Green synthesis of titanium dioxide nanoparticles using green tea (Camellia sinensis) extract: Characteristics and applications
-
Thi Kim Ngan Tran
, Le Khanh Van Tran
Abstract
Titanium dioxide (TiO2) nanoparticles were successfully synthesized using titanium isopropoxide as a substrate and green tea extract as a reducing agent in the synthesis process. The structural characteristics of TiO2 were analyzed using advanced techniques such as scanning electron microscope, energy dispersive X-ray, X-ray diffraction analysis, and N2 adsorption/desorption isotherm. TiO2 nanoparticles with a crystal size of 21.04 nm were calculated using the Debye–Scherer equation, indicating a dominant anatase structure. The synthesized TiO2 nanomaterial exhibited a spherical shape and formed aggregates, with a surface area of 18.33 m2·g−1. Additionally, the antibacterial activity of TiO2 nanoparticles was evaluated using the disk diffusion method, and the minimum inhibitory concentration against gram-positive bacteria Staphylococcus aureus and gram-negative bacteria Escherichia coli was found to be 7.5 and 11.25 mg·mL−1, respectively. In addition, the photocatalytic degradation of TiO2 activates and generates free radicals such as ˙OH and
Graphical abstract

1 Introduction
Green tea, with the scientific name Camellia sinensis, is a plant commonly grown to harvest leaves and buds, which are used in green tea production. Green tea leaves contain several useful components such as polyphenols, flavonoids, catechins, and other antioxidants, making it an excellent raw material in biological and chemical applications. Green tea is reported to contain nearly 4,000 bioactive compounds, of which one-third is contributed by polyphenols [1]. Green tea extract, typically derived from different parts of the tea plant, such as leaves, buds, or seeds, is a popular ingredient in cosmetics and skincare products due to its potential benefits [2,3]. Green tea extract is rich in polyphenols, especially catechins, which are powerful antioxidants. These antioxidants protect the skin from oxidative stress caused by free radicals, which can contribute to premature aging and skin damage. Anti-inflammatory properties can help calm and soothe the skin, making it suitable for products designed to reduce redness, irritation, and inflammation [4]. Green tea extract has been found to have some sun protection properties and may be included in sunscreens and sun protection products to enhance protection against ultraviolet (UV) radiation. In addition to its skincare benefits, green tea extract can be used in cosmetics and perfumes to provide a pleasant fragrance [2,3,5].
Metal oxide nanoparticles are nanometer-sized particles with a chemical structure based on metal oxide. These particles have numerous important applications in fields such as biomedicine, nanotechnology, the environment, and industry due to their unique properties [6,7]. Metal oxide nanoparticles typically range in size from 1 to 100 nm. Their small size increases their surface area relative to volume, enhancing their effectiveness in various applications. There are several methods for synthesizing metal oxide nanoparticles, each with its own advantages and disadvantages. Specifically, sol–gel, precipitation, hydrothermal, and solvothermal methods, among others, are commonly used [8]. Depending on the specific requirements of each application and the type of metal oxide, selecting the appropriate synthesis method can optimize product performance and quality. Additionally, green synthesis of metal oxide nanoparticles is an environmentally friendly approach that utilizes biological agents such as plants, microorganisms, and enzymes instead of toxic chemicals. This method is gaining increasing attention due to its safety, cost-effectiveness, and ability to minimize negative environmental impacts [9]. There are numerous reports on the preparation of TiO2 nanoparticles from various plants, such as Mangifera indica, Citrus reticulata, Azadirachta indica, Murraya koenigii, and Curcuma longa, for biomedical, antibacterial, or cosmetic applications [10,11,12]. Plant extracts can serve as reducing agents to convert metal precursors into metal oxide nanoparticles. This process leverages biological compounds in plant extracts, such as polyphenols, flavonoids, alkaloids, and antioxidants, to reduce metal ions and form nanoparticles.
Besides, TiO2 nanoparticles have attracted widespread attention in research and applications thanks to their outstanding properties and significant advantages [13,14]. TiO2 nanoparticles are a versatile material with several superior properties such as low production cost, mechanical and chemical stability, high transparency, biological and chemical inertness, hydrophilicity, and high light conversion and corrosion resistance performance. These properties make nano-TiO2 the ideal choice for a wide range of applications from cosmetics, energy, and environment to medicine and construction materials. Additionally, TiO2 nanoparticles exhibit strong antibacterial activity against both Gram-positive and Gram-negative bacteria. This antibacterial effect is due to its ability to break down the lipid outer membrane and generate powerful free radicals. For example, studies have demonstrated the ability of TiO2 nanoparticles to kill Escherichia coli and Pseudomonas aeruginosa [15].
The introduction of TiO2 particles has significantly improved the performance of TiO2 in cosmetics. Ultra-fine TiO2 particles maintain the ability to provide broad-spectrum protection against both UVA and UVB rays. Studies have shown that even at the nanoscale, TiO2 particles do not significantly penetrate the skin barrier, a fact confirmed by the Scientific Committee on Consumer Safety (SCCS) in 2014 [16]. In a study by Xie et al., conducted on mice, it was shown that nano-TiO2 did not penetrate the stratum corneum in intact or mildly damaged skin with a 2% sodium lauryl sulfate solution, both in vitro and in vivo [17]. Therefore, nanoparticles are often added at specific concentrations to achieve the desired UV protection factor (SPF) and other cosmetic benefits. However, the synthesis of TiO2 through traditional chemical methods often requires harsh reaction conditions and the use of toxic chemicals, which have negative impacts on the environment and human health. To overcome this problem, “green” TiO2 synthesis methods are being developed using natural raw materials such as plant extracts. Among them, green tea extract (C. sinensis) stands out due to its rich content of polyphenol compounds, especially catechin and epigallocatechin gallate, which act as reducing agents and stabilizers in the synthesis of nanoparticles. Expanding the potential for application in smart cosmetics, it also contributes to environmental remediation through its ability to decompose organic pollutants under light.
2 Materials and methods
2.1 Chemicals
The chemical synthesis of materials in this study includes the following: Folin Ciocalteu’s phenol reagent (2,2-diphenyl-1-picrylhydrazyl [DPPH], C18H12N5O6) and titanium(iv) isopropoxide (C12H28O4Ti, 99%), which were purchased from Sigma-Aldrich Co. (St. Louis, MO, USA).
Sodium hydroxide (NaOH, 99%), Potassium acetate (CH3COOK, 99%), and ethanol (C2H5OH, 99.7%) were obtained from Xilong Chemical Co., Ltd. (Shantou, China).
Cosmetic chemicals, including non-ionic surfactants, emollients, moisturizers, and preservatives, were purchased from Nguyen Ba Trading Production Co., Ltd, Tan Binh District, HCMC, Vietnam.
2.2 Plant materials
Green tea is harvested in Lam Dong province, Vietnam, and then transported to the laboratory. The materials are pre-treated to remove dirt, damaged leaves, and cleaned with running water to remove any remaining chemicals from the surface. Tea leaves are completely dried to avoid the growth of bacteria and mold using specialized drying equipment at a temperature of 50°C for an appropriate time.
2.3 Extraction process of the green tea plant
After drying, tea leaves can be ground into powder or left whole and then stored in airtight containers to avoid exposure to air and moisture. Take 1 g of ground green tea leaf powder and add 100 mL of solvent (water, ethanol, or water:ethanol in a 1:1 ratio). This solution is then stirred at 300 rpm for 90 min at a temperature of 50°C. The solution is then filtered to recover the extract for use in the synthesis of TiO2 nanoparticles.
2.4 Green synthesis of TiO2 nanoparticles using green tea leaf extract
The synthesis of TiO2 nanoparticles using 20 mL of green tea leaf extract and 20 mL of titanium tetraisopropoxide (0.1 mol·L−1) is based on the research of Pushpamalini et al. [18]. The ingredients were stirred well using a magnetic stirrer maintained at 60°C for 5 h. The solution was then cooled to reach room temperature (30°C). The resulting suspension was separated by centrifugation at 6,000 rpm for 15 min. After the reaction process, continue washing with ethanol solvent to remove impurities. The product is then dried at 120°C overnight. Subsequently, the sample was calcined in a JeioTech furnace at 600°C for 3 h. The TiO2 nanoparticles were cooled to room temperature, and their structural characteristics were analyzed.
2.5 Characterization of TiO2 nanoparticles
Characterization of TiO2 nanoparticle structure was performed through various analytical methods to determine important properties such as particle size, surface morphology, crystal structure, and chemical properties. Details of the surface structure, size, and morphology of the nanoparticles were analyzed using a scanning electron microscope (SEM) S4800 (JEOL, Japan). The D8 Advanced equipment (Bruker, Germany) with Cu Kα radiation (λ = 1.5418 Å) and a 2θ range from 10° to 70° was used for X-ray diffraction (XRD) analysis to determine the crystal structure and lattice parameters. Fourier transform infrared spectroscopy was conducted on a Nicolet 6700 (Thermo Fisher Scientific, USA) to determine functional groups and chemical bonds in TiO2. The N2 adsorption–desorption isotherm method based on the Brunauer–Emmett–Teller (BET) theory was used to determine the specific surface area and pore size distribution of nanoparticles, performed on Micromeritics 2020 equipment (USA). A UV–Vis spectrophotometer (Metash UV-5100, China) was used to analyze the phytochemical composition of green tea extract.
2.6 Antibacterial activity
Antibacterial activity was evaluated using the agar disk diffusion method with two bacterial standards, Gram-negative (E. coli ATCC 8739) and Gram-positive (Staphylococcus aureus ATCC 6538). Bacteria were activated in Mueller Hinton medium for 24 h with a bacterial density of approximately 106–107 CFU·mL−1 [19]. Different concentrations of the sample solution were pipetted into 8 mm diameter wells on agar plates spread with test bacteria. Chloramphenicol (1 mg·mL−1) was used as a positive control and solvent as a negative control. After 24 h, results were recorded by image and diameter of the sterile zone. The minimum inhibitory concentration (MIC) value was determined as the lowest concentration of TiO2 particles that inhibited bacterial growth (no change in color of the resazurin reagent in the well) [20].
2.7 Application orientation for foaming face wash products
For the preparation of foaming facial cleanser products, the process typically involves carefully combining the oil and water phases to create a stable and effective product based on the formula provided in Table 1. The oil phase preparation includes weighing emollients, emulsifiers, thickeners, preservatives, and oil-soluble active ingredients, heating the mixture to 70–75°C until all ingredients are melted and homogenous. The water phase preparation involves mixing water-soluble ingredients such as humectants, surfactants, chelating agents, and preservatives, heating the mixture to the same temperature as the oil phase, and stirring continuously until all ingredients are dissolved.
Formula ingredients of foaming face wash products
Ingredients | Formulations (%) | |||||
---|---|---|---|---|---|---|
F0 | F1 | F2 | F3 | F4 | F5 | |
Sodium methyl cocoyl taurate | 7.5 | 7.5 | 7.5 | 7.5 | 7.5 | 7.5 |
Diisostearyl malate | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 |
Cocamidopropyl betaine | 5 | 5 | 5 | 5 | 5 | 5 |
Glycerin | 1 | 1 | 1 | 1 | 1 | 1 |
d-panthanol | 0.1 | 0.1 | 0.1 | 0.1 | 0.1 | 0.1 |
Cetyl alcohol | 4 | 4 | 4 | 4 | 4 | 4 |
PEG-40 | 2 | 2 | 2 | 2 | 2 | 2 |
TiO2 | 0 | 0.1 | 0.2 | 0.3 | 0.4 | 0.5 |
Citric acid | 0.25 | 0.25 | 0.25 | 0.25 | 0.25 | 0.25 |
Vitamin E | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 |
Aloe vera extract | 0.3 | 0.3 | 0.3 | 0.3 | 0.3 | 0.3 |
Tea tree essential oil | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 | 0.5 |
Deionized water | qs 100 | qs 100 | qs 100 | qs 100 | qs 100 | qs 100 |
2.8 Physicochemical properties of raw materials and cosmetic products
Parameters for evaluating the physical and chemical properties and biological activity of raw materials and cosmetic products are shown in Table 2.
Physicochemical analysis methods
Targets | Analytical method |
---|---|
pH | Hanna HI-2211 pH meter (Hanna Romania) |
Viscosity | Viscosity Brookfield DVEELVTJ0 Viscometer (USA) |
Foamability | Shaking test |
DPPH antioxidant activity (mg/100 g) | DPPH free radical scavenging ability [21] |
Total flavonoid content (TFC) (mgQE·g−1 DW) | Aluminum chloride colorimetric method [22] |
Total polyphenols content (TPC) (mgAAE/100 g DW) | Folin–Ciocalteu’s method [22] |
2.9 Analyzing SPF value
The SPF value is analyzed based on the level of skin protection from UVB rays in the absorption range from 290 to 320 nm (UVB range). The sample was prepared at a concentration of 0.2 g and titrated to 100 mL with ethanol solvent. Based on the publication by Mansur et al. [23], the absorbance results were used to determine the SPF value according to the formula:
where EE (λ) × 1 is the erythemal effect of radiation with wavelength λ, Abs (λ) is the value of spectral absorbance at wavelength λ, and CF is the correction factor [10]. The values of EE × λ are constant.
2.10 Photocatalytic activity
Tetracycline is an antibiotic used to evaluate the photocatalytic ability of TiO2 in aqueous medium. The light source used is a 40 W LED lamp. At room temperature, the reaction system is carried out with a solution mixture including a catalyst sample (5 mg) and tetracycline (20 ppm) stirred in a 250 mL beaker. About 1 mL of the solution mixture is taken out at a time point, including stirring in dark and light conditions. The sample after being taken out is centrifuged at 6,000 rpm for 10 min to completely remove solids. Then, the concentration of the solution is determined using a UV–Vis spectrophotometer, at the maximum absorption wavelength λ = 357 nm.
3 Results and discussion
3.1 Physicochemical of green tea extract
To date, conventional organic solvents such as hexane, acetone, and chloroform remain the most commonly used solvents for extracting bioactive compounds from plant materials. However, some organic solvents have distinct disadvantages, such as high cost, residue levels, and toxicity [24]. Recently, the concepts of ecological sustainability and a green economy have sparked interest in solvents like water or ethanol as potential substitutes for organic solvents.
Green tea extraction can be performed using various solvents to obtain desired chemical compounds. Common and effective solvents for extracting polyphenols and flavonoids and determining antioxidant activity are listed in Table 3. The extraction efficiency is presented in Table 4, showing the highest TFC, total polyphenols content (TPC), and DPPH content when extracted with ethanol solvent as 12.425 ± 0.941, 351.45 ± 7.546, and 109.757 ± 0.184, respectively, compared to water solvent and a water–alcohol solvent mixture.
Quantification of compounds in green tea extract
Parameters | Results | Phenomena |
---|---|---|
Alkaloids | + | Reddish brown precipitate |
Saponin | + | Durable foam |
Flavonoids | + | Yellow precipitate |
Terpenoid/steroid | − | Brick red color |
Tannins | + | Dark green precipitate |
Phytochemistry of green tea extract
Water | Ethanol | Water:ethanol | |
---|---|---|---|
TFC (mgQE/g1 DW) | 7.535 ± 0.543 | 12.425 ± 0.941 | 11.889 ± 0.073 |
DPPH radical scavenging activity (mg 100/g DW) | 104.88 ± 0.591 | 109.757 ± 0.184 | 96.809 ± 0.771 |
TPC (mgAAE 100/g DW) | 317.403 ± 1.007 | 351.45 ± 7.546 | 324.357 ± 5.113 |
Green tea is rich in polyphenols that can reduce metal ions to form metal nanoparticles. Polyphenols can bind to the surfaces of nanoparticles, stabilize them, and prevent agglomeration. Phenolic groups in polyphenols form strong interactions with metal atoms on the nanoparticle surfaces, helping maintain their size and stability, making them durable, and preventing agglomeration of the resulting nanoparticles. Polyphenols in green tea extract can interact with the surface of TiO2 nanoparticles, preventing agglomeration through electrostatic and/or steric stabilization mechanisms. This is due to the chelating (binding) ability of polyphenols with metal ions, as well as hydrogen interactions between the hydroxyl group of polyphenols and the TiO2 surface [25,26].
3.2 Characterization of TiO2 material from green tea extract
The XRD pattern of the synthesized TiO2 clearly shows the crystal structure of the nanoparticles (Figure 1). Sharp diffraction peaks are observed at 2θ values of 25.47°, 27.60°, 36.24°, 37.95°, 38.72°, 48.23°, 54.08°, 55.25°, 62.88°, and 69.05°. The diffraction peaks were compared with standard reference samples such as rutile Ti (JCPDS 88-1175) and anatase Ti (JCPDS 78-2486). The anatase phase shows the most intense diffraction peak corresponding to the Miller index (hkl) value of (101), along with other peaks responsible for anatase such as (004), (200), (105), (211), (204), and (116). The rutile phase is represented by a peak at approximately 27.60° (110), along with peaks of small intensity such as (101), 41.20° (111), 54.30° (211), and 56.60° (220). The Scherrer equation was used to estimate the average nanoparticle size of 21.04 nm (Table 5):
where K is a constant (usually ∼0.9), λ is the X-ray wavelength, β is the full width at half maximum (FWHM) of the peak, and θ is the Bragg angle.

XRD pattern of ZnO.
XRD analysis data and TiO2 nanoparticle size
Peak position (2θ) | FWHM (β) | Miller indices | Particle size (nm) |
---|---|---|---|
25.46525 | 0.35351 | (101) | 23.03017885 |
27.59778 | 0.29724 | (110) | 27.51042242 |
36.24086 | 0.31934 | (103) | 26.16518912 |
37.95483 | 0.38802 | (004) | 21.64226835 |
38.72484 | 0.60023 | (112) | 14.02341327 |
48.22608 | 0.4492 | (200) | 19.36860982 |
54.07888 | 0.49996 | (105) | 17.8329081 |
55.2549 | 0.48512 | (211) | 18.47617307 |
62.87807 | 0.51377 | (204) | 18.11625785 |
69.04867 | 0.3975 | (116) | 24.24833065 |
Similarly, nanoparticles with different crystal sizes depending on the reducing agent from plant extracts such as neem extract (A. indica) are about 10–50 nm [27], orange peel extract (C. reticulata) is 24 nm [28], and jasmine extract has an average crystal size of 31–42 nm [29]. Calcination at 600°C can cause the conversion of anatase to rutile; high calcination temperatures often increase the crystallinity of TiO2, resulting in sharper and more well-defined peaks in the XRD patterns. At 600°C, TiO2 crystals grow larger, reducing the width of the diffraction peaks.
For TiO2 synthesized using green tea extract, FT-IR can help determine the presence of organic compounds from green tea extract and the formation of TiO2 before and after heating at 600°C (Figure 2). In the TiO2 sample before calcination, the spectrum was dominated by peaks related to organic functional groups from green tea extract and hydroxyl groups [30]. O–H stretching vibrations at broad bands around 3,200–3,600 cm−1 indicate the presence of hydroxyl groups (from green tea extract and adsorbed water). C–H bonds in organic molecules from green tea extract are shown through peaks at 2,800–3,000 cm−1. Peaks around 1,600–1,650 cm−1 (aromatic ring C═C) and 1,000–1,300 cm−1 (C–O stretching) reflect the organic components in green tea extract [31]. The removal of organic compounds is evident from the disappearance of the peaks associated with C–H, C═C, and C–O vibrations after heating to 600°C [32]. The peaks of Ti–O–Ti (493 cm−1) and Ti–O (614 cm−1) become more obvious, and the prolonged absorption band of Ti–O at peak 763 cm−1 shows the anatase phase of the structure of TiO2. The TiO2 structure was determined through green synthesis [33]. The presence of the Ti–O–Ti bond proves the interaction of biological molecules with the TTIP precursor by the phytochemical constituents polyphenol, flavonoids, alkaloids, tannins, and terpenoids.

FT-IR diagram of TiO2.
SEM analysis provides detailed images of the surface morphology and size of TiO2 nanoparticles. When TiO2 nanoparticles are heated at 600°C, their morphology and structure can change significantly. TiO2 nanoparticles synthesized through green methods typically appear as small, uniformly distributed particles. There may be some degree of agglomeration, where the nanoparticles clump together. The surface may appear relatively smooth with some organic residue from the green synthetic agents. After calcination at 600°C, there is a decrease in the degree of agglomeration as some smaller particles combine into larger particles, indicating increased crystallinity. There is less agglomeration compared to the unburnt sample. These changes are indicative of thermal processes occurring during firing, such as sintering and phase transformation. When analyzing TiO2 nanoparticles, energy dispersive X-ray (EDX) can provide valuable information on the presence and proportions of Ti and O elements and can detect any impurities or residues from the synthesis such as carbon in Figure 3. The sample before calcination in Figure 3 shows characteristic peaks of Ti commonly observed at about 4.51 and 4.93 keV. The O peak typically appears at about 0.52 keV, and there are small peaks for carbon (C at about 0.28 keV), indicating the presence of organic residues from the green synthesis. After calcination, the EDX spectrum still shows prominent peaks for Ti (64.23%) and O (35.77%), but there is no presence of the C peak, which indicates that these residues have been removed during the firing process.

SEM (a) and (b) and EDX (d) and (c) images of TiO2 nanoparticles. “a” and “c” at initial condition and “b” and “d” at 600°C.
The BET surface area is used to determine the surface area, pore size distribution, and porosity of TiO2 materials. When TiO2 is synthesized via green methods, Figure 4 isotherms can provide detailed information on how the synthesis process affects the nanoparticle properties. According to the IUPAC classification of physical absorption isotherms, TiO2 nanoparticles correspond to type IV isotherms. This type of isotherm often shows a hysteresis loop, characteristic of capillary condensation occurring in mesopores. The significant reduction in the BET surface area from 691.24 to 18.33 m²·g−1 upon calcination at 600°C can be attributed to the removal of organic samples used in the synthesis [34]. Green tea extract acts as a template, creating a highly porous structure with small mesopores or micropores, resulting in a high BET surface area (Table 6). Firing at 600°C decomposes and removes the green tea extract. Heat treatment causes sintering, leading to the collapse of the pores and the formation of larger pores. The change from many small pores to fewer large pores significantly reduces the overall surface area. The pore size confirms significant structural changes in the TiO2 nanoparticles before and after calcination, increasing from 4.1 to 19.5 nm. These changes reflect the typical effects of calcination of a TiO2 material sample, where the organic components are removed, and the inorganic framework undergoes sintering.

N2 adsorption–desorption isotherm (a) and (c) curve and pore size distribution (b) and (d) of TiO2: “a” and “c” at initial condition and “b” and “d” at 600°C.
BET analysis data of TiO2 nanoparticles
Samples | BET surface area (m2·g−1) | Pore volume (cm3·g−1) | Pore size (nm) |
---|---|---|---|
TiO2 initial | 691.24 | 0.72 | 4.1 |
TiO2 600°C | 18.33 | 0.12 | 19.5 |
3.3 Evaluation of the antibacterial ability of TiO2 nanoparticles
TiO2 nanoparticles were used to determine the antibacterial activity against two bacterial strains, E. coli and S. aureus. The results of determining the antibacterial activity of TiO2 are presented in Figure 5. TiO2 nanoparticles did not show significant antibacterial activity against E. coli and S. aureus at the concentrations tested by the disk diffusion method. No inhibition zones were observed around the wells impregnated with TiO2 nanoparticles. The absence of an inhibition zone for the two bacterial strains indicates that the concentration of TiO2 nanoparticles used in this study was not effective against E. coli and S. aureus in the disk diffusion assay. Another study by Nguyen et al. tested antibacterial activity and showed that pure TiO2 nanoparticles did not show any inhibitory effect on E. coli and S. aureus bacteria [35].

Antibacterial activity of TiO2 nanoparticles by the agar diffusion method: (a) Escherichia coli and (b) Staphylococcus aureus.
Therefore, determining the MIC is an important assessment to determine the lowest concentration of TiO2 nanoparticles required to inhibit the visible growth of E. coli and S. aureus after 24 h of incubation with TiO2 (Table 7 and Figure 6). The MIC values of TiO2 against E. coli were 7.5 mg·mL−1. This indicated that a concentration of 7.5 mg·mL−1 TiO2 was sufficient to inhibit the visible growth of E. coli after 24 h of incubation. Similarly, a concentration of 11.25 mg·mL−1 could inhibit the growth of S. aureus. The MIC values indicate that E. coli is more susceptible to inhibition by TiO2 than S. aureus, as it requires a lower concentration of TiO2 to inhibit its growth. Based on the results of Table 7, it shows that S. aureus bacteria are less sensitive than E. coli bacteria. The higher MIC value for S. aureus indicates that this bacterium is more resistant to the effects of TiO2, possibly due to a thicker peptidoglycan layer, which may provide additional protection against invasion by TiO2 nanoparticles [36].
MIC of TiO2 nanoparticles
MIC (mg·mL−1) | ||
---|---|---|
Samples | E. coli ATCC 8739 | S. aureus ATCC 6538 |
TiO2 | 7.5 ± 0.14 | 11.25 ± 0.14 |
Chloramphenicol (μg·mL−1) | 0.2 | 0.2 |

Broth dilution method of the 96-well microplate on tested bacterial strains.
Azizi-Lalabadi et al. have shown that the thick peptidoglycan layer in the cell wall of the Gram-positive bacteria S. aureus will create more resistance than other bacteria. Gram-negative bacteria only have a thin peptidoglycan layer, which easily interacts with the charge of TiO2 nanoparticles [37]. The publication also shows that the negative charge on the cell surface of Gram-negative bacteria is higher than the negative charge on the cell surface of Gram-positive bacteria [38].
3.4 Photocatalytic activity of TiO2 nanoparticles
The efficiency of tetracycline removal in the photochemical reaction by TiO2 catalyst is shown through the light absorption intensity in the UV–Vis spectrum in Figure 7. As the irradiation time increases, the intensity of the characteristic peaks of tetracycline at 357 nm wavelength decreases abruptly after 240 min of irradiation, with a removal efficiency of about 73%. However, there is a shift in the maximum wavelength to the left at about 180 and 240 min at the end of the photochemical stage. The longer the reaction time, the higher the decomposition rate of tetracycline. This is because there is more time for photochemical reactions to take place, which helps to break down the structure of tetracycline into smaller decomposition products. When the reaction time is too long, the decomposition efficiency can reach a saturation level, meaning that there is no significant increase in the decomposition rate. This may be because all tetracycline molecules have been degraded or the degradation products have become saturated.

Effect of catalytic time (a) and adsorption spectrum of tetracycline (b).
The point of zero charge (pHpzc) is an important parameter in understanding the surface properties and photocatalytic activity of TiO2. For TiO2 with a pHpzc of 5.8, the surface charge characteristics at different pH values will affect its interaction with tetracycline antibiotics and its overall photocatalytic performance. Tetracycline is an amphoteric molecule, meaning it can act as both an acid and a base. Its ionization state depends on the pH of the solution. In general, tetracycline has different protonation states under acidic, neutral, and alkaline conditions, which affect its interaction with the TiO2 surface. Based on Figure 8b, the TiO2 surface is positively charged when pH < 5.8; TiO2 can adsorb negative charges (anions) due to electrostatic attraction. In contrast, solutions with pH > 5.8 have a negatively charged TiO2 surface, which is favorable for the adsorption of cationic species. The effective pH for the photodegradation of tetracycline by TiO2 (pHpzc 5.8) is usually slightly higher than pHpzc, ranging from 6 to 8 (Figure 8a). This allows for the efficient adsorption of positively charged tetracycline molecules and the generation of reactive oxidant species.

Effect of solution pH (a) and pHpzc value (b) of TiO2 nanoparticles.
In this study, increasing the dosage from 3 to 5 mg resulted in an increase in the tetracycline degradation efficiency from 48% to 80% after 240 min of illumination, respectively. The increase in decolorization efficiency with increasing amount of catalyst in the reaction is due to the increase in available surface area or active sites of the catalyst; increasing the amount of catalyst in the reaction will generate more ˙OH radicals. However, when the amount of photocatalyst was increased further (7 and 9 mg), the degradation efficiency tended to decrease due to the increased opacity of the suspension, leading to increased adsorption of tetracycline onto the material surface. Therefore, the light transmission efficiency to the material will be reduced, and the surface part of the photocatalyst will no longer be able to absorb light.
The tetracycline concentration significantly affects the photocatalytic degradation process using TiO2 (Figure 9b). At a tetracycline concentration of 20 mg·L−1 after 240 min, the degradation efficiency was approximately 76%. At higher concentrations, the degradation efficiency decreased. This was due to increased competition for available reactive oxygen species and active sites on the TiO2 surface. The reaction rate slowed as the TiO2 surface became saturated with tetracycline molecules. In addition, intermediate products formed during the degradation process may adsorb onto the TiO2 surface, further inhibiting the reaction.

Effect of catalyst dosage (a) and tetracycline antibiotic concentration (b).
The stability of TiO2 after multiple uses is a crucial aspect to consider when assessing the viability of the catalyst for environmental remediation purposes, particularly in the breakdown of antibiotics like tetracycline. The decrease in photocatalytic efficiency of TiO2 from 78.51% to 53.22% after four reuses indicates that various factors are impacting the reuse process and leading to a decline in catalyst performance (Figure 10). This decline is attributed to the build-up of degradation byproducts and impurities on the material’s surface, hindering light penetration to the catalyst. To address this issue, ethanol was employed to cleanse TiO2 after each cycle and remove impurities. However, over the reuse cycles, the nano-TiO2 particles tended to aggregate, diminishing the surface area and catalytic effectiveness. Additionally, some of the TiO2 particles may transition from the anatase phase (high activity) to the rutile phase (lower activity) after four reuse cycles. Similarly, some publications on the photocatalytic activity of green synthesized TiO2 particles have been made and presented in Table 8.

Recycling efficiency for green synthesis TiO2 catalyst.
Some publications on photocatalytic activity of green-synthesized TiO2 particles
Material precursor | Reducing agent (plant extraction) | % Removal | References |
---|---|---|---|
Titanium chloride | Jatropha curcas L. | 82.26% COD and 76.48% Cr (tannery wastewater) | [39] |
Titania bulk powder | Lemon peel | 70% Rhodamine B (RhB) | [40] |
Titanium(iv) tetraisopropoxide | Soluble starch | 13.3% Methylene blue (MB) dye | [41] |
Titanium isopropoxide | Phyllanthus emblica | 93% Coralline red dye | [42] |
Titanium isopropoxide | Syzygium cumini | 82.53% Lead (Pb2+) | [43] |
Titanium isopropoxide | Tinospora cordifolia | 94.43% Acid blue 113 | [44] |
Titanium(iii) chloride | Fresh aloe | 90% Ciprofloxacin (CIP) | [45] |
The photocatalytic mechanism of TiO2 (titanium dioxide) with tetracycline antibiotic occurring on the surface of TiO2 nanoparticles TiO2 nanoparticles can absorb UV light with a wavelength of 357 nm (Figure 11). When exposed to UV light, electrons in the valence band of TiO2 are excited to the conduction band, generating pairs of holes (h+) and electrons (e−). These holes and electrons can react with water and oxygen molecules on the TiO2 surface to produce reactive species like hydroxyl (˙OH) and superoxide (

Photocatalysis mechanism for TiO2 nanoparticles.
The efficiency of electron transfer can be enhanced by ensuring that the electron donor and acceptor molecules have appropriate chemical structures. Electron-rich organic compounds, like polyphenols or flavonoids found in natural extracts, can serve as electron donors, reducing the activation energy and enhancing the separation of electron–hole pairs in TiO2. Acceptor molecules with strong electron capture abilities, such as O2 or potent oxidants, can aid in reducing electron–hole recombination.
3.5 Application orientation for foaming face wash products
Specifications of facial cleanser products containing TiO2 may include factors such as pH, viscosity, and oxidation activity (Table 9). The pH of facial cleansers is usually in the range of 5–6.5. Research shows that products with a pH in the range of 5–6 are suitable for the natural pH of the skin, helping to maintain the acid mantle that protects the skin and prevents acne by inhibiting the growth of harmful bacteria. The viscosity of the product is in the range of 500–560 cP. The viscosity is adjusted relatively low to ensure the product has a moderate consistency for easy use in foam spray form and is easy to clean. Titanium dioxide is commonly used in skincare products, including facial cleansers, because of its various beneficial properties. TiO2 is an effective physical blocker of UV rays, both UVA and UVB. When incorporated into facial cleansers, it provides a layer of protection against the harmful effects of UV radiation. This helps enhance the product’s SPF value, helping to protect the skin from the sun more effectively. TiO2 is highly stable and does not easily decompose when exposed to light, heat, or air, ensuring the longevity and effectiveness of the facial cleanser. Table 10 shows the product’s effectiveness in protecting against UVB rays and the UVB-blocking ability of TiO2. Formula F0, without TiO2 nanoparticles, has the lowest SPF index of 8.02, while formulas F1–F5 have SPF values that change according to the content of TiO2 nanoparticles, but the values do not show significant deviation. Additionally, the presence of plant extracts and active skincare ingredients contributes to protecting the skin from the harmful effects of free radicals caused by sunlight, which can help enhance the product’s SPF index due to its antioxidant and skin-protecting properties. Studies have consistently shown that TiO2 nanoparticles do not penetrate beyond the outermost layer of the skin (stratum corneum) to reach skin cells. Based on a detailed analysis of studies, the SCCS has concluded that the use of nano-TiO2 in cosmetic products, including facial cleansers, is safe for use on healthy, intact skin [16].
Basic parameters of surveyed face wash products
Targets | F0 | F1 | F2 | F3 | F4 | F5 |
---|---|---|---|---|---|---|
pH | 5.8 | 6.1 | 6 | 5.4 | 5.6 | 6.0 |
Viscosity | 563.5 | 551.1 | 533.4 | 535.6 | 546.7 | 525.2 |
Foamability | 0.407 | 0.416 | 0.458 | 0.432 | 0.438 | 0.451 |
DPPH radical scavenging activity (mg/100 g) | 6.55 ± 0.071 | 6.763 ± 0.143 | 6.678 ± 0.078 | 6.548 ± 0.316 | 6.711 ± 0.192 | 6.956 ± 0.137 |
SPF value and sun protection ability of the product
Formulation | SPF value | Sunlight resistance (%) |
---|---|---|
F0 | 8.02 | 87.53 |
F1 | 9.80 | 89.79 |
F2 | 9.59 | 89.58 |
F3 | 10.28 | 90.27 |
F4 | 11.11 | 91.00 |
F5 | 11.31 | 91.16 |
4 Conclusions
Successfully synthesized TiO2 nanoparticles in the presence of a plant-reducing agent, green tea extract, and TTIP precursor. The nanoparticles exhibit a spherical structure and tend to aggregate. The TiO2 nanoparticles are predominantly in the anatase phase crystallized after heating at 600°C, with a particle size of 21.04 nm and a surface area of 18.33 m2·g−1. The antibacterial activity of the TiO2 nanoparticles was evaluated against Gram-positive (S. aureus) and Gram-negative (E. coli) bacteria. The nanoparticles demonstrated the ability to inhibit the visible growth of both E. coli and S. aureus bacteria after 24 h of incubation at their respective MIC. After 240 min of light irradiation, TiO2 nanoparticles were able to decompose 76% of tetracycline antibiotics at pH 6 with a concentration of 20 mg·L−1. Additionally, the cleanser product maintained its phase stability when stored at room temperature. The formulas exhibited SPF values that were not significantly different, ranging from 8 to 11, as calculated using the Mansur equation. In conclusion, green-synthesized TiO2 nanoparticles show promise for various applications due to their unique properties and environmentally friendly production methods, warranting further research across different industries.
-
Funding information: This study was supported by grants from Nguyen Tat Thanh University, Ho Chi Minh City, Viet Nam.
-
Author contributions: Ngoc Cat Thuyen Vo and Thi Nhu Dung Nguyen: conceptualization; Thi Kim Ngan Tran and Thi Que Minh Doan: data curation; Hoang Danh Pham and Thi Nhu Dung Nguyen: formal analysis; Le Khanh Van Tran and Ngoc Cat Nguyen Vo: methodology; Thi Kim Ngan Tran and Thi Que Minh Doan: writing – original draft; writing – review & editing.
-
Conflict of interest: The authors state no conflict of interest.
-
Data availability statement: All data generated or analyzed during this study are included in this published article.
References
[1] Mahmood T, Akhtar N, Khan BA. The morphology, characteristics, and medicinal properties of Camellia sinensis tea. J Med Plants Res. 2010;4(19):2028–33.10.5897/JMPR10.010Suche in Google Scholar
[2] Jhansi D, Kola M. The antioxidant potential of Centella asiatica: A review. J Med Plants Stud. 2019;18(2):18–20. 10.1016/j.etap.Suche in Google Scholar
[3] Namita P, Mukesh R, Vijay KJ. Camellia sinensis (green tea): A review. Glob J Pharmacol. 2012;6(2):52–9.Suche in Google Scholar
[4] Rahardiyan D. Antibacterial potential of catechin of tea (Camellia sinensis) and its applications. Food Res. 2019;3(1):1–6.10.26656/fr.2017.3(1).097Suche in Google Scholar
[5] Koch W, Zagórska J, Marzec Z, Kukula-Koch W. Applications of tea (Camellia sinensis) and its active constituents in cosmetics. Molecules. 2019;24(23):1–28.10.3390/molecules24234277Suche in Google Scholar PubMed PubMed Central
[6] Stark WJ, Stoessel PR, Wohlleben W, Hafner A. Industrial applications of nanoparticles. Chem Soc Rev. 2015;44(16):5793–805.10.1039/C4CS00362DSuche in Google Scholar
[7] Arruda SCC, Silva ALD, Galazzi RM, Azevedo RA, Arruda MAZ. Nanoparticles applied to plant science: a review. Talanta. 2015;131:693–705.10.1016/j.talanta.2014.08.050Suche in Google Scholar PubMed
[8] Pareek V, Bhargava A, Gupta R, Jain N, Panwar J. Synthesis and applications of noble metal nanoparticles: a review. Adv Sci Eng Med. 2017;9(7):527–44.10.1166/asem.2017.2027Suche in Google Scholar
[9] Rafique M, Sadaf I, Rafique MS, Tahir MB. A review on green synthesis of silver nanoparticles and their applications. Artif Cells, Nanomed Biotechnol. 2017;45(7):1272–91. 10.1080/21691401.2016.1241792.Suche in Google Scholar PubMed
[10] Hussein A, Al-Abodi EE. A review article: Green synthesis by using different plants to preparation oxide nanoparticles. Ibn AL-Haitham J Pure Appl Sci. 2023;36(1):246–59.10.30526/36.1.2933Suche in Google Scholar
[11] Chatterjee A, Kwatra N, Abraham J. Nanoparticles fabrication by plant extracts. Phytonanotechnology. Elsevier; 2020. p. 143–57. 10.1016/B978-0-12-822348-2.00008-5.Suche in Google Scholar
[12] Irshad MA, Nawaz R, ur Rehman MZ, Adrees M, Rizwan M, Ali S, et al. Synthesis, characterization and advanced sustainable applications of titanium dioxide nanoparticles: A review. Ecotoxicol Env Saf. 2021;212:111978. 10.1016/j.ecoenv.2021.111978.Suche in Google Scholar PubMed
[13] Rajaram P, Jeice AR, Jayakumar K. Review of green synthesized TiO2 nanoparticles for diverse applications. Surf Interfaces. 2023 Jul;39:102912, https://linkinghub.elsevier.com/retrieve/pii/S2468023023002821.10.1016/j.surfin.2023.102912Suche in Google Scholar
[14] Rajaram P, Jeice AR, 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. Biomass Convers Biorefinery. 2024 Sep;14(17):20665–78, https://link.springer.com/10.1007/s13399-023-04212-x.10.1007/s13399-023-04212-xSuche in Google Scholar
[15] Selvi JM, Murugalakshmi M, Sami P, Gnanaprakash M, Thanalakshmi R. Green synthesis, characterization and applications of TiO2 nanoparticles using aqueous extract of Erythrina variegata leaves. Curr Sci. 2022;123(1):59–66.10.18520/cs/v123/i1/59-66Suche in Google Scholar
[16] Dréno B, Alexis A, Chuberre B, Marinovich M. Safety of titanium dioxide nanoparticles in cosmetics. J Eur Acad Dermatol Venereol. 2019;33(S7):34–46.10.1111/jdv.15943Suche in Google Scholar PubMed
[17] Xie G, Lu W, Lu D. Penetration of titanium dioxide nanoparticles through slightly damaged skin in vitro and in vivo. J Appl Biomater Funct Mater. 2015;13(4):356–61.10.5301/jabfm.5000243Suche in Google Scholar PubMed
[18] Pushpamalini T, Keerthana M, Sangavi R, Nagaraj A, Kamaraj P. Comparative analysis of green synthesis of TiO2 nanoparticles using four different leaf extract. Mater Today Proc. 2020;40:S180–4. 10.1016/j.matpr.2020.08.438.Suche in Google Scholar
[19] Palaksha MN, Ahmed M, Das S. Antibacterial activity of garlic extract on streptomycin-resistant Staphylococcus aureus and Escherichia coli solely and in synergism with streptomycin. J Nat Sci Biol Med. 2010;1(1):12.10.4103/0976-9668.71666Suche in Google Scholar PubMed PubMed Central
[20] Cockerill FR. Performance standards for antimicrobial susceptibility testing. Approv Stand. 2010;30(1):M100–S20.Suche in Google Scholar
[21] Brand-Williams W, Cuvelier M-E, Berset C. Use of a free radical method to evaluate antioxidant activity. LWT-Food Sci Technol. 1995;28(1):25–30.10.1016/S0023-6438(95)80008-5Suche in Google Scholar
[22] Pham HNT, Tang Nguyen V, Van Vuong Q, Bowyer MC, Scarlett CJ. Bioactive compound yield and antioxidant capacity of Helicteres hirsuta Lour. stem as affected by various solvents and drying methods. J Food Process Preserv. 2017;41(1):9.10.1111/jfpp.12879Suche in Google Scholar
[23] Mansur MC, Leitão SG, Cerqueira-Coutinho C, Vermelho AB, Silva RS, Presgrave OA, et al. In vitro and in vivo evaluation of efficacy and safety of photoprotective formulations containing antioxidant extracts. Rev Bras Farmacogn. 2016;26(2):251–8. 10.1016/j.bjp.2015.11.006.Suche in Google Scholar
[24] Koch W, Kukuła-Koch W, Czop M, Helon P, Gumbarewicz E. The role of extracting solvents in the recovery of polyphenols from green tea and its antiradical activity supported by principal component analysis. Molecules. 2020;25(9):1–14.10.3390/molecules25092173Suche in Google Scholar PubMed PubMed Central
[25] Deng Z, Yi Z, Chen G, Ma X, Tang Y, Li X. Green tea polyphenol nanoparticle as a novel adsorbent to remove Pb2+ from wastewater. Mater Lett. 2021;284:128986, https://www.sciencedirect.com/science/article/pii/S0167577X20316931.10.1016/j.matlet.2020.128986Suche in Google Scholar
[26] Qiu H, Ni W, Yang L, Zhang Q. Remarkable ability of Pb(II) capture from water by self-assembled metal-phenolic networks prepared with tannic acid and ferric ions. Chem Eng J. 2022;450:138161, https://www.sciencedirect.com/science/article/pii/S1385894722036452.10.1016/j.cej.2022.138161Suche in Google Scholar
[27] Shekhar S, Singh S, Gandhi N, Gautam S, Sharma B. Green chemistry based benign approach for the synthesis of titanium oxide nanoparticles using extracts of Azadirachta Indica. Clean Eng Technol. 2023;13(2):1007–11.10.1016/j.clet.2023.100607Suche in Google Scholar
[28] Rao KG, Ashok CH, Rao KV, Chakra CS, Rajendar V. Synthesis of TiO2 nanoparticles from orange fruit waste. Synth (Stuttg). 2015;2(1):1.Suche in Google Scholar
[29] Aravind M, Amalanathan M, Mary MSM. Synthesis of TiO2 nanoparticles by chemical and green synthesis methods and their multifaceted properties. SN Appl Sci. 2021;3(4):1–10. 10.1007/s42452-021-04281-5.Suche in Google Scholar
[30] Kaur H, Goyal V, Singh J, Kumar S, Rawat M. Biomolecules encapsulated TiO2 nano-cubes using Tinospora cordifolia for photodegradation of a textile dye. Micro Nano Lett. 2019;14(12):1229–32.10.1049/mnl.2019.0340Suche in Google Scholar
[31] Khan M, Naqvi AH, Ahmad M. Comparative study of the cytotoxic and genotoxic potentials of zinc oxide and titanium dioxide nanoparticles. Toxicol Rep. 2015;2:765–74.10.1016/j.toxrep.2015.02.004Suche in Google Scholar PubMed PubMed Central
[32] Ahmad MM, Mushtaq S, Al Qahtani HS, Sedky A, Alam MW. Investigation of TiO2 nanoparticles synthesized by sol-gel method for effectual photodegradation, oxidation and reduction reaction. Crystals. 2021;11(12):1456.10.3390/cryst11121456Suche in Google Scholar
[33] Ahmad MZ, Alasiri AS, Ahmad J, Alqahtani AA, Abdullah MM, Abdel-Wahab BA, et al. Green synthesis of titanium dioxide nanoparticles using Ocimum sanctum leaf extract: in vitro characterization and its healing efficacy in diabetic wounds. Molecules. 2022;27(22):7712.10.3390/molecules27227712Suche in Google Scholar PubMed PubMed Central
[34] López-Mercado J, González-Domínguez MI, Reynoso-Marin FJ, Acosta B, Smolentseva E, Nambo A. Green synthesis of TiO2 for furfural production by photohydrolysis of tortilla manufacturing waste. Sci Rep. 2023;13(1):1–11. 10.1038/s41598-023-41529-z.Suche in Google Scholar PubMed PubMed Central
[35] Nguyen VT, Vu VT, Nguyen TH, Nguyen TA, Tran VK, Nguyen-Tri P. Antibacterial activity of TiO2-and zno-decorated with silver nanoparticles. J Compos Sci. 2019;3(2):61.10.3390/jcs3020061Suche in Google Scholar
[36] Yen L-T, Weng C-H, Than NAT, Tzeng J-H, Jacobson AR, Iamsaard K, et al. Mode of inactivation of Staphylococcus aureus and Escherichia coli by heated oyster-shell powder. Chem Eng J. 2022;432:134386, https://www.sciencedirect.com/science/article/pii/S138589472105957X.10.1016/j.cej.2021.134386Suche in Google Scholar
[37] Azizi-Lalabadi M, Ehsani A, Divband B, Alizadeh-Sani M. Antimicrobial activity of Titanium dioxide and Zinc oxide nanoparticles supported in 4A zeolite and evaluation the morphological characteristic. Sci Rep. 2019;9(1):1–10.10.1038/s41598-019-54025-0Suche in Google Scholar PubMed PubMed Central
[38] Nagalakshmi M, Karthikeyan C, Anusuya N, Brundha C, Basu MJ, Karuppuchamy S. Synthesis of TiO2 nanofiber for photocatalytic and antibacterial applications. J Mater Sci Mater Electron. 2017;28(21):15915–20.10.1007/s10854-017-7487-0Suche in Google Scholar
[39] Goutam SP, Saxena G, Singh V, Yadav AK, Bharagava RN, Thapa KB. Green synthesis of TiO2 nanoparticles using leaf extract of Jatropha curcas L. for photocatalytic degradation of tannery wastewater. Chem Eng J. 2018 Mar;336:386–96, https://linkinghub.elsevier.com/retrieve/pii/S1385894717321423.10.1016/j.cej.2017.12.029Suche in Google Scholar
[40] Nabi G, Ain Q-U-, Tahir MB, Nadeem Riaz K, Iqbal T, Rafique M, et al. Green synthesis of TiO2 nanoparticles using lemon peel extract: their optical and photocatalytic properties. Int J Env Anal Chem. 2022 Jan;102(2):434–42, https://www.tandfonline.com/doi/full/10.1080/03067319.2020.1722816.10.1080/03067319.2020.1722816Suche in Google Scholar
[41] Muniandy SS, Mohd Kaus NH, Jiang ZT, Altarawneh M, Lee HL. Green synthesis of mesoporous anatase TiO2 nanoparticles and their photocatalytic activities. RSC Adv. 2017;7(76):48083–94, https://xlink.rsc.org/?DOI=C7RA08187A.10.1039/C7RA08187ASuche in Google Scholar
[42] Singh A, Goyal V, Singh J, Rawat M. Structural, morphological, optical and photocatalytic properties of green synthesized TiO2 NPs. Curr Res Green Sustain Chem. 2020 Jun;3:100033, https://linkinghub.elsevier.com/retrieve/pii/S2666086520300369.10.1016/j.crgsc.2020.100033Suche in Google Scholar
[43] Sethy NK, Arif Z, Mishra PK, Kumar P. Green synthesis of TiO2 nanoparticles from Syzygium cumini extract for photo-catalytic removal of lead (Pb) in explosive industrial wastewater. Green Process Synth. 2020 Feb;9(1):171–81, https://www.degruyter.com/document/doi/10.1515/gps-2020-0018/html.10.1515/gps-2020-0018Suche in Google Scholar
[44] Saini R, Kumar P. Green synthesis of TiO2 nanoparticles using Tinospora cordifolia plant extract & its potential application for photocatalysis and antibacterial activity. Inorg Chem Commun. 2023 Oct;156:111221, https://linkinghub.elsevier.com/retrieve/pii/S138770032300833X.10.1016/j.inoche.2023.111221Suche in Google Scholar
[45] Li Y, Fu Y, Zhu M. Green synthesis of 3D tripyramid TiO2 architectures with assistance of aloe extracts for highly efficient photocatalytic degradation of antibiotic ciprofloxacin. Appl Catal B Env. 2020 Jan;260:118149, https://linkinghub.elsevier.com/retrieve/pii/S0926337319308963.10.1016/j.apcatb.2019.118149Suche in Google Scholar
© 2025 the author(s), published by De Gruyter
This work is licensed under the Creative Commons Attribution 4.0 International License.
Artikel in diesem Heft
- Research Articles
- Optimized green synthesis of silver nanoparticles from guarana seed skin extract with antibacterial potential
- Green adsorbents for water remediation: Removal of Cr(vi) and Ni(ii) using Prosopis glandulosa sawdust and biochar
- Green approach for the synthesis of zinc oxide nanoparticles from methanolic stem extract of Andrographis paniculata and evaluation of antidiabetic activity: In silico GSK-3β analysis
- Development of a green and rapid ethanol-based HPLC assay for aspirin tablets and feasibility evaluation of domestically produced bioethanol in Thailand as a sustainable mobile phase
- A facile biodegradation of polystyrene microplastic by Bacillus subtilis
- Enhanced synthesis of fly ash-derived hydrated sodium silicate adsorbents via low-temperature alkaline hydrothermal treatment for advanced environmental applications
- Impact of metal nanoparticles biosynthesized using camel milk on bacterial growth and copper removal from wastewater
- Preparation of Co/Cr-MOFs for efficient removal of fleroxacin and Rhodamine B
- Applying nanocarbon prepared from coal as an anode in lithium-ion batteries
- Improved electrochemical synthesis of Cu–Fe/brass foil alloy followed by combustion for high-efficiency photoelectrodes and hydrogen production in alkaline solutions
- Precipitation of terephthalic acid from post-consumer polyethylene terephthalate waste fractions
- Biosynthesized zinc oxide nanoparticles: Multifunctional potential applications in anticancer, antibacterial, and B. subtilis DNA gyrase docking
- Anticancer and antimicrobial effects of green-synthesized silver nanoparticles using Teucrium polium leaves extract
- Green synthesis of eco-friendly bioplastics from Chlorella and Lithothamnion algae for safe and sustainable solutions for food packaging
- Optimizing coal water slurry concentration via synergistic coal blending and particle size distribution
- Green synthesis of Ag@Cu and silver nanowire using Pterospermum heterophyllum extracts for surface-enhanced Raman scattering
- Green synthesis of copper oxide nanoparticles from Algerian propolis: Exploring biochemical, structural, antimicrobial, and anti-diabetic properties
- Simultaneous quantification of mefenamic acid and paracetamol in fixed-dose combination tablet dosage forms using the green HPTLC method
- Green synthesis of titanium dioxide nanoparticles using green tea (Camellia sinensis) extract: Characteristics and applications
- Pharmaceutical properties for green fabricated ZnO and Ag nanoparticle-mediated Borago officinalis: In silico predications study
- Synthesis and optimization of gemcitabine-loaded nanoparticles by using Box–Behnken design for treating prostate cancer: In vitro characterization and in vivo pharmacokinetic study
- A comparative analysis of single-step and multi-step methods for producing magnetic activated carbon from palm kernel shells: Adsorption of methyl orange dye
- Sustainable green synthesis of silver nanoparticles using walnut septum waste: Characterization and antibacterial properties
- Efficient electrocatalytic reduction of CO2 to CO over Ni/Y diatomic catalysts
- Greener and magnetic Fe3O4 nanoparticles as a recyclable catalyst for Knoevenagel condensation and degradation of industrial Congo red dye
- Recycling of HDPE-giant reed composites: Processability and performance
- Fabrication of antibacterial chitosan/PVA nanofibers co-loaded with curcumin and cefadroxil for wound healing
- Cost-effective one-pot fabrication of iron(iii) oxychloride–iron(iii) oxide nanomaterials for supercapacitor charge storage
- Novel trimetallic (TiO2–MgO–Au) nanoparticles: Biosynthesis, characterization, antimicrobial, and anticancer activities
- Green-synthesized chromium oxide nanoparticles using pomegranate husk extract: Multifunctional bioactivity in antioxidant potential, lipase and amylase inhibition, and cytotoxicity
- Therapeutic potential of sustainable zinc oxide nanoparticles biosynthesized using Tradescantia spathacea aqueous leaf extract
- Chitosan-coated superparamagnetic iron oxide nanoparticles synthesized using Carica papaya bark extract: Evaluation of antioxidant, antibacterial, and anticancer activity of HeLa cervical cancer cells
- Antioxidant potential of peptide fractions from tuna dark muscle protein isolate: A green enzymatic approach
- Clerodendron phlomoides leaf extract-mediated synthesis of selenium nanoparticles for multi-applications
- Optimization of cellulose yield from oil palm trunks with deep eutectic solvents using response surface methodology
- Nitrogen-doped carbon dots from Brahmi (Bacopa monnieri): Metal-free probe for efficient detection of metal pollutants and methylene blue dye degradation
- High energy density pseudocapacitor based on a nanoporous tungsten(VI) oxide iodide/poly(2-amino-1-mercaptobenzene) composite
- Green synthesized Ag–Cu nanocomposites as an improved strategy to fight multidrug-resistant bacteria by inhibition of biofilm formation: In vitro and in silico assessment study
- In vitro evaluation of antibacterial activity and associated cytotoxicity of biogenic silver nanoparticles using various extracts of Tabernaemontana ventricosa
- Fabrication of novel composite materials by impregnating ZnO particles into bacterial cellulose nanofibers for antimicrobial applications
- Solidification floating organic drop for dispersive liquid–liquid microextraction estimation of copper in different water samples
- Kinetics and synthesis of formation of phosphate composites from low-grade phosphorites in the presence of phosphate–siliceous shales and oil sludge
- Removal of minocycline and terramycin by graphene oxide and Cr/Mn base metal–organic framework composites
- Microfluidic preparation of ceramide E liposomes and properties
- Therapeutic potential of Anamirta cocculus (L.) Wight & Arn. leaf aqueous extract-mediated biogenic gold nanoparticles
- Review Article
- Sustainable innovations in garlic extraction: A comprehensive review and bibliometric analysis of green extraction methods
- Rapid Communication
- In situ supported rhodium catalyst on mesoporous silica for chemoselective hydrogenation of nitriles to primary amines
- Special Issue: Valorisation of Biowaste to Nanomaterials for Environmental Applications
- Valorization of coconut husk into biochar for lead (Pb2+) adsorption
- Corrigendum
- Corrigendum to “An updated review on carbon nanomaterials: Types, synthesis, functionalization and applications, degradation and toxicity”
Artikel in diesem Heft
- Research Articles
- Optimized green synthesis of silver nanoparticles from guarana seed skin extract with antibacterial potential
- Green adsorbents for water remediation: Removal of Cr(vi) and Ni(ii) using Prosopis glandulosa sawdust and biochar
- Green approach for the synthesis of zinc oxide nanoparticles from methanolic stem extract of Andrographis paniculata and evaluation of antidiabetic activity: In silico GSK-3β analysis
- Development of a green and rapid ethanol-based HPLC assay for aspirin tablets and feasibility evaluation of domestically produced bioethanol in Thailand as a sustainable mobile phase
- A facile biodegradation of polystyrene microplastic by Bacillus subtilis
- Enhanced synthesis of fly ash-derived hydrated sodium silicate adsorbents via low-temperature alkaline hydrothermal treatment for advanced environmental applications
- Impact of metal nanoparticles biosynthesized using camel milk on bacterial growth and copper removal from wastewater
- Preparation of Co/Cr-MOFs for efficient removal of fleroxacin and Rhodamine B
- Applying nanocarbon prepared from coal as an anode in lithium-ion batteries
- Improved electrochemical synthesis of Cu–Fe/brass foil alloy followed by combustion for high-efficiency photoelectrodes and hydrogen production in alkaline solutions
- Precipitation of terephthalic acid from post-consumer polyethylene terephthalate waste fractions
- Biosynthesized zinc oxide nanoparticles: Multifunctional potential applications in anticancer, antibacterial, and B. subtilis DNA gyrase docking
- Anticancer and antimicrobial effects of green-synthesized silver nanoparticles using Teucrium polium leaves extract
- Green synthesis of eco-friendly bioplastics from Chlorella and Lithothamnion algae for safe and sustainable solutions for food packaging
- Optimizing coal water slurry concentration via synergistic coal blending and particle size distribution
- Green synthesis of Ag@Cu and silver nanowire using Pterospermum heterophyllum extracts for surface-enhanced Raman scattering
- Green synthesis of copper oxide nanoparticles from Algerian propolis: Exploring biochemical, structural, antimicrobial, and anti-diabetic properties
- Simultaneous quantification of mefenamic acid and paracetamol in fixed-dose combination tablet dosage forms using the green HPTLC method
- Green synthesis of titanium dioxide nanoparticles using green tea (Camellia sinensis) extract: Characteristics and applications
- Pharmaceutical properties for green fabricated ZnO and Ag nanoparticle-mediated Borago officinalis: In silico predications study
- Synthesis and optimization of gemcitabine-loaded nanoparticles by using Box–Behnken design for treating prostate cancer: In vitro characterization and in vivo pharmacokinetic study
- A comparative analysis of single-step and multi-step methods for producing magnetic activated carbon from palm kernel shells: Adsorption of methyl orange dye
- Sustainable green synthesis of silver nanoparticles using walnut septum waste: Characterization and antibacterial properties
- Efficient electrocatalytic reduction of CO2 to CO over Ni/Y diatomic catalysts
- Greener and magnetic Fe3O4 nanoparticles as a recyclable catalyst for Knoevenagel condensation and degradation of industrial Congo red dye
- Recycling of HDPE-giant reed composites: Processability and performance
- Fabrication of antibacterial chitosan/PVA nanofibers co-loaded with curcumin and cefadroxil for wound healing
- Cost-effective one-pot fabrication of iron(iii) oxychloride–iron(iii) oxide nanomaterials for supercapacitor charge storage
- Novel trimetallic (TiO2–MgO–Au) nanoparticles: Biosynthesis, characterization, antimicrobial, and anticancer activities
- Green-synthesized chromium oxide nanoparticles using pomegranate husk extract: Multifunctional bioactivity in antioxidant potential, lipase and amylase inhibition, and cytotoxicity
- Therapeutic potential of sustainable zinc oxide nanoparticles biosynthesized using Tradescantia spathacea aqueous leaf extract
- Chitosan-coated superparamagnetic iron oxide nanoparticles synthesized using Carica papaya bark extract: Evaluation of antioxidant, antibacterial, and anticancer activity of HeLa cervical cancer cells
- Antioxidant potential of peptide fractions from tuna dark muscle protein isolate: A green enzymatic approach
- Clerodendron phlomoides leaf extract-mediated synthesis of selenium nanoparticles for multi-applications
- Optimization of cellulose yield from oil palm trunks with deep eutectic solvents using response surface methodology
- Nitrogen-doped carbon dots from Brahmi (Bacopa monnieri): Metal-free probe for efficient detection of metal pollutants and methylene blue dye degradation
- High energy density pseudocapacitor based on a nanoporous tungsten(VI) oxide iodide/poly(2-amino-1-mercaptobenzene) composite
- Green synthesized Ag–Cu nanocomposites as an improved strategy to fight multidrug-resistant bacteria by inhibition of biofilm formation: In vitro and in silico assessment study
- In vitro evaluation of antibacterial activity and associated cytotoxicity of biogenic silver nanoparticles using various extracts of Tabernaemontana ventricosa
- Fabrication of novel composite materials by impregnating ZnO particles into bacterial cellulose nanofibers for antimicrobial applications
- Solidification floating organic drop for dispersive liquid–liquid microextraction estimation of copper in different water samples
- Kinetics and synthesis of formation of phosphate composites from low-grade phosphorites in the presence of phosphate–siliceous shales and oil sludge
- Removal of minocycline and terramycin by graphene oxide and Cr/Mn base metal–organic framework composites
- Microfluidic preparation of ceramide E liposomes and properties
- Therapeutic potential of Anamirta cocculus (L.) Wight & Arn. leaf aqueous extract-mediated biogenic gold nanoparticles
- Review Article
- Sustainable innovations in garlic extraction: A comprehensive review and bibliometric analysis of green extraction methods
- Rapid Communication
- In situ supported rhodium catalyst on mesoporous silica for chemoselective hydrogenation of nitriles to primary amines
- Special Issue: Valorisation of Biowaste to Nanomaterials for Environmental Applications
- Valorization of coconut husk into biochar for lead (Pb2+) adsorption
- Corrigendum
- Corrigendum to “An updated review on carbon nanomaterials: Types, synthesis, functionalization and applications, degradation and toxicity”