence of MnFe2O4/HAP nanocomposite to mineralize BV3 dye molecules under the influence of visible light illumination has become notable. Under the optimum conditions the MnFe2O4/HAP nanocomposite achieved a degradation of 93.24% at the end of 210 min of visible light illumination. The Langmuir-Hinshelwood kinetic plot could logically represent the rate of mineralization of BV3 dye molecules showcasing a surface phenomenon with pseudo-first order rate constants. The reusability study reveals that the MnFe2O4/HAP nanocomposite can be reused up to three cycles without any significant loss of photocatalytic activity. Phytotoxicity results demonstrated that the treated BV3 dye water has a very little toxic effect in the growth of Vigna radiata plant. The findings of this study provide valuable insights for utilizing MnFe2O4/HAP nanocomposite in the degradation of industrial effluents discharged specifically from textile and dyeing industries.

Acknowledgement

The authors sincerely acknowledge HAIF centre, Bishop Heber College, Tiruchirappalli, 620017, Tamil Nadu, India for the technical support.

Conflict of Interest

The content of this paper is the sole responsibility of the authors.

References

  1. Soliman A. M., Elsuccary S. A. A., Ali I. M., and Ayesh A. I. Journal of Water Process Engineering, 2017, 17, 245–255.
    CrossRef
  2. Jayasantha Kumari H., Krishnamoorthy P., Arumugam T. K., Radhakrishnan S., and Vasudevan D. International Journal of Biological Macromolecules, 2017, 96, 324–333.  
    CrossRef
  3. Reeta Mary I., Leethiyal R., Sekar P., Mangalaraj D., Viswanathan C., and Ponpandian N. Materials Today: Proceedings, 2019, 18, 1729–1734.
    CrossRef
  4. Hu Z. G., Zhang J., Chan W. L., and Szeto Y. S. Polymer, 2006, 47(16), 5838–5842.
    CrossRef
  5. Wojnárovits L., Pálfi T., and Takács E. Radiation Physics and Chemistry, 2007, 76(8–9), 1497–1501.
    CrossRef
  6. Hema E., Manikandan A., Karthika P., Antony S. A., and Venkatraman B. R. Journal of Superconductivity and Novel Magnetism, 2015, 28(8), 2539–2552.
    CrossRef
  7. Su L., Feng J., Zhou X., Ren C., Li H., and Chen X. Analytical Chemistry, 2012, 84(13), 5753–5758.
    CrossRef
  8. Ramli R. A., Adnan R., Bakar M. A., and Masudi S. A. M. J. Phys. Sci, 2011, 22(1), 20-37.
  9. Zarei Z., and Akhlaghinia B. RSC Advances, 2016, 6(108), 106473–106484.
    CrossRef
  10. Tedsree K., Temnuch N., Sriplai N., and Pinitsoontorn S. Materials Today: Proceedings, 2017, 4(5), 6576–6584.
    CrossRef
  11. Nguyen V. C. Advances in Natural Sciences: Nanoscience and Nanotechnology, 2014, 5(3), 035014.
    CrossRef
  12. Xia Y., and Yin L. Physical Chemistry Chemical Physics, 2013, 15(42), 18627.
    CrossRef
  13. Yang Z., Gong X., and Zhang C.Chemical Engineering Journal, 2010, 165(1), 117–121.
    CrossRef
  14. Jayalakshmi R., Jeyanthi J., Microchemical Journal, 2019, 145, 791–800.
    CrossRef
  15. Kafshgari L. A., Ghorbani M., and Azizi A. Particulate Science and Technology, 2018, 37(7), 904–910.
    CrossRef
  16. Joshi P., Patel C., and Vyas M. AIP Conference Proceedings. 2018, pubs.aip.org.
  17. Algethami J., Hassan M., Alorabi A., Alhemiary N., Fallatah A., Alnaam Y., Almusabi S., and Amna T. Nanomaterials, 2022,  12(10), 1631.
    CrossRef
  18. Guesmi Y., Agougui H., Lafi R., Jabli M., and Hafiane A. Journal of Molecular Liquids, 2018, 249, 912–920.
    CrossRef
  19. Kakavandi B., Takdastan A., Jaafarzadeh N., Azizi M., Mirzaei A., and Azari A. Journal of Photochemistry and Photobiology A: Chemistry, 2016,  314, 178–188.
    CrossRef
  20. Federation W. E., and Aph Association. American Public Health Association (APHA): 2005, Washington, DC, USA, 21.
  21. Liew K. H., Rocha M., Pereira C., Pires A. L., Pereira A. M., Yarmo M. A., Juan J. C., Yusop R. M., Peixoto A. F., and Freire C. Chem Cat Chem, 2017, 9(20), 3930–3941. Portico.
    CrossRef
  22. Bhowmik M., Kanmani M., Debnath A., and Saha B. Powder Technology, 2019, 354, 496–504.
    CrossRef
  23. Yang, Z., Jiang Y., Yu L. xin, Wen B., Li F., Sun S., and Hou T. Journal of Materials Chemistry, 2005, 15(18), 1807.
    CrossRef
  24. Wang S., Gao H., Li J., Wang Y., Chen C., Yu X., Tang S., Zhao X., Sun G., and Li. D Journal of Physics and Chemistry of Solids, 2021, 150, 109891.
    CrossRef
  25. Rocha J. H. G., Lemos A. F., Kannan S., Agathopoulos S., and Ferreira J. M. F. Journal of Materials Chemistry, 2005, 15(47), 5007-5011.
    CrossRef
  26. Ghobadi M., Gharabaghi M., Abdollahi H., Boroumand Z., and Moradian M. Journal of Hazardous Materials, 2018, 351, 308–316.
    CrossRef
  27. Zhou Y., Xiao B., Liu S.-Q., Meng Z., Chen Z.-G., Zou C.-Y., Liu C.-B., Chen F., and Zhou X. Chemical Engineering Journal, 2016, 283, 266–275.
    CrossRef
  28. Das K. Ch., Das B., and Dhar S. S. Water, Air, & Soil Pollution, 2020, 231(2).
    CrossRef
  29. Paris E. C., Malafatti J. O. D., Moreira A. J., Santos L. C., Sciena C. R., Zenatti A., Escote M. T., Mastelaro V. R., and Joya M. R. Environmental Science and Pollution Research, 2022, 29(27), 41505–41519.
    CrossRef
  30. Tatarchuk T. R., Paliychuk N. D., Bououdina M., Al-Najar B., Pacia M., Macyk W., and Shyichuk A. Journal of Alloys and Compounds, 2018, 731, 1256–1266.
    CrossRef
  31. André R. S., Paris E. C., Gurgel M. F. C., Rosa I. L. V., Paiva-Santos C. O., Li M. S., Varela J. A., and Longo E. Journal of Alloys and Compounds, 2012, 531, 50–54.
    CrossRef
  32. Kafshgari L. A., Ghorbani M., and Azizi A. Applied Surface Science, 2017, 419, 70–83.
    CrossRef
  33. Lee D., Kwon O.-S., and Song S. H. RSC Adv., 2017, 7(72), 45595–45600.
    CrossRef
  34. Yamashita T., and Hayes P. Applied Surface Science, 2008, 254(8), 2441–2449.
    CrossRef
  35. Negrila C., Predoi M., Iconaru S., and Predoi D. Molecules, 2018, 23(11), 2986.
    CrossRef
  36. Nelson A. E., Hildebrand N. K. S., and Major P. W. Surface Science Spectra, 2002, 9(1), 250–259.
    CrossRef
  37. Chen Y., Huang W., He D., Situ Y., and Huang H. ACS Applied Materials & Interfaces, 2014, 6(16), 14405–14414.
    CrossRef
  38. Abroshan E., Farhadi S., and Zabardasti A. Solar Energy Materials and Solar Cells, 2018,  178, 154–163.
    CrossRef
  39. Hou P., Shi C., Wu L., and Hou X. Microchemical Journal, 2016, 128, 218–225.
    CrossRef
  40. Valizadeh S., Rasoulifard M. H., and Dorraji M.S.S. Applied Surface Science, 2014,319, 358–366.
    CrossRef
  41. Pare B., Jonnalagadda S. B., Tomar H., Singh P., and Bhagwat V. W. Desalination, 2008,232(1–3), 80–90.
    CrossRef
  42. Zhang F., Yin X., Zhang W., and Ji Y. Journal of the Taiwan Institute of Chemical Engineers, 2016, 65, 269–275.
    CrossRef
  43. Shariffuddin J. H., Jones M. I., and Patterson D. A. Chemical Engineering Research and Design, 2013, 91(9), 1693–1704.
    CrossRef
  44. Ertugay, N., & Acar, F. N.  Arabian Journal of Chemistry, 2017, 10, S1158–S1163.
    CrossRef
  45. Chai Y., Ding J., Wang L., Liu Q., Ren J., and Dai W.-L. Applied Catalysis B: Environmental, 2015, 179, 29–36.
    CrossRef
  46. Garcia J. C., de Souza Freitas T. K. F., Palácio S. M., Ambrósio E., Souza M. T. F., Santos L. B., de Cinque Almeida V., and de Souza N. E. Environmental Monitoring and Assessment, 2012, 185(3), 2179–2187.
    CrossRef

Creative Commons License
This work is licensed under a Creative Commons Attribution 4.0 International License.

About The Author