192
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Synthesis of TiO2 nanoparticles using red spinach leaf extract (Amaranthus Tricolor L.) for photocatalytic of methylene blue degradation

, ORCID Icon, ORCID Icon, , ORCID Icon, , ORCID Icon, ORCID Icon, ORCID Icon & ORCID Icon show all
Article: 2352571 | Received 16 Feb 2024, Accepted 03 May 2024, Published online: 16 May 2024

ABSTRACT

Textile industry waste such as methylene blue can pose a serious threat to health and ecosystems. One solution to overcome this problem is to utilize photocatalytic technology using TiO2 semiconductors. This research aims to make green synthetic TiO2 nanoparticles (NPs) using red spinach (Amaranthus Tricolor L.) leaf extract and used for photodegradation of methylene blue. Extraction of red spinach leaves showed that 30 min and a temperature of 95°C showed the highest phenolic content. In this research, several phenolic compounds were also identified in red spinach extract. For TiO2 NPs, the X-ray diffraction results show that TiO2 has a pure anatase phase. Electron microscope results show that TiO2 has the shape of agglomerated spherical clusters. The particle size distribution shows that TiO2 has an average size range of 77–90 nm. The methylene blue photodegradation showed that TiO2 had an efficiency of 90.36% for 180 min, higher than commercial TiO2 Evonic. This result can be attributed to the small particle size and broadband spectra so that photocatalysis is more effective.

GRAPHICAL ABSTRACT

1. Introduction

Industrial production and use of synthetic dyes for textile coloring has become a giant industry today (Citation1). In Indonesia, textile industry is one of the sectors with rapid growth, reaching 15.6% in 2022 according to data from the China Research and Intelligence (CRI) (Citation2). Dye waste generated from the textile industry poses environmental problems. Its toxic nature has become a serious concern for the environment. The use of synthetic dyes harms all forms of life which are usually not biodegradable (Citation3). Therefore, further treatment of textile industry waste is needed before it reaches final disposal. Dyes are among the organic compounds with complex aromatic structures, making them stable and resistant to degradation (Citation4). Methylene blue is one of the azo dyes commonly used by industries, and it can be harmful to the environment (Citation5). Hence, there is a need for environmentally friendly and effective methods to treat wastewater from the textile industry.

Numerous wastewater treatment methods have been developed, but they often raise new environmental issues (Citation6). An alternative method that can be used is the combination of efficient photocatalysis principles for wastewater purification. In photocatalysis, dyes are broken down into simpler components, making them environmentally safe (Citation7). Nanomaterials semiconductor such as ZnO, MnO2, and TiO2 are efficient for various reactions due to their large surface area and high surface activity (Citation8). Nanotechnology can be applied in photocatalysis methods by synthesizing TiO2 NPs. TiO2 NPs-based photocatalysis are favored because nano-sized TiO2 (ranging from 1–100 nm) exhibits superior chemical and physical properties compared to larger materials (Citation9). TiO2 also have better UV light activity, better stability, lower reaction time and wider band gap (3.2 eV). Wider band gap in TiO2 accelerate photo charge recombination process making it stand-out among other metal oxide NPs (Citation10).

Synthesis of TiO2 nanoparticles (NPs) can be carried out in various ways, such as the sol–gel, hydrothermal, polyol synthesis, and precipitation (Citation11). The sol–gel method uses a low-temperature process. However, if organic reagents are used during the experiment, the final product will contain a lot of carbon. To synthesize metal oxides directly from the solution, hydrothermal methods have been used. However, most previous studies on the synthesis of TiO2 NPs via the hydrothermal method required high temperatures, which resulted in the formation of polydisperse powders (Citation12). Until now, not much research has been carried out regarding the synthesis of TiO2 NPs through a precipitation process using environmentally friendly chemical methods.

Green synthesis is environmentally friendly, safe, and cost-effective. It involves using plant extracts as capping agents in nanoparticle synthesis (Citation13,Citation14). Green synthesis with plant extracts utilizes organic compounds like secondary metabolites such as flavonoids, triterpenoids, polyphenols, and quercetin found in plants to reduce metal ions (Citation15–17). In the presence of plant extract and suitable conditions, such as concentration and temperature, metal nanoparticles can be synthesized and the quality can even surpass those that use chemical method (Citation18). Previous research used extracts from various plants, including white snakeroot (Ageratina altissima), ginger (Curcuma longa), Calotropis gigantea, and Aloe vera as capping agents for synthesis TiO2 NPs (Citation19–22). In other research, TiO2 NPs were synthesized using Averrhoa carambola and Nervilia aragoana leaf extracts which were used as a photocatalyst for the degradation of textile wastewater and as an antibacterial (Citation23,Citation24). TiO2 NPs using the help of the extract have good properties such as smaller size compared to chemical methods, thereby having a better ability of photocatalytic activity (Citation10).

In this research, TiO2 NPs is synthesizes using red spinach (Amaranthus Tricolor L.) leaf extract for photocatalytic degradation of methylene blue. Red spinach leaves were chosen because they contain higher levels of secondary metabolites such as flavonoids and polyphenols than green spinach leaves (Citation25). TiO2 NPs were synthesized using titanium tetraisopropoxide as a titanium source, water as a solvent, and red spinach extract as a template. The effect of varying extract concentration and precursor concentration on TiO2 NPs and their potential in photocatalytic applications for methylene blue degradation compare to commercial TiO2 were also explored.

2. Materials and methods

2.1 Materials

Red spinach (Amaranthus Tricolor L.) was obtained from a local market in Bandung, Indonesia. Red spinach leaves were separated from their stems, finely chopped, and then dried at 45°C for 48 h before being pulverized. The materials used in this study include distilled water, gallic acid (C7H6O5, Merck), absolute ethanol (C2H5OH, 99%, Merck), Folin–Ciocalteu reagent (Merck), methylene blue (Merck), sodium hydroxide (NaOH, Merck), hydrochloric acid (HCl, 37%, Merch), sodium carbonate (Na2CO3, Merck), and titanium tetraisopropoxide (TTIP 97%, Sigma-Aldrich). All the materials used were of analytical grade.

2.2 Preparation of gallic acid standard

Gallic acid (0.05 grams) was dissolved in 50 mL of methanol and homogenized. Gallic acid was prepared in several standard variations, namely 30, 50, 100, 300, 500, and 750 mg L−1. A 50 µL gallic acid 300 mg L−1 was added to 2.5 mL of 10% Folin–Ciocalteu reagent and 2 mL of 7.5% Na2CO3 solution. The solution was incubated at 45°C for 15 min, and the absorbance was measured at maximum wavelength.

2.3 Determination of polyphenol content in red spinach leaf extract

Red spinach leaves (3.6 g) pulverized were added to 100 mL of distilled water. They were then heated at 50°C for 30, 60, 90, and 120 min, filtered, and concentrated. Several grams of concentrated extract were then dissolved in 1 mL of methanol and centrifuged. Next, 50 µL of the solution was added to 2.5 mL of 10% Folin–Ciocalteu reagent, and 2 mL of 7.5% Na2CO3 solution was added. The solution was incubated at 45°C for 15 min, and the absorbance was measured at the maximum wavelength. The polyphenol content was then calculated. Once the time yielding the highest polyphenol content was determined, the temperature that produces the highest polyphenol content was sought using the same method at temperatures of 20, 40, 60, 80, and 95°C. The selected extract was then characterized using gas chromatography-mass spectrometry (GC-MS, Asquisition 10.0.368). The total polyphenol content was calculated using Equation (1), where TPC is the total phenolic content (mg GAE g−1), C is the concentration of tea leaves (mg L−1), V represents the volume of solvent (L), and m is the weight of the tea extract used (g). (1) TPC=(C×V)/m(1)

2.4 Synthesis and characterization of TiO2 NPs

A total of 2 mL of TTIP was taken and added to 30 mL of ethanol, which was stirred for 1 min. Subsequently, red spinach leaf extract was added in volume ratios TTIP:red spinach leaf extract of 1:3, 1:5, 1:7, 1:10, and 1:15, and the mixture was stirred for 3 h. The solution was then centrifuged for 10 min, and the precipitate was collected. The precipitate was washed three times with distilled water and ethanol, followed by centrifugation for 10 min. Next, the precipitate was dried in an oven for 18 h. The dried precipitate was then calcined at a temperature of 500°C for 3 h. Synthesized TiO2 NPs for a volume ratio of TTIP:red spinach leaf extract of 1:3, 1:5, 1:7, 1:10, and 1:15 is named as TiO2-3, TiO2-5, TiO2-7, TiO2-10, and TiO2-15 respectively.

To determine the crystal structure, phase composition, and crystallite size, X-ray diffraction (XRD, Rigaku/miniflex 600, Tokyo, Japan) was used at room temperature using Cu Kα emission (wavelength λ = 0.15418 nm) and scanning was carried out in the range 20–70° (2θ Bragg) with a scanning speed of 10° min−1. The resulting data was analyzed using the Rietveld method with HighScore Plus software (PANalytical) to determine the crystal structure and phase composition (Citation26). The crystal size of TiO2 NPs was calculated using the Scherrer equation. To determine the light absorption and band gap energy properties, UV-vis diffuse reflectance spectroscopy (UV-Vis-DRS, Jasco V-550, Tokyo, Japan) was used from the range 220–800 nm. To determine the functional groups contained in the TiO2 NPs, Fourier-transform infrared (FTIR, Perkin Elmer Spectrum 100) characterization was used in the range 500–4000 cm−1 from the average of 64 times scanning. To determine the thermal properties thermal gravimetric-differential thermal analysis (TG-DTA, Shimadzu, Japan) was used at a temperature of 20–900°C with a heating rate of 10°C/min and heated in air. To determine the morphology, size, and elements contained in the compound, scanning electron microscopy-energy dispersive X-ray spectroscopy (SEM-EDS, Tabletop Microscope-1000, Hitachi, Tokyo, Japan) and transmission electron microscope (Talos F200X G2 TEM, Thermo Fisher Scientific, Tokyo, Japan) was used.

2.5 Photocatalytic activity test

The photocatalysis activity was analyzed using a methylene blue dye solution. A 100 mL of methylene blue 10 mg L−1 was prepared, and 0.1 g of synthesized TiO2 nanoparticles was added. The pH was adjusted using 0.1 M HCl and NaOH. The solution was stirred at room temperature for 60 min for reach adsorption–desorption and 180 min for UV-B irradiation (Exo Terra Reptile UVB150), with samples taken every 30 min. The precipitate and solution were separated by centrifugation. The absorbance at the maximum wavelength was measured using UV-vis spectroscopy.

3. Result and discussion

3.1 Analysis of secondary metabolite content in red spinach leaf extract

3.1.1 Polyphenol content

In this experiment, the part of red spinach used was the leaf. In the extraction process, distilled water was used with the aim of reducing organic solvents in the extraction process and having the same polarity as phenolic compounds (Citation27). In the determination of polyphenols, the principle of the Folin–Ciocalteu method is that when Folin–Ciocalteu reacts with phenolic compounds originating from polyphenols (Fig. S1). The reaction between the Folin–Ciocalteu reagent and polyphenols in the solution can be seen in Fig. S2 (Citation28).

The maximum polyphenol content was obtained by heating the red spinach leaf powder at various time intervals at a temperature of 50°C. (a) shows that the maximum total polyphenol content (TPC) was 2.37 mg GAE g−1 for 30 min. After 30 min, there was a decrease in polyphenols. This can happen because the polyphenols in red spinach will decompose, thereby reducing the polyphenol content obtained. After determining the optimal time, heating was performed at various temperatures for 30 min to find the best temperature. (b) shows that the highest polyphenol content in the red spinach leaf extract was at 95°C, which was 1.59 mg GAE g−1. At this temperature the polyphenol content is highest because at high temperatures the polyphenols released are high because the solubility of phenolic compounds is greater. Therefore, the optimum time and temperature extraction was carried out at 95°C for 30 min.

Figure 1. Total polyphenol content (TPC) for (a) various of time at temperature of 50°C and (b) various of temperature for 30 min.

Figure 1. Total polyphenol content (TPC) for (a) various of time at temperature of 50°C and (b) various of temperature for 30 min.

3.1.2 Secondary metabolites compounds

To determine the secondary metabolites compounds in the red spinach leaf extract, GC/MS analysis was performed. The analysis results using GC/MS can be seen in and Fig. S3, which shows the identification of 8 compounds in the red spinach leaf extract. Based on the results, the compounds that play a role in the synthesis of TiO2 nanoparticles are identified. 2-amino-5-methyl benzoic acid is a derivative of anthranilic acid (Citation29). 3-buten-2-ol is an organic compound containing hydroxyl groups and a double bond as an alkene. 2-(3,4-dimethoxyphenyl)−5-hydroxy-6,7,8-trimethoxy-4H-chromen-4-one is polymethoxyflavone, and 3’,4’,5,6,7,8-hexamethoxyflavone is flavonoid group (Citation30). Polymethoxyflavones and flavonoids are polyphenol category (Citation31). As for 4,4'-bi-4H-pyran, 2,2’,6,6'-tetrakis(1,1-dimethylethyl)−4,4'-dimethyl is an organic compound with 15 carbons, which is a cyclic sesquiterpene alcohol (Citation32). β-Amyrin is a chemical compound belonging to the triterpene class (Citation33). Based on previous research, red spinach leaves contain secondary metabolites including anthocyanins, flavonoids, tannins, saponins, and squalene (Citation34). This result is also similar to Fatimah and Aftrid’s research which shows that in spinach leaves there are at least 13 types of flavonoids and groups of phenolic compounds in spinach, namely gallic acid, caffeic acid, routine, ferulic acid, and quercetin (Citation35).

Table 1. Compounds detected in the red spinach leaf extract.

3.2 Properties of TiO2 NPs

The XRD pattern of the TiO2 NPs are presented in (a). The XRD pattern of all synthesized TiO2 NPs closely matches the anatase Inorganic Crystal Structure Database (ICSD) reference code 98-002-4276, indicating that all samples are pure anatase and adopt a tetragonal structure. The peaks obtained at 2θ = 25.3°, 37.9°, 48.4°, 53.9°, 55.3°, and 62.7° are equivalent to the (011), (004), (020), (015), (121), and (024) planes (Citation36). The lattice parameters of TiO2 NPs were determined, representing the physical dimensions of the unit cell in the crystal lattice. shows the lattice parameters of the synthesized TiO2 NPs. Differences in extract volume variations did not significantly change the lattice parameters. However, there was a slight change in the a and b-axis which slightly increased with the amount of extract in the synthesis stage. On the other hand, the c-axis slightly decreases with the amount of extract, indicating the distortion of the crystal of green synthesis. To confirm the accuracy of the Rietveld refinement calculations, the goodness of fit (GoF) value is calculated. GoF is a statistical measure that assesses how well the experimental results match a set of observations (Citation37). The GoF values from all samples are below 4 and indicate the accuracy of the calculations. The crystallite size of TiO2 NPs after calcination is shown in , which indicates that the larger the amount of extract used, the larger the crystal size obtained. These results are smaller than in the TiO2 synthesis using a commercial capping agent polyethylene glycol (PEG), ranging from 22–31 nm (Citation38), which concluded that green synthesis produces smaller crystallite sizes.

Figure 2. (a) XRD pattern, (b) UV-vis spectra, (c) FTIR spectra, and (d) TGA graph of TiO2 NPs.

Figure 2. (a) XRD pattern, (b) UV-vis spectra, (c) FTIR spectra, and (d) TGA graph of TiO2 NPs.

Table 2. Crystal properties of TiO2 NPs.

The optical properties of the samples were investigated by diffuse reflectance UV – vis spectroscopy. The absorption spectra in the range of 220–800 nm of the synthesized TiO2 NPs is shown in (b). In general, all TiO2 absorbs light in the UV range (220–400 nm). However, there is a slight difference in absorption intensity at 325 nm. This difference in absorption is likely due to differences in crystal lattice parameters (Citation39). Band gap energy from TiO2 NPs is calculated using the Davis – Mott plot for indirect band gap (Fig. S4). All samples’ band gap energy values are around 2.98–3.00 eV. These results show that all variations of TiO2 NPs do not show significant differences in band-gap values.

FTIR analysis was carried out to identify the functional groups of the synthesized TiO2 NPs shown in (c). The FTIR spectra of TiO2 NPs shows a broad and strong absorption band in the wavenumber range 505–530 cm−1, where this absorption band indicates the vibration of the O−Ti−O bond, thus confirming that the synthesized nanoparticles are TiO2 compounds (Citation40). In addition, there is an absorption band at the wavenumber 2350 cm−1 which can be caused by saturated carbon dioxide bonds adsorbed on the catalyst (Citation41). These results indicate that the TiO2 NPs synthesized are pure and no other bonds are formed.

In this FTIR data, secondary metabolite compounds from red spinach leaf extract are no longer present in TiO2 NPs. This is because secondary metabolite compounds disappear during the calcination process at a temperature of 500°C. This result is proven by TGA data on TiO2 NPs before calcination (Fig. S5). The mass of TiO2 NPs decreases drastically at a temperature of 50–500°C which is probably due to the disappearance of secondary metabolite compounds from red spinach leaf extract. Mass loss in TiO2-3 before calcination to a temperature of 500°C was 25.6%, while in TiO2-15 it was 28.7%. This shows that the greater the volume ratio of red spinach leaf extract used during the synthesis process, the more metabolites are contained in TiO2 NPs.

(d) shows the TGA graph of TiO2 NPs after calcination. TiO2 experiences surface water loss up to 200°C. After that, weight loss occurred at a temperature of 200–400°C mainly due to evaporation, partial dehydration of structural water, or due to the presence of some impurities from red spinach extract (Citation42). There are slight differences between TiO2 variations. TiO2-3 appears to have more water contained in the surface and structural water than TiO2-15. This may also be related to particle size (see next section particle size distribution). TiO2-3 has a smaller particle size distribution so it is possible to absorb more water. After that, there was an increase in mass caused by the oxidation of TiO2. The result of this oxidation is usually a TiO2 phase transformation from anatase to rutile which has a different density (Citation43). Fig. S6 shows the TG-DTA curve of TiO2-3 and TiO2-15. Both show exothermic reactions during the heating of TiO2 to 900°C.

(a) shows the SEM image of TiO2-3. TiO2 NPs show an agglomerated spherical particle shape. In addition, other variations of TiO2 NPs are shown in Fig. S7a-e. All variations of TiO2 NPs have a homogeneous round particle shape. Additionally, EDS mapping confirmed the high purity of the TiO2 sample, which contains only titanium and oxygen atoms (Fig. S7f). The percentage of Ti and O atoms from the EDS results is shown in Table S1. Further characterization was carried out using TEM (b) to observe TiO2-3 more closely which showed the presence of a spherical structure. It is known that the particle size ranges from 90–100 nm.

Figure 3. (a) SEM and (b) TEM image of TiO2-3.

Figure 3. (a) SEM and (b) TEM image of TiO2-3.

The particle size distribution of TiO2 NPs is presented in . The smallest average particle size distribution of TiO2 NPs is TiO2-3, namely 77 ± 21 nm. The particle size distribution increased with the amount of extract added. This shows that if too much extract is added, it will make making nanoparticles ineffective. Apart from that, TiO2-3 also has relatively high homogeneity compared to other variations, which shows that excess extract results in the growth of nanoparticles not being evenly distributed.

Figure 4. Particle size distribution of (a) TiO2-3, (b) TiO2-5, (c) TiO2-7, (d) TiO2-10, and (e) TiO2-15.

Figure 4. Particle size distribution of (a) TiO2-3, (b) TiO2-5, (c) TiO2-7, (d) TiO2-10, and (e) TiO2-15.

In the process of forming TiO2 NPs, red spinach leaf extract is used as a capping agent or stabilizing agent to stabilize the nanostructure (Citation44). Commercially, commonly used capping agents include surfactants (cetyltrimethylammonium bromide (CTAB) and sodium dodecyl sulfate (SDS)), polymers (polyvinylpyrrolidone (PVP), polyethylene glycol (PEG), and poly-α, γ, L-glutamic acid (PGA)), and thiols (dodecanethiol and thioglycerol) (Citation45,Citation46). In red spinach leaf extract, secondary metabolites act as capping agents or stabilizers by coating the TiO2 NPs formed, preventing agglomeration, reducing particle size, and forming a stable and homogeneous structure. The mechanism of the synthesis process that occurs in TiO2 NPs can be seen in . Functional groups contained in secondary metabolites are attached to the outer surface of TiO2 nanoparticles through covalent or hydrogen bond interactions. As a result, the hydrophilic structure of secondary metabolites sticks to the TiO2 NPs, while the hydrophobic structure remains outside (Citation47,Citation48).

Figure 5. Mechanism of formation green synthesis TiO2 NPs with red spinach leaf extract.

Figure 5. Mechanism of formation green synthesis TiO2 NPs with red spinach leaf extract.

3.3 Photocatalytic activity of TiO2 NPs

In this research, methylene blue was used as a photocatalysis target because it is a common compound that can be found in industrial waste (Citation49). Fig. S8 shows the maximum wavelength and standard curve of methylene blue. Before testing, photolysis was carried out in the absence of a photocatalyst. As shown in (a), no activity was observed for photolysis for methylene blue degradation. When TiO2 NPs is added, degradation occurs which indicates photocatalytic activity. TiO2-3 has the highest activity with a degradation efficiency of 90.36% after 180 min. The Langmuir – Hinshelwood first-order kinetic model is shown in (b). Green synthesized TiO2 NPs showed better reaction rate kinetics for all variations compared to P25 Evonic (commercial), with the highest being TiO2-3. This is due to the properties described previously, that TiO2-3 has a smaller average particle size, allowing it to have a larger surface area. This large area causes photocatalysis reactions on the surface to occur efficiently. shows several previous reports of the photocatalytic activities of TiO2 synthesized using plant extracts, clearly showing that our data on photoactivity are compared with those published in the literature.

Figure 6. (a) Photocatalytic activity, (b) kinetics value for all TiO2 NPs, and (c) photocatalytic activity of TiO2-3 on different pH for methylene blue degradation.

Figure 6. (a) Photocatalytic activity, (b) kinetics value for all TiO2 NPs, and (c) photocatalytic activity of TiO2-3 on different pH for methylene blue degradation.

Table 3. Comparison of the photocatalytic activity with several previous reports of TiO2 synthesized using plant extracts.

When TiO2 is exposed to UV light, electrons (e ) are excited from the valence band (VB) to the conduction band (CB) and leave holes (h+) in the VB. Electron accumulation in CB reduces O2 to ⋅O2 (Citation36). In addition, h+ accumulation can develop ⋅OH because the VB position is more positive than the OH/⋅OH potential (+2.38 eV vs. NHE) (Citation55). These ⋅OH and ⋅O2 species can degrade MB in solution. Meanwhile, photogenerated h+ due to electron transitions from the valence band (VB) has strong oxidation capabilities and can directly degrade MB to produce less harmful products such as water and carbon dioxide (Citation56).

Photocatalytic activity can be influenced by several factors including band gap energy, size, morphology, recombination rate, pH, and surface area of the material (Citation57). The difference between pH 7 and pH 10 is shown in (c). An initial pH of 10 showed better methylene blue degradation efficiency than pH 7. The properties of methylene blue and TiO2 may explain this. The TiO2 photocatalyst has a point of zero charge (pHPZC) of 6.25. As a result, the surface is mostly negatively charged over pHPZC (TiOH + OH ↔ TiO  + H2O) (Citation36). In addition, methylene blue, a cationic dye, is almost completely cationic (Citation58). Thus, methylene blue and TiO2 show an attractive effect in alkaline pH, favoring the adsorption ability. The occurrence of electrostatic interactions is useful for improving adsorption properties thereby increasing the efficiency of concentration reduction. Meanwhile, at low pH, the efficiency of concentration reduction tends to be low due to electrostatic repulsion between TiO2 NPs and methylene blue dye (Citation59).

4. Conclusions

TiO2 NPs were successfully made by green synthesis using red spinach leaf extract. Red spinach extract contains many different polyphenolic compounds, which can help the formation of nanoparticles. The extraction process of red spinach leaves shows that the highest amount of polyphenol is within thirty min and at a temperature of 95°C. The crystallite size of the resulting TiO2 NPs ranges from 10–16 nm and has a pure anatase structure. The presence of O – Ti – O groups and CO2 bonds that can be adsorbed on TiO2 NPs was discovered through FTIR analysis. The agglomerated and spherical shape of the TiO2 synthesis results is shown in the SEM image. Methylene blue (MB) as a dye was used to check the photocatalytic ability of the prepared TiO2 NPs. The results showed that the sample synthesized using a volume ratio of TTIP and extract of 1:3 had a smaller particle size (77 nm), contributing to a degradation efficiency of 90.36% after two hours of ultraviolet irradiation. The method developed in this research is fast and environmentally friendly for producing TiO2 NPs with good photocatalytic activity.

Supplemental material

Supplemental Material

Download ()

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

This work was supported by [Universitas Padjadjaran] Grant Hibah Riset Unpad (HRU) by Riset Kompetensi Dosen Unpad (RKDU) the Academic Leadership Grant (ALG) under Grant the Academic Leadership Grant (ALG), Prof. Iman Rahayu [number 1764/UN6.3.1/PT.00/2024] and [Universitas Padjadjaran] under Grant Hibah Riset Unpad (HRU) by Riset Kompetensi Dosen Unpad (RKDU), Diana Rakhmawaty Eddy [number 1764/ UN6.3.1/PT.00/2024]. The authors extend their appreciation to the Deanship of Scientific Research at Northern Border University, Arar, KSA for funding this research “work through the project number “NBU-FFR-2024-540-04”.

References

  • Prasetyani, D.; Abidin, A.Z.; Purusa, N.A.; Sandra, F.A. The Prospects and the Competitiveness of Textile Commodities and Indonesian Textile Product in the Global Market. Etikonomi 2020, 19, 1–18. doi:10.15408/etk.v19i1.12886
  • China Research and Intelligence. Indonesia Garment Manufacturing Industry Research Report 2023-2032, (2023) CRI Publisher. https://www.giiresearch.com/report/cri1339032-indonesia-garment-manufacturing-industry-research.html.
  • Khattab, T.A.; Abdelrahman, M.S.; Rehan, M. Textile Dyeing Industry: Environmental Impacts and Remediation. Environ Sci Pollut Res 2020, 27, 3803–3818. doi:10.1007/s11356-019-07137-z
  • Daneshvar, N.; Ayazloo, M.; Khataee, A.R.; Pourhassan, M. Biological Decolorization of dye Solution Containing Malachite Green by Microalgae Cosmarium sp. Bioresour. Technol. 2007, 98, 1176–1182. doi:10.1016/j.biortech.2006.05.025
  • Khan, I.; Saeed, K.; Zekker, I.; Zhang, B.; Hendi, A.H.; Ahmad, A.; Ahmad, S.; Zada, N.; Ahmad, H.; Shah, L.A.; Shah, T. Review on Methylene Blue: Its Properties, Uses, Toxicity and Photodegradation. Water. (Basel) 2022, 14, 242. doi:10.3390/w14020242
  • Utubira, Y.; Wijaya, K.; Triyono, T.; Sugiharto, E. Preparation and Characterization of TiO2;-Zeolite and its Application to Degrade Textille Wastewater by Photocatalytic Method. Ind J Chem 2006, 6, 231–237. doi:10.22146/ijc.21724
  • Saeed, M.; Muneer, M.; Haq, A.U.; Akram, N. Pesticide and Agro-Ecological Transition: Assessing the Environmental and Human Impacts of Pesticides and Limiting Their use. Environ Sci Pollut Res 2022, 29, 1–5. https://doi.org/10.1007/s11356-021-16389-7.
  • Roy, P.; Ho, L.; Periasamy, A.P.; Lin, Y.; Huang, M.; Chang, H. Graphene-ZnO-Au Nanocomposites Based Photocatalytic Oxidation of Benzoic Acid. Scijet 2015, 4, 120.
  • M. Ismael. A Review and Recent Advances in Solar-to-Hydrogen Energy Conversion Based on Photocatalytic Water Splitting Over Doped-TiO2 Nanoparticles. Sol. Energy 2020, 211, 522–546. doi:10.1016/j.solener.2020.09.073
  • Rathi, V.H.; Jeice, A.R.; Jayakumar, K. Green Synthesis of Ag/CuO and Ag/TiO2 Nanoparticles for Enhanced Photocatalytic dye Degradation, Antibacterial, and Antifungal Properties. Appl Surf Sci Adv 2023, 18, 100476. doi:10.1016/j.apsadv.2023.100476
  • Nabi, I.; Li, K.; Cheng, H.; Wang, T.; Liu, Y.; Ajmal, S.; Yang, Y.; Feng, Y.; Zhang, L. Complete Photocatalytic Mineralization of Microplastic on TiO2 Nanoparticle Film. Iscience 2020, 23, 1–12. doi:10.1016/j.isci.2020.101326
  • Nemiwal, M.; Kumar, D. TiO2 and SiO2 Encapsulated Metal Nanoparticles: Synthetic Strategies, Properties, and Photocatalytic Applications. Inorg. Chem. Commun. 2021, 128, 108602. doi:10.1016/j.inoche.2021.108602
  • Iravani, S. Green Synthesis of Metal Nanoparticles Using Plants. Green Chem. 2011, 13, 2638–2650. doi:10.1039/c1gc15386b
  • Rajaram, P.; Jeice, A.R.; Jayakumar, K. Review of Green Synthesized TiO2 Nanoparticles for Diverse Applications. Surfaces and Interfaces 2023, 39, 102912. doi:10.1016/j.surfin.2023.102912
  • Akhtar, M.S.; Panwar, J.; Yun, T.S. Biogenic Synthesis of Metallic Nanoparticles by Plant Extracts. ACS. Sustain. Chem. Eng. 2013, 1, 591–602. doi:10.1021/sc300118u
  • Ying, S.; Guan, Z.; Ofoegbu, P.C.; Clubb, P.; Rico, C.; He, F.; Hong, J. Green Synthesis of Nanoparticles: Current Developments and Limitations. Environ Tech Innov 2022, 26, 102336. doi:10.1016/j.eti.2022.102336
  • Gour, A.; Jain, N.K. Advances in Green Synthesis of Nanoparticles. Artif. Cells. Nanomed. Biotechnol. 2019, 47, 844–851. doi:10.1080/21691401.2019.1577878
  • Ying, S.; Guan, Z.; Ofoegbu, P.C.; Clubb, P.; Rico, C.; He, F.; Hong, J. Green Synthesis of Nanoparticles: Current Developments and Limitations. Environ Techn Innov 2022, 26, 102336. doi:10.1016/j.eti.2022.102336
  • Ganesan, S.; Babu, I.G.; Mahendran, D.; Arulselvi, P.I.; Elangovan, N.; Geetha, N.; Venkatachalam, P. Green Engineering of Titanium Dioxide Nanoparticles Using Ageratina Altissima (L.) King & H.E. Robines. Medicinal Plant Aqueous Leaf Extracts for Enhanced Photocatalytic Activity. Ann Phyto Inter J 2016, 5, 69–75. doi:10.21276/ap.2016.5.2.8
  • Abdul Jalill, R.D.; Nuaman, R.S.; Abd, A.N. Biological Synthesis of Titanium Dioxide Nanoparticles by Curcuma Longa Plant Extract and Study its Biological Properties, World Sci. News 2016, 49, 204–222.
  • Pavithra, S.; Bessy, T.C.; Bindhu, M.R.; Venkatesan, R.; Parimaladevi, R.; Alam, M.M.; Mayandi, J.; Umadevi, M. Photocatalytic and Photovoltaic Applications of Green Synthesized Titanium Oxide (TiO2) Nanoparticles by Calotropis Gigantea Extract. J. Alloys Compd. 2023, 960, 170638. doi:10.1016/j.jallcom.2023.170638
  • Ahmed, N.K.; Abbady, A.; Elhassan, Y.A.; Said, A.H. Green Synthesized Titanium Dioxide Nanoparticle from Aloe Vera Extract as a Promising Candidate for Radiosensitization Applications. Bionanoscience. 2023, 13, 730–743. doi:10.1007/s12668-023-01085-2
  • Rajaram, P.; Jeice, A.R.; Jayakumar, K., Influences of Calcination Temperature on Titanium Dioxide Nanoparticles Synthesized Using Averrhoa Carambola Leaf Extract: In Vitro Antimicrobial Activity and UV-Light Catalyzed Degradation of Textile Wastewater. Bio Conver Bio. 2023, 1–14. doi:10.1007/s13399-023-04212-x
  • Rathi, V.H.; Jeice, A.R. Green Fabrication of Titanium Dioxide Nanoparticles and Their Applications in Photocatalytic dye Degradation and Microbial Activities. Chem Phys Imp 2023, 6, 100197. doi:10.1016/j.chphi.2023.100197
  • Paranthaman, R.; Praveen, K.P.; Kumaravel, S. GC-MS Analysis of Phytochemicals and Simultaneous Determination of Flavonoids in Amaranthus Caudatus (Sirukeerai) by RP-HPLC. J. Anal. Bioanal. Tech. 2012, 03, 1–4.
  • Permana, M.D.; Noviyanti, A.R.; Lestari, P.R.; Kumada, N.; Eddy, D.R.; Rahayu, I. Enhancing the Photocatalytic Activity of TiO2/Na2Ti6O13 Composites by Gold for the Photodegradation of Phenol. Chemengineering 2022, 6, 69. doi:10.3390/chemengineering6050069
  • Eddy, D.R.; Nursyamsiah, D.; Permana, M.D.; Solihudin; Noviyanti, A.R.; Rahayu, I., Green Production of Zero-Valent Iron (ZVI) Using tea-Leaf Extracts for Fenton Degradation of Mixed Rhodamine B and Methyl Orange Dyes, Materials. (Basel) 2022, 15, 332. doi:10.3390/ma15010332
  • Attard, E. A Rapid Microtitre Plate Folin-Ciocalteu Method for the Assessment of Polyphenols. Open Life Sciences 2013, 8, 48–53. doi:10.2478/s11535-012-0107-3
  • Converso, A.; Hartingh, T.; Garbaccio, R.M.; Tasber, E.; Rickert, K.; Fraley, M.E.; Yan, Y.; Kreatsoulas, C.; Stirdivant, S.; Drakas, B.; Walsh, E.S. Development of Thioquinazolinones, Allosteric Chk1 Kinase Inhibitors. Bioorg. Med. Chem. Lett. 2009, 19, 1240–1244. doi:10.1016/j.bmcl.2008.12.076
  • Wang, H.M.; Qu, L.Q.; Ng, J.P.; Zeng, W.; Yu, L.; Song, L.L.; Wong, V.K.W.; Xia, C.L.; Law, B.Y.K. Natural Citrus Flavanone 5-Demethylnobiletin Stimulates Melanogenesis Through the Activation of cAMP/CREB Pathway in B16F10 Cells. Phytomedicine 2022, 98, 153941. doi:10.1016/j.phymed.2022.153941
  • Ksouri, R.; Megdiche, W.; Debez, A.; Falleh, H.; Grignon, C.; Abdelly, C. Salinity Effects on Polyphenol Content and Antioxidant Activities in Leaves of the Halophyte Cakile Maritima. Plant Physiol. Biochem. 2007, 45, 244–249. doi:10.1016/j.plaphy.2007.02.001
  • Zenkevich, I.G.; Makarov, A.A. Identification of Alkylarene Chloromethylation Products Using gas-Chromatographic Retention Indices. Russ. J. Gen. Chem. 2007, 77, 611–619. doi:10.1134/S1070363207040196
  • Babalola, I.T.; Shode, F.O. Ubiquitous Ursolic Acid: A Potential Pentacyclic Triterpene Natural Product. J. Pharmacogn. Phytochem 2013, 2, 214–222.
  • Sani, H.A.; Rahmat, A.; Ismail, M.; Rosli, R.; Endrini, S. Potential Anticancer Effect of red Spinach (Amaranthus Gangeticus) Extract. Asia Pac. J. Clin. Nutr 2004, 13, 396–400.
  • Fatimah, I.; Aftrid, Z.H.V.I. Characteristics and Antibacterial Activity of Green Synthesized Silver Nanoparticles Using red Spinach (Amaranthus Tricolor L.) Leaf Extract. Green Chemistry Letters and Reviews 2019, 12, 25–30. doi:10.1080/17518253.2019.1569729
  • Eddy, D.R.; Sheha, G.A.N.; Permana, M.D.; Saito, N.; Takei, T.; Kumada, N.; Rahayu, I.; Abe, I.; Sekine, Y.; Oyumi, T.; Izumi, Y. Study on Triphase of Polymorphs TiO2 (Anatase/Rutile/Brookite) for Boosting Photocatalytic Activity of Metformin Degradation. Chemosphere 2024, 351, 141206. doi:10.1016/j.chemosphere.2024.141206
  • González-Manteiga, W.; R.M. Crujeiras. An Updated Review of Goodness-of-Fit Tests for Regression Models. Test 2013, 22, 361–411. doi:10.1007/s11749-013-0327-5
  • Singh, S.; Maurya, I.C.; Srivastava, P.; Bahadur, L. Synthesis of Nanosized TiO2 Using Different Molecular Weight Polyethylene Glycol (PEG) as Capping Agent and Their Performance as Photoanode in dye-Sensitized Solar Cells. J. Solid State Electrochem. 2020, 24, 2395–2403. doi:10.1007/s10008-020-04768-y
  • Jha, A.K.; Prasad, K.; Kulkarni, A.R. Synthesis of TiO2 Nanoparticles Using Microorganisms. Colloids Surf., B 2009, 71, 226–229. doi:10.1016/j.colsurfb.2009.02.007
  • Catauro, M.; Tranquillo, E.; Dal Poggetto, G.; Pasquali, M.; Dell’Era, A.; Ciprioti, S.V. Influence of the Heat Treatment on the Particles Size and on the Crystalline Phase of TiO2 Synthesized by the sol-gel Method. Materials. (Basel) 2018, 11, 2364. doi:10.3390/ma11122364
  • Dong, H.; Zhao, F.; He, Q.; Xie, Y.; Zeng, Y.; Zhang, L.; Tang, L.; Zeng, G. Physicochemical Transformation of Carboxymethyl Cellulose-Coated Zero-Valent Iron Nanoparticles (nZVI) in Simulated Groundwater Under Anaerobic Conditions. Sep. Purif. Technol. 2017, 175, 376–383. doi:10.1016/j.seppur.2016.11.053
  • Ying, L.; Hon, L.S.; White, T.; Withers, R.; Hai, L.B. Controlled Nanophase Development in Photocatalytic Titania. Mater. Trans. 2003, 44, 1328–1332. doi:10.2320/matertrans.44.1328
  • Goyal, A.; Rumaiz, A.K.; Miao, Y.; Hazra, S.; Ni, C.; Shah, S.I. Synthesis and Characterization of TiO2–Ge Nanocomposites. J Vac Sci Tech B: Microelect Nanomet Struct Process, Measure Pheno 2008, 26, 1315–1320. doi:10.1116/1.2939262
  • Waghmode, M.S.; Gunjal, A.B.; Mulla, J.A.; Patil, N.N.; Nawani, N.N. Studies on the Titanium Dioxide Nanoparticles: Biosynthesis, Applications and Remediation. SN Applied Sciences 2019, 1, 310. doi:10.1007/s42452-019-0337-3
  • Arularasu, M.V. Effect of Organic Capping Agents on the Optical and Photocatalytic Activity of Mesoporous TiO2 Nanoparticles by sol–gel Method. SN Applied Sciences 2019, 1, 393. doi:10.1007/s42452-019-0424-5
  • Kumari, Y.; Jangir, L.K.; Kumar, A.; Kumar, M.; Awasthi, K. Investigation of Thermal Stability of TiO2 Nanoparticles Using 1-Thioglycerol as Capping Agent. Solid State Commun. 2017, 263, 1–5. doi:10.1016/j.ssc.2017.07.001
  • Khatoon, N.; Mazumder, J.A.; Sardar, M. Biotechnological Applications of Green Synthesized Silver Nanoparticles. J Nanosci: Curr Res 2017, 02, 2572–0813. doi:10.4172/2572-0813.1000107
  • Malik, A.Q.; Mir, T.U.G.; Kumar, D.; Mir, I.A.; Rashid, A.; Ayoub, M.; Shukla, S., A Review on the Green Synthesis of Nanoparticles, Their Biological Applications, and Photocatalytic Efficiency Against Environmental Toxins. Environ Sci Pollut Res. 2023, 30, 69796–69823. doi:10.1007/s11356-023-27437-9
  • Begum, R.; Najeeb, J.; Sattar, A.; Naseem, K.; Irfan, A.; Al-Sehemi, A.G.; Farooqi, Z.H. Chemical Reduction of Methylene Blue in the Presence of Nanocatalysts: A Critical Review. Rev. Chem. Eng. 2020, 36, 749–770. doi:10.1515/revce-2018-0047
  • Kaur, H.; Kaur, S.; Singh, J.; Rawat, M.; Kumar, S. Expanding Horizon: Green Synthesis of TiO2 Nanoparticles Using Carica Papaya Leaves for Photocatalysis Application. Mater. Res. Express 2019, 6, 095034. doi:10.1088/2053-1591/ab2ec5
  • Srujana, S.; Anjamma, M.; Alimuddin; Singh, B.; Dhakar, R.C.; Natarajan, S.; Hechhu, R. A Comprehensive Study on the Synthesis and Characterization of TiO2 Nanoparticles Using Aloe Vera Plant Extract and Their Photocatalytic Activity Against MB Dye. Adsorpt. Sci. Technol. 2022, 2022, 7244006.
  • Al-hamoud, K.; Shaik, M.R.; Khan, M.; Alkhathlan, H.Z.; Adil, S.F.; Kuniyil, M.; Khan, M. Pulicaria Undulata Extract-Mediated Eco-Friendly Preparation of TiO2 Nanoparticles for Photocatalytic Degradation of Methylene Blue and Methyl Orange. ACS Omega 2022, 7, 4812–4820. doi:10.1021/acsomega.1c05090
  • Shimi, A.K.; Ahmed, H.M.; Wahab, M.; Katheria, S.; Wabaidur, S.M.; Eldesoky, S.M.; Rane, K.P., Synthesis and Applications of Green Synthesized TiO2 Nanoparticles for Photocatalytic dye Degradation and Antibacterial Activity. J. Nanomater. 2022, 2022, 7060388.
  • Nabi, G.; Ain, Q.-U.; Tahir, M.B.; Nadeem Riaz, K.; Iqbal, T.; Rafique, M.; Rizwan, M. Green Synthesis of TiO2 Nanoparticles Using Lemon Peel Extract: Their Optical and Photocatalytic Properties. Int. J. Environ. Anal. Chem. 2022, 102, 434–442. doi:10.1080/03067319.2020.1722816
  • Rosman, N.; Salleh, W.N.W.; Mohamed, M.A.; Harun, Z.; Ismail, A.F.; Aziz, F. Constructing a Compact Heterojunction Structure of Ag2CO3/Ag2O in-Situ Intermediate Phase Transformation Decorated on ZnO with Superior Photocatalytic Degradation of Ibuprofen. Sep. Purif. Technol. 2020, 251, 117391. doi:10.1016/j.seppur.2020.117391
  • Muniandy, S.S.; Kaus, N.H.M.; Jiang, Z.T.; Altarawneh, M.; Lee, H.L. Green Synthesis of Mesoporous Anatase TiO2 Nanoparticles and Their Photocatalytic Activities. RSC Adv. 2017, 7, 48083–48094. doi:10.1039/C7RA08187A
  • Alvaro, M.; Aprile, C.; Benitez, M.; Carbonell, E.; García, H. Photocatalytic Activity of Structured Mesoporous TiO2 Materials. J. Phys. Chem. B 2006, 110, 6661–6665. doi:10.1021/jp0573240
  • Neethu, N.; Choudhury, T. Treatment of Methylene Blue and Methyl Orange Dyes in Wastewater by Grafted Titania Pillared Clay Membranes, Recent Pat. Nanotechnol 2018, 12, 200–207.
  • Azeez, F.; Al-Hetlani, E.; Arafa, M.; Abdelmonem, Y.; Nazeer, A.A.; Amin, M.O.; Madkour, M. The Effect of Surface Charge on Photocatalytic Degradation of Methylene Blue dye Using Chargeable Titania Nanoparticles. Sci. Rep. 2018, 8, 7104. doi:10.1038/s41598-018-25673-5