Next Article in Journal
The Effect of Coronary Artery Bypass Surgery on Interleukin-18 Concentration and Biomarkers Related to Vascular Endothelial Glycocalyx Degradation
Previous Article in Journal
In Vitro Determination of Cytotoxic Effects of Ten Essential Oils on Prototheca bovis, Which Causes Mastitis in Dairy Cows
Previous Article in Special Issue
An Advanced Sensing Approach to Biological Toxins with Localized Surface Plasmon Resonance Spectroscopy Based on Their Unique Protein Quaternary Structures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Green Synthesis of Titanium Dioxide Nanoparticles: Physicochemical Characterization and Applications: A Review

1
Department of Materials Technology and Chemistry, Faculty of Chemistry, University of Łódź, Pomorska 163, 90-236 Lodz, Poland
2
Doctoral School of Exact and Natural Sciences, University of Łódź, Jana Matejki 21/23, 90-237 Lodz, Poland
3
Government Graduate College, Chowk Azam, Layyah, Multan 31-450, Pakistan
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(12), 5454; https://doi.org/10.3390/ijms26125454
Submission received: 12 April 2025 / Revised: 27 May 2025 / Accepted: 31 May 2025 / Published: 6 June 2025
(This article belongs to the Collection Feature Papers in Molecular Nanoscience)

Abstract

Nanotechnology is an emerging field in science that exhibits significant promise in the synthesis of nanomaterials for diverse applications. Traditionally, these nanomaterials were manufactured using hazardous and labor-intensive physical and chemical processes. Nevertheless, in recent years, researchers have developed safer, more scalable, and environmentally friendly methods for green synthesis. The problem addressed in this study is the need for an environmentally friendly and efficient synthesis process for titanium dioxide nanoparticles (TiO2 NPs) with enhanced properties. The aim of this work is to describe the synthesis of TiO2 NPs with various plant extracts using a green approach and to evaluate the physicochemical characteristics and potential applications of the resulting nanoparticles. This study focuses on understanding how the integration of plant extracts influences the properties of TiO2 NPs, particularly in terms of their structural, optical, and functional characteristics. The novelty lies in the use of plant extracts as bio-reductants and capping agents, which not only provides a safer and more sustainable synthesis method but also enhances the functional properties of TiO2 NPs. This green synthesis approach reduces the use of harmful chemicals, making the process more environmentally friendly and economically viable, with potential applications in photocatalysis, antibacterial, and antioxidant activities. The TiO2 NPs possess diverse functionalities, including photocatalysis, antibacterial properties, and antioxidant properties. The initial precursor, such as a metal salt, undergoes transformation into the desired nanoparticles through the actions of plants exactly. Bio-reduction and capping processes are carried out by secondary metabolites found in bacteria and plants. The results demonstrated that the plant extract-mediated TiO2 NPs exhibited enhanced photocatalytic activity, superior antibacterial effects, and higher antioxidant potential compared to chemically synthesized TiO2 NPs. This highlights the potential of green synthesis methods in producing nanomaterials with improved functional properties for a wide range of applications.

Graphical Abstract

1. Introduction

Global environmental degradation and an energy dilemma are currently endangering humanity’s long-term steady growth. Seeking high-efficiency and environmentally friendly protocols to address environmental pollution and energy constraint challenges has become an urgent endeavor [1,2]. It is critical for researchers to develop efficient, stable, and green solutions to successfully regulate environmental pollution and tackle energy concerns [3]. Figure 1 shows the potential applications of TiO2 nanoparticles in various fields of life.
Fujishima and Honda [4] were the first to demonstrate photocatalytic water splitting on a TiO2 electrode in 1972. Their work advances semiconductor photocatalysis, which is widely employed in the fields of environment and energy, opening the door to photocatalysis and attracting the attention of many scientific scholars. In recent years, the research and development of nano-TiO2 (nano-TiO2 commonly refers to particles smaller than 100 nm in at least one dimension) [5,6,7] photocatalysts has progressed rapidly, with increasing accomplishments. TiO2, one of the most researched photocatalysts, can only absorb UV light due to its relatively high band gap of 3.0–3.2 eV, which reduces photocatalysis efficiency [8,9].
It is intended to decompose waste products efficiently under UV light irradiation due to its band gap above 3.0 eV. However, due to its broad band gap, visible light activity is often hampered, and it is known that solar energy contains around 4–5% UV light and 50% visible light. As a result, with so little UV light reaching the earth’s surface, the usage of UV-active photocatalytic materials is not economically possible [10].
TiO2 mixed with metal NPs has been researched to overcome this problem, where the metal NPs enhance the catalytic capabilities of TiO2 by tuning the Fermi level of TiO2 and acting as an electron sink [11]. Various ways for changing their properties have been developed, including the formation of nanoparticles by the addition of other nanomaterials.
Many remarkable investigations have been conducted on the synthesis and modification of TiO2-based photocatalysts, as well as their applications in alleviating energy and environmental challenges [12,13,14]. However, few papers have classified the environmental uses of nano-TiO2 and compared the specific applications of nano-TiO2-based photocatalysts [15,16,17]. Phytochemical doping, surface state modifications, and heterojunction formation are some of the mechanisms that contribute to the significantly lower band gaps seen in plant-mediated TiO2 nanoparticles (NPs) when compared to the typical anatase phase (~3.2 eV), rather than just quantum size effects. Phytochemicals such as flavonoids and polyphenols can replace oxygen sites in TiO2 by acting as sources of carbon (C) or nitrogen (N). For instance, the hybridization of N 2p and O 2p orbitals results in a smaller band gap in N-doped TiO2 [18,19]. Extracts such as Taraxacum officinale include functional groups (-OH, -COOH) that stabilize oxygen vacancies (Vo), which lower the effective band gap by introducing defect states below the conduction band [20]. By increasing interfacial interactions between TiO2 and secondary phases (such as carbon coatings), phytochemicals such as inulin from dandelion roots reduce recombination and alter band edges [21]. As demonstrated in Ag-doped TiO2 NPs, where plasmonic effects predominate, plant-mediated synthesis frequently counteracts the effects of quantum confinement, which normally increases band gaps in ultrasmall NPs (<5 nm), by inducing surface states that are mitigated by phytochemicals [22]. The band gap of TiO2 was modified to improve its photocatalytic activity by doping it with metal and nonmetal ions, introducing vacancy, and constructing composites with other semiconductors [23].
Various synthesis approaches, such as chemical, physical, and green methods, have been utilized in the manufacture of NPs (as shown in Figure 2). These approaches are now encountering issues such as aggregation, stability, crystal growth control, size, shape, and distribution due to their ongoing development stage. In addition, the separation of the generated NPs is a crucial factor to be considered for future applications. The use of bacteria, actinomycetes, yeasts, fungi, algae, and plant extracts in green synthesis provides a superior alternative to chemical and physical methods in the development of green technology. Furthermore, it has the capability to avoid the utilization of dangerous chemicals, is cost-effective and environmentally friendly, and can be easily expanded for large-scale synthesis [24]. It is more energy- efficient without the need for high pressure and temperature [25]. While chemical and physical procedures are commonly employed for the manufacture of NPs, they necessitate the utilization of reactive and poisonous reducing agents [26].
Efforts are ongoing to establish a straightforward, dependable, and efficient method for producing NPs through environmentally friendly synthesis processes. The collective endeavors of all researchers are crucial in generating stable, highly reliable, and well-functionalized NPs. Hence, it is imperative to investigate a sustainable methodology to produce NPs. The concern may encompass the economic viability, environmental sustainability, social adaptation, and availability of local resources. Green synthesis methods commonly involve natural, sustainable materials such as plant extracts, microorganisms, or other biocompatible substances. This method decreases the utilization of dangerous chemicals and solvents linked to conventional industrial techniques, leading to a decreased environmental footprint. Nanoparticles of TiO2 synthesized through environmentally friendly techniques exhibit greater biocompatibility in comparison to those manufactured using traditional chemical methods.
Although the goal of both green and chemical synthesis methods is to create TiO2 nanoparticles, these approaches vary significantly in terms of their environmental impact, safety, scalability, and the characteristics of the materials and processes used. Green synthesis methods are preferred due to their sustainability and potential for biocompatibility, which makes them more appealing for applications that priorities eco-friendliness and safety. When TiO2 nanoparticles are produced through environmentally friendly methods (such as using plant extracts or microorganisms) the resulting nanoparticles maintain their original chemical properties. The selection of the synthesis technique can impact distinct attributes such as particle dimensions, structure, and surface characteristics. However, the fundamental chemical properties of TiO2 remain constant regardless of the chosen synthesis approach.
The purity and crystalline phase of TiO2 NPs produced through environmentally friendly methods may differ based on the synthesis conditions and parameters. The purity of the resulting TiO2 NPs can be affected by the purity of the starting materials, such as the titanium precursor, and the efficiency of the green synthesis process in reducing impurities. By manipulating synthesis parameters such as the reaction temperature, precursor concentration, pH of the reaction mixture, and nature of the green reducing or stabilizing agents, one can regulate the crystalline phase of TiO2 NPs, including anatase, rutile, or mixed phases. Phytochemicals (such as flavonoids, polyphenols, and -OH and -COOH groups) found in plant extracts serve as capping and reducing agents. These biomolecules regulate growth and nucleation to stabilize stages. Because of their strong chelation of Ti4+ surface sites, which slows hydrolysis and allows for ordered atomic organization, extracts high in polyphenols (such as Tinospora cordifolia [27] and Impatiens rothii [28]) favor anatase. By changing surface energy or promoting phase transitions at higher temperatures, amine-rich extracts may encourage rutile [29]. Because of kinetic control, green technologies frequently generate anatase at moderate temperatures (e.g., 50–150 °C) [30]. Phase transitions between anatase and rutile are triggered by post-synthesis heating (>500 °C). For instance, mixed anatase–rutile phases with increased photocatalytic activity were observed in TiO2 NPs calcined at 500 °C [31].
Titanium dioxide is categorized as a substance with a moderate suspicion of causing cancer through inhalation (Category 2) according to the Classification, Labelling, and Packaging (CLP) Regulation. Regulation (CLP) (EC) No. 1272/2008 imposes precise labeling obligations for products containing TiO2, which include the use of hazard pictograms, signal words, and precautionary statements. The disparity in waste generation between green and industrial TiO2 synthesis methods primarily stems from the sustainability and environmental factors associated with the respective processes. Green synthesis methods aim to reduce waste and environmental impact by utilizing renewable resources and eco-friendly processes. In contrast, industrial methods can produce more waste and have a larger environmental footprint due to their larger scale and reliance on traditional chemical processes. Utilizing green synthesis methods can thus aid in achieving a more sustainable and environmentally conscientious production of TiO2. Green synthesis decreases expenses by utilizing plant extracts, frequently derived from agricultural waste, while minimizing dangerous chemicals and reducing energy demands relative to conventional processes. Agricultural by-products are utilized for nanoparticle manufacturing, hence minimizing waste. The procedure mitigates environmental effects by recycling natural resources and reducing toxic chemicals, promoting long-term sustainability and supporting a circular economy [32,33,34,35]. The circular economy aspect of chemical and specific green synthesis (based on plant extracts like Rhizophora apiculata, Ocimum sanctum and Butea monosperma) is shown in Table 1.
Waste containing TiO2 must be handled and disposed of in compliance with the Waste Framework Directive (WFD) (Directive 2008/98/EC) to minimize environmental effects and ensure appropriate waste management and recycling. The energy balance in the green synthesis of TiO2 entails a comprehensive assessment of energy inputs, process efficiency, the utilization of renewable energy, and environmental impacts to guarantee the sustainability and eco-friendliness of production methods. Prioritizing the reduction of energy consumption and the utilization of renewable resources are crucial in green synthesis methods for TiO2 NPs.
However, numerous obstacles persist in the eco-friendly production of TiO2 NPs. A primary problem is the reproducibility of results, as discrepancies in plant species, cultivation conditions, and extraction techniques might result in inconsistencies in nanoparticle size and functioning [36]. By multipurpose interactions that are absent in traditional chemical synthesis, plant-derived capping agents (such as polyphenols, -COOH, and -OH groups) improve the durability and reusability of TiO2 nanoparticles (NPs). By physically adhering to TiO2 surfaces, large biomolecules (such as proteins and polysaccharides) form a dense organic shell that inhibits aggregation through steric hindrance. Aloe vera polysaccharides, for instance, envelop NPs in a hydrogel-like layer that defies van der Waals forces [37]. Undercoordinated Ti3+ surface sites are saturated by plant biomolecules (such as flavonoids and tannins) that chelate Ti atoms via –OH/–COOH groups. This preserves NP integrity during reuse by lowering oxidative degradation and ROS production (such as OH radicals) [38]. By scavenging free radicals, polyphenols (such as quercetin and catechins) protect TiO2 surfaces from photocorrosion. This defense is absent from chemically produced NPs, which break down more quickly in reactive settings [39]. Moreover, expanding the green synthesis technique for industrial use while preserving its ecological advantages presents considerable technical challenges. The precise mechanisms by which plant metabolites promote nanoparticle production are poorly understood, complicating optimization efforts.
Despite these challenges, the prospects in this domain are extensive. The utilization of plant extracts for nanoparticle synthesis reduces the reliance on harmful chemicals and facilitates the utilization of agricultural waste or underutilized plant species, thereby establishing a sustainable cycle. The capacity to refine the characteristics of TiO2 NPs by choosing plant extracts offers promising prospects for specialized applications in photocatalysis, environmental remediation, antimicrobial therapies, and biomedical devices [40].
The prospects of this research appear favorable. Improvements in comprehending the function of various phytochemicals in nanoparticle formation may result in more predictable and controllable methodologies. Furthermore, integrating plant-mediated synthesis with alternative green technologies, such as solar or microwave-assisted synthesis, may augment efficiency and scalability.
This review focuses on environmentally friendly methods for synthesizing TiO2 NPs. It also analyses the vast range of plant species that can be used in this process. Utilizing green methods not only decreases the expense of synthesis but also lessens the necessity for dangerous chemicals and promotes environmentally friendly synthesis. The novelty of this research comes in its in-depth comparison of multiple plant-based green synthesis strategies, stressing the specific impact of distinct plant species on the morphology, size, and functional aspects of TiO2 NPs. Unlike previous studies that focused on a limited range of plant extracts or conventional synthesis methods, this review introduces the concept of tailoring nanoparticle properties through the selection of specific plant metabolites, offering insights into how the unique biochemical composition of each plant can optimize TiO2 NPs for applications. This strategy presents new paths for increasing nanoparticle efficacy in sectors like photocatalysis and antibacterial activities. These discoveries could motivate researchers to explore more diversified natural resources and develop improved, secure, and cost-effective nanomaterial production processes for a wider range of industrial and biological applications.
This paper specifically examines the distinctive characteristics that come from the output of this environmentally friendly technique, which enhances its suitability for certain applications such as photocatalysis and antimicrobial uses. These remarkable discoveries are likely to inspire researchers and attract new individuals to further investigate and expand the potential applications of natural resources and the development of advanced and more secure techniques for creating nanomaterials. Figure 2 gathers the main approaches applied in the preparation of TiO2 nanoparticles, including the green method.

2. Synthesis of TiO2 NPs from Green Plant Extracts

The synthesis of TiO2 NPs presents challenges that can impact their structure, properties, and performance, especially for applications like photocatalysis, solar cells, and sensors. Attaining exact control over phase formation during synthesis is difficult [41,42]. Methods such as sol–gel synthesis, hydrothermal processing, and chemical vapor deposition (CVD) can be intricate and costly, complicating the scalability of manufacturing for commercial applications [43]. Synthesizing TiO2 at the nanoscale can markedly augment its features, such as elevated surface area and enhanced photocatalytic activity. Nonetheless, regulating the dimensions and consistency of nanoparticles poses significant challenges [44]. A limited number of pilot-scale studies have demonstrated the feasibility of scaling up such processes [45]. However, significant bottlenecks persist. Moreover, unlike conventional chemical synthesis, green methods often suffer from lower production efficiency and require careful optimization of pH, temperature, and precursor concentration to maintain consistency [46]. Standardization of extract preparation and the establishment of quality control protocols remain critical hurdles for scaling up these methods reliably. Until these challenges are systematically addressed, the commercial viability of green TiO2 synthesis remains limited to small-scale applications [47].
Large-scale productivity is further constrained by slower reaction kinetics in comparison to chemical synthesis [48]. Although green synthesis of TiO2 is economically appealing and theoretically feasible, practical scale-up is still in early stages. The standardization of plant extracts, the repeatability of the process, clogging, and yield optimization are major obstacles. To transfer lab-scale gains to industry, future research should concentrate on reactor engineering, extract characterization, and techno-economic modelling.
The synthesis of preferred TiO2 phases frequently necessitates high-temperature processing, resulting in elevated energy consumption and production expenses [49]. Certain synthesis pathways utilize hazardous solvents or precursors, presenting environmental and safety issues. The pursuit of more environmentally friendly and sustainable synthesis processes remains a persistent task [50]. Plant extracts’ functional groups balance reduction, capping, and stabilization to control the synthesis of TiO2 NPs. By giving electrons to Ti4+ surface sites, -OH groups function as reducing agents (such as polyphenols and flavonoids) and start nucleation. Smaller particle sizes are preferred due to the large number of tiny nuclei produced by rapid reduction. Adsorbing onto TiO2 surfaces, bulky biomolecules such as polysaccharides (found in aloe vera, for example) physically prevent aggregation and restrict development. As a result, NPs have a large surface area and are stable. Surface defects are covered by strong Ti–O bonds made by –OH and –COOH groups (found, for example, in Cannabis sativa), which inhibit oxidative degradation and Ostwald ripening. Green tea polyphenols and other extracts high in -OH groups quickly reduce Ti4+, generating a large number of nucleation sites. Growth is limited to sub-10 nm diameters because of competition for scarce Ti4+. Atoms can organize into ordered lattices due to slow ion release via –COOH complexation (e.g., aloe vera extracts), which promotes the production of anatase [51]. By forming compounds with a Ti4+ center (like citric acid),–COOH groups slow down nucleation through regulated ion release. This promotes consistency by resulting in fewer, well-defined nuclei. Citrus peel is one example of an extract that is rich in –COOH and forms stable Ti–carboxylate complexes that tightly cap NPs and inhibit growth [52]. Certain crystal facets are stabilized by functional groups. For instance, extracts rich in amines encourage rutile under calcination, whereas extracts high in –OH favor anatase [53].
The chemical approaches for the synthesis of TiO2 NPs have been found to be harmful to the environment, and the synthesis demands high temperatures and pressure, limiting TiO2 mass manufacturing. As a result, green nanotechnology has been investigated as an alternative and environmentally responsible strategy for TiO2 NPs synthesis since it employs reducing agents obtained from plants sources, and the same reducing agent may be used for the synthesis of numerous metallic compounds. The use of plants, their waste materials, fruit extracts, and microorganisms in synthesis reduces the use of hazardous and costly chemicals [54]. Furthermore, it can be used for low-cost mass synthesis of TiO2 NPs [55]. Green synthesis approaches are therefore critical for producing stable, suitably sized, and dispersible NPs with low energy consumption [56]. The advantages of safety and feasibility commonly lead to the use of plant extract as a significant component in the production of TiO2 NPs. Plants contain proteins, carbohydrates, enzymes, phenolic acids, and alkaloids, all of which perform reduction and stabilization activities throughout synthesis. Moreover, leaves have a high concentration of metabolites, and they are widely used in green synthesis. The utilization of plant extracts in the processes of nanoparticle production shows more stability due to their higher reaction rate as a reducing agent (capable of reducing metal or metal oxide ions) and as a stabilizing agent, surpassing that of fungus, bacteria, or other organisms. In addition, plant extract gives a significant benefit due to the organism’s need for precise control over cell culture. The green technique is a viable option for scaling up the production of well-dispersed NPs in an industrial setting [24]. Figure 3 shows the sources of extracts for green synthesis of TiO2 and its applications.
There are many recent examples of successful synthesis procedures where plant extracts were used, e.g., TiO2 NPs synthesized from several Commenlina benghalensis plant concentrations [57]. Spherical and agglomerated NPs with a 30–200 nm particle size dispersion and 3.69–3.80 eV band gap were produced. The materials’ surfaces became smoother and more spherical as plant concentration rose. At the greatest concentration, surface area increased from 18 to 174 m2/g. The authors of [52] created spherical anatase TiO2 NPs with a size range of 15–28 nm utilizing peels, flowers, and leaves. Different extracts affected the size and optical characteristics of the materials. Tulbhagia violacea was utilized by Mbenga et al. [58] to create irregular and rectangular NPs. These anatase-structured materials were highly crystalline and varied in shape. Shimi et al. [59] prepared spherical TiO2 materials from Phyllanthus niruri leaves with a band gap of 2.8 eV. Despite several works of research showing a lower TiO2 band gap, there is no specific synthesis pattern to explain it. Perhaps some phytochemicals found in certain plants could affect optical qualities like excitation and band gap shifts. Kaur et al. generated irregularly shaped TiO2 NPs with 303 nm particle sizes and 2.1 eV band gap from Alcea plant flowers [60]. Aravind et al. prepared TiO2 NPs capped and reduced with jasmine flower extract. The rutile phase was confirmed by XRD, while SEM showed randomly organized spherical TiO2 NPs having a size of 31–42 nm. UV–Vis measurements revealed the excitation wavelength at 385 nm [61]. Sethy et al. synthesized TiO2 NPs from Syzygium cumini leaf extract. A strong absorption peak at 356 nm and 3.48 eV band gap were found. Synthesized TiO2 NPs were round, irregular, and had an average size of 18 nm in diameter. They removed lead from wastewater with 82.53% efficiency [62]. Arabi et al. formed irregular and spherical TiO2 NPs with under 15 nm particle sizes using Thyme and Alcea [63]. Thakur et al. synthesized spherical, 124 nm sized TiO2 NPs from A. indica leaf extract; UV–Vis confirmed absorption in the range wavelength of 270–320 nm [64]. Pomegranate peel was used to synthesize TiO2 NPs, which disinfected water without harming the environment [65].

3. The Mechanism of Synthesis of TiO2 NPs Prepared with the Use of Green Methods

Recently, researchers have focused on understanding the mechanisms involved in green synthesis of TiO2. Currently, the precise process by which NPs are produced through plant extract is not fully understood. However, numerous studies have demonstrated that certain biomolecules found in plant materials, such as alkaloids, polysaccharides, alcoholic, and phenolic compounds, may play a role in the creation, stabilization, and reduction of these NPs [66]. The formation of NPs is influenced by the capacity of plants to reduce ions and the reduction potential of these ions. This process is dependent on the presence of biomolecules in plants. Nasrollahzadeh and Sajadi [67] have identified a potential method for the bio-reduction of titanyl hydroxide, as depicted in Figure 4.
Initially, they conducted FTIR analysis to ascertain the presence of biomolecules. Consequently, the researchers determined that the hydroxyl groups of phenolics in the root extract of E. heteradena Jaub are accountable for the reduction of TiO(OH)2 and serve as capping ligands on the surfaces of the NPs. Moreover, the FT-IR data clearly indicates distinct variations in the position and morphology of signals, suggesting a connection between the phytochemicals and the sites involved in the green synthesis process. This implies that polyphenolics are likely to be adsorbed onto the surface of the NPs. Biomolecules such as flavonoids, alkaloids, polyphenols, terpenoids, heterocyclic compounds, and polysaccharides have been found to play important roles in reducing, stabilizing, and acting as capping agents for highly uniform NPs. The authors of [68] demonstrated that carboxylic acids found in Aloe vera leaf extract, which are identified by broad -OH stretching in FTIR analysis, indicate the presence of terpenoids, flavonoids, and proteins. These compounds can aid in the formation and production of biosynthesized TiO2 NPs. Additionally, they act as both a capping and reducing agent during the biosynthesis process. The work conducted by Goutam et al. [69] demonstrated the existence of hydroxyl groups in Jatropha curcas L., suggesting that the surface of TiO2 NPs synthesized through biological means was covered by phenols or polyphenolic tannins. Even though the biochemical mechanism is unknown, those biochemical substances are crucial to the reduction, stabilization, and capping of the reaction. Understanding the mechanisms will boost nanoparticle productivity. Figure 4 presents the formation mechanism of TiO2 NPs obtained from two types of precursors: titanyl hydroxide and titanium tetrachloride, respectively.
Figure 4. Possible reaction mechanism for the formation of biosynthesized TiO2 NPs from different Ti precursors ((A): Titanyl hydroxide, (B): TTIP, and (C): TiCl4) [69].
Figure 4. Possible reaction mechanism for the formation of biosynthesized TiO2 NPs from different Ti precursors ((A): Titanyl hydroxide, (B): TTIP, and (C): TiCl4) [69].
Ijms 26 05454 g004aIjms 26 05454 g004b

4. Limitations of Green Method for NPs Synthesis

The process of green synthesis involves the utilization of environmentally friendly and biocompatible ingredients as both reducing agents and stabilizing agents during the manufacture of NPs. However, there are several limitations to green synthesis, some of which are the following:
  • Green synthesis is hindered by the inability to precisely manipulate nanoparticle (NP) size and form depending on reaction duration, temperature, pH, and reactant concentration [70].
  • Green synthesis has a lower NPs yield than conventional methods. The instability of the reducing agent, poor precursor ion reduction, and agglomerates can cause low yields. When industrial nanoparticle demand is high, insufficient production output can be a major issue [71,72].
  • Green synthesis can struggle to scale up production to synthesize large volumes of NPs. Scalability may be constrained by the limited availability and unpredictability of natural extracts or microbes used as reducing agents [24,71].
  • Green synthesis also faces nanoparticle stability and surface chemistry issues. Nanoparticle surface chemistry greatly affects their interactions with biological systems and other materials. The use of natural extracts or microorganisms as reducing agents can make nanoparticle surface chemistry difficult to manage. Natural extracts used as reducing agents can also absorb contaminants onto nanoparticle surfaces, changing their characteristics [73].

5. Characterization of TiO2 NPs

The primary objective of nanoparticle characterization is the examination of critical parameters including charge, size, shape, and degree of aggregation. Various techniques are frequently employed to characterize NPs, including Fourier transform infrared spectroscopy (FTIR), X-ray photoelectron spectroscopy (XPS), energy-dispersive X-ray spectroscopy (EDS), and scanning electron microscopy (SEM). Additionally, transmission electron microscopy (TEM) and dynamic light scattering (DLS) are also utilized.

6. Optical Properties of TiO2 NPs

Quantification of the amount of light that is absorbed and scattered by a sample, e.g., TiO2, is a technique that is frequently utilized these days. When a sample is positioned between the UV–Vis light source and the detector, light is permitted to fall on and go through the sample. The intensity of the beam of light is measured both before and after it passes through the sample [74]. UV–Vis spectroscopy can be considered as a method that not only confirms the presence of TiO2 NPs in aqueous solution as an effect of green preparation but first shows the absorption intensity (and range) of the resulting material. Different scientific research indicated that titanium ions are reduced, and TiO2 NPs are generated within a few hours at room temperature. The solution turns light green when titanium ions decrease. Absorption peaks at 245–265 nm are seen in plant extract solutions used for TiO2 NPs preparation, while chemically synthesized TiO2 NPs exhibit a little absorbance shift, with a peak at 303 nm (λ-max).

7. Functional Group Characterization of TiO2 NPs Prepared Using Green Method

Fourier transform infrared spectroscopy (FTIR) is employed to analyze the impact of reducing and stabilizing agents on the surface chemistry of nanomaterials. Additionally, it plays a crucial role in elucidating the chemical bonding present within materials and the functional molecules that are attached to the surface of nanomaterials. In FTIR analysis, the samples undergo exposure to infrared (IR) radiation. The IR radiation subsequently exerts influence on the atomic vibrations of a molecule within the specimen, leading to the selective absorption and/or transmission of energy [75].
FT-IR spectroscopy was utilized to analyze the synthesized TiO2 NPs obtained using various methods, including green synthesis. The purpose was to identify the potential functional groups contained in the synthesized samples. The FT-IR spectra were recorded within the wavelength range of 4000 to 400 cm−1. It has been observed that titanium dioxide NPs synthesized using a green technique exhibit a comparable FTIR spectrum. Typically, samples exhibiting a green color demonstrated an absorption band within the spectral region of 650–700 cm−1, indicative of the stretching link between titanium (Ti) and oxygen (O) atoms (Ti-O-Ti). The absorption peak observed within the wavenumber range of 650–700 cm−1 is recognized as the distinctive peak of the anatase phase of TiO2 NPs, as reported in the literature [62,76]. The observed peaks within spectral range of 1000–1300 cm−1 can be attributed mostly to Ti-O-Ti vibrations, indicating the presence of O-Ti-O bonding. Additional carbon functional groups, such as CH, CH2, CH3, or O-CH3, may also be observed within the mentioned frequency range. These functional groups could potentially originate from plant extracts. An intriguing distinction may be observed in the synthesis processes, namely in the choice of precursors employed during the synthesis process. The distinctive characteristic of green TiO2 NPs can be attributed to the hydroxyl (-OH) groups found in leaf extracts of plants, which are present because of the prevalence of phenols. The likely explanation for the presence of the phenolic group in both extracts is the presence of polyphenolic tannins. These tannins are found on the surface of TiO2 NPs that were synthesized using environmentally friendly materials [77,78,79]. Surface-bound biomolecules derived from plant extracts improve the long-term stability of TiO2 NPs by inhibiting aggregation through steric and electrostatic interactions, as well as chelation [20]. They enhance biocompatibility by decreasing reactive oxygen species (ROS) and cytotoxicity by surface passivation, as shown by PEG-coated TiO2 nanoparticles [80]. These biomolecules might improve visible-light catalytic activity by dye sensitization (e.g., chlorophyll) or electron transfer (–OH groups); nevertheless, excessive organic layers can block active sites. Functional groups confirmed by FTIR (e.g., C=O, –OH stretching) are associated with these effects, as seen in TiO2 synthesized from orange peel extract [20].
Figure 5 shows an example of the FT-IR spectra of TiO2 NPs synthetized with the use of plant extracts.

8. Morphology of TiO2 NPs

The scanning electron microscope (SEM) is an instrument that utilizes a focused electron beam to scan across a surface to generate a picture. The scanning electron microscope (SEM) and transmission electron microscope (TEM) vary fundamentally in their operational principles. While TEM involves the transmission of electrons through a sample, SEM operates by directing a beam of electrons onto the surface of the sample and collects secondary electrons (SE) and/or backscattered electrons (BSE) in detectors to create an image. Through interactions between the electron beam and the sample, an image is generated pixel by pixel. The SEM preserves the three-dimensional topography of the material because of its greater depth and width compared to the TEM [81]. The green TiO2 NPs are composed of porous structures that exhibit relatively large crystals, displaying an intriguing and distinctive surface appearance. Due to their porous architecture, TiO2 NPs exhibit significant potential for revolutionary applications within the environmental applications. The distinctive composition of the leaf extracts of the plant material is attributed to the presence of phytochemicals, including terpenoids, steroids, polyphenols, flavonoids, alkaloids, antioxidants, free amino acids, and tannins that affect the topography and morphology of TiO2 NPs. In this context, the investigation of NPs morphology with the use of microscopy techniques is indispensable. In a recent study conducted by Shimpi et al., the authors successfully synthesized TiO2 NPs using a solution of titanium tetra-isopropoxide (TTIP) as a precursor, together with a leaf extract obtained from Murraya koenigii, often known as the Curry tree. The researcher utilized an aqueous extract of a plant and an ethanolic leaf extract to synthesize TiO2 NPs. A comparison was made between the obtained results and the standard parameters. It was observed through FESEM and TEM analysis that the NPs produced from the ethanolic leaf extract exhibited a perfectly spherical shape with a smaller size range of 2–15 nm. In contrast, the NPs synthesized using the aqueous leaf extract displayed a scattered spherical crystallite shape with a larger size range of 15–30 nm. In addition, the author also conducted a study of TiO2 NPs synthesized using chemical and green processes, revealing a disparity in their respective band gap energies of 4.3 eV and 4.5 eV [82].
The chemical production process exhibits the highest percentage of yield, whereas the percentage of TiO2 NPs is comparatively lower. This method of synthesis can be considered environmentally friendly and sustainable. To obtain TiO2 NPs from environmentally friendly sources, a method that eliminates precursors or reduces synthetic chemicals is needed [62,76,83]. Figure 6 presents the morphology (SEM) and chemical composition (EDX) of TiO2 NPs obtained using flower, leaf, and peel extracts of various plants.

9. Analysis of Crystallographic and Phase Structure of TiO2 NPs

The technique employed for the characterization of NPs is widely utilized. X-ray diffraction (XRD) is a commonly used technique in materials science that offers valuable insights on the structural characteristics, dimensions, composition, and lattice parameters of a material. The powdered material, which has been thoroughly dried, is utilized in X-ray diffraction (XRD) analysis.
The TiO2 NP standards describe the tetragonal anatase phase of crystallite, with space group I41/amd and lattice parameters a = b = 3.792 Å and c = 9.554 Å. The TiO2 NPs demonstrate a prominent peak observed at the 2θ value ranging from 24° to 28°, with a specific occurrence at 25.6°, which corresponds to the anatase (101) crystallographic plane. The XRD graphs revealed additional minor peaks at 2θ values of (38°, 48°, 54°, and 63°), corresponding to the (112), (200), (211), and (204) planar angles of the anatase phase. The XRD pattern of crystals indicates that this value is associated with the anatase form of TiO2 NPs, along with the phases of brookite and rutile. Furthermore, it is worth noting that the observation of the anatase form of TiO2 NPs at a certain plane angle is consistent across the study community. The property of TiO2 NPs, namely the 2θ peak at 25.6°, is not influenced by the synthesis process, as demonstrated in a study conducted by [76]. The X-ray diffraction (XRD) pattern of TiO2 NPs synthesized using plant extracts from T. portulaca strum and C. quinoa was compared with the sol–gel approach. The results demonstrate that the synthesis method does not influence the crystallinity of TiO2 NPs. In addition, [83,84] synthesized TiO2 NPs using extracts from Carica papaya and Cymbopogon citrates, respectively, and the XRD study yielded similar results. In addition to crystallinity, the average particle size of green TiO2 NPs has mostly been determined using the Debye–Scherrer equation (1) after successful synthesis [69,85]. The Scherrer equation is commonly regarded as the primary and extensively employed equation for determining particle size through the utilization of 2θ and FWHM values derived from X-ray diffraction (XRD) data.
D = Kλ βcosθ
In the given equation, the symbol D is used to indicate the particle size, specifically the diameter. The symbol K represents Scherrer’s constant, which has a value of 0.9. The symbol λ denotes the wavelength of the X-ray source, specifically 0.15406 nm. The symbol b is used to represent the full width at half-maximum intensity (FWHM) in radians. Lastly, the symbol q is utilized to designate the peak locations, measured in radians. Figure 7 presents the XRD patterns of green synthesized TiO2 NPs with a JCPDS card and Figure 8 presents the XRD patterns of TiO2(1), TiO2(2), and TiO2(3) at different T. violacea bulb extract volumes of 10, 20, and 40 mL, respectively.

10. Applications for Green TiO2 NPs

TiO2 NPs synthesized using green methods have been investigated for their potential applications in dye-sensitized solar cells (DSSCs), as well as in various biomedical fields such as drug delivery, imaging, and antimicrobial treatments. Additionally, these nanoparticles have shown effectiveness in removing heavy metals and other pollutants from polluted water and soil. Their photocatalytic characteristics allow for the decomposition of organic pollutants, while their large surface area facilitates the adsorption of metal ions. This distinctive behavior of TiO2 NPs differentiates it from other metallic nanoparticles. The comparison of the application of green synthesized TiO2 NPs and other green synthesized metal and metal oxide nanomaterials is presented in Table 2.
One of the important current applications of TiO2 NPs is the ability to produce reactive oxygen species (ROS) that subsequently take part in water purification processes. ROS generated by titanium dioxide can effectively degrade organic pollutants and disinfect water, making it a promising technology for clean drinking water production. Additionally, the future applications of titanium dioxide’s ROS include its potential use in cancer treatment, as ROS have been shown to induce cell death in cancer cells while sparing healthy cells.

11. Antimicrobial Activity of Green Synthesis TiO2 NPs

Antimicrobial drugs inhibit or eradicate the proliferation of germs. Extensive research has been conducted on the antibacterial characteristics of metal oxide NPs, specifically TiO2, yielding encouraging outcomes. Psidium guajava-derived TiO2 NPs were evaluated for their antimicrobial activity against human pathogenic bacteria using the disc diffusion method. The antioxidant activity of TiO2 NPs is significantly higher than that of ascorbic acid. Staphylococcus aureus (S. aureus) exhibited the highest level of zone inhibition, measuring 25 mm, while E. coli had a little lower zone inhibition of 23 mm. The synthesized TiO2 NPs exhibited superior antibacterial efficacy compared to the typical antibiotic disc, tetracycline, resulting in a considerable reduction in the effectiveness of antibiotics in protecting against bacterial infections [86]. The antibacterial effectiveness of TiO2 NPs synthesized using Azadirachta indica leaf extract was investigated against various pathogens, including E. coli, Klebsiella pneumonia, Salmonella typhi (S. typhi), and Bacillus subtilis (B. subtilis). The NPs exhibited the lowest Minimum Inhibitory Concentrations (MIC) value of 10.42 μg/mL against E. coli and S. typhi. However, the lowest Minimum bactericidal concentration (MBC) value of 83.3 μg/mL was seen against K. pneumonia [64]. The synthesis of green TiO2 NPs from Mentha arvensis leaf extract results in the manifestation of antifungal and antibacterial properties against several microorganisms including Aspergillus niger (A. niger), Proteus vulgaris (P. vulgaris), E. coli, Arthrographis cuboid, S. aureus, and Aspergillus fumigatus (A. fumigatus). The TiO2 NPs exhibit remarkable antibacterial efficacy against P. vulgaris, owing to their ability to dissolve the outermost bacterial membrane, which is the primary cause of bacterial demise [87]. Various laboratory techniques were employed to evaluate and test the antibacterial properties of an extract derived from a valuable chemical. In addition to commonly used and essential techniques like agar dilution and disc diffusion, other methods such as poisoned food are employed for antifungal testing. The antibacterial activity of green synthesized TiO2 NPs is attributed to the biomolecules derived from the Kniphofia foliosa root extract. Furthermore, it donates an extra electron to the surface of TiO2, resulting in the formation of superoxide radicals O2•−. This, in turn, leads to the production of reactive oxygen species (ROS) within the bacterial cell. This ROS disrupts the integrity of bacterial cell membranes. Elevating the production of the superoxide radical also enhances the generation of reactive oxygen species (ROS) within the bacterial cell. The generation of electrons from TiO2 NPs is responsible for the antibacterial action exhibited against both Gram-negative and Gram-positive bacteria strains [88]. The significant porosity and average particle size of TiO2 NPs further promote bacterial mortality. Bulk materials lack the ability to connect with cellular components such as proteins, carbohydrates, nucleic acids, and fatty acids. However, small NPs can interact effectively due to their large specific surface area (SSA). Consequently, they exhibit higher toxicity and induce detrimental effects on bacterial pathogens. Moreover, due to its minuscule dimensions, it can penetrate cells and inflict damage upon them. Particle size can have an impact on the rate at which a substance dissolves. The little particle dissolves more rapidly due to its extensive surface area. The pharmaceutical industry utilizes this understanding when formulating new drugs, as substances with a larger surface area (smaller grain size) may undergo digestion at a faster rate. Due to the production of free radical oxides and peroxide, TiO2 NPs have potent antibacterial activity and broad reactivity against various microbial pathogens. Consequently, they have been shown to have a beneficial antimicrobial effect. The orange peel extract facilitates the eco-friendly production of TiO2 NPs. These NPs exhibit a superior inhibitory effect on the growth of P. aeruginosa compared to E. coli and S. aureus. Furthermore, they demonstrate an enhanced ability to combat cancer when tested against the A549 lung cancer cell line. The presence of TiO2 NPs in quantities ranging from 40 to 400 mg can induce cell membrane damage, oxidative stress, decreased levels of glutathione, and elevated levels of lipid peroxidation stress [89]. Figure 9 presents antimicrobial activity of Trigonella foenum-graecum TiO2 NPs against different microorganisms. Some examples of NPs synthesis, characterization, and application by using different plant extract are shown in Table 3.

12. Mechanism of Antimicrobial Activity

Although recent papers have postulated and explained various routes of cell death using reactive radicals, the precise mechanism by which NPs exhibit their antibacterial effects remains unknown and requires additional research. Generalizing the precise method by which NPs kill germs is challenging because to the variability in the types of NPs, strains, and environmental conditions employed in antibacterial experiments. Various NPs display varying antibacterial activity against different strains. Therefore, it is crucial to include these subjects in the debate forum to conduct a thorough investigation and clarification of the precise mechanism of action of NPs. Several difficulties, such as the entry of NPs into cells with varying structural integrity, need to be clarified. Furthermore, the precise mechanism by which NPs affect protein breakdown, DNA damage, and the modification of fundamental metabolic processes in cells remains largely unclear.
The main mechanisms by which NPs exhibit antibacterial activity, as suggested in the literature, are as follows: (a) the induction of oxidative stress through the production of ROS [93]—this oxidative process leads to lipid membrane peroxidation, resulting in damage to proteins and DNA within bacterial cells; (b) the release of metal ions from metal or metal oxide NPs, which can penetrate bacterial cell walls and directly interact with nucleic acids and proteins containing –SH, –NH, and –COOH groups, ultimately leading to cell death [94,95].
In 1985, ref. [96] published the first results on the antibacterial and photoelectrochemical properties of TiO2 powders loaded with platinum (TiO2/Pt). These powders were found to effectively kill Lactobacillus acidophilus, Saccharomyces cerevisiae, and Escherichia coli (E. coli) bacteria. Researchers have made significant efforts to enhance the photocatalytic bactericidal activity of TiO2 NPs. TiO2 nanostructures possess a broad range of uses in various industries, environmental processes, and energy-related applications. These include but are not limited to water purification, food preservation, the breakdown of dyes, chemical sensing, dye-sensitized solar cells, and the creation of antimicrobial agents. TiO2 doped with metals and non-metals that respond to visible light have been found to have bactericidal effects on various bacterial species, such as Gram-negative E. coli, Acinetobacter baumannii, and Shigella flexneri, as well as Gram-positive S. aureus, Bacillus subtilis, Listeria monocytogenes, and Bacillus anthracis spores [97]. These photocatalysts can sterilize dangerous germs, therefore effectively stopping the transmission of diseases caused by microbes. A general scheme showing the antibacterial action mechanism of TiO2 NPs is presented in Figure 10.

13. Green TiO2 as a Photocatalyst for the Degradation of Dyes

Green TiO2 has emerged as a promising photocatalyst for the degradation of dyes due to its enhanced efficiency and environmental friendliness. Its unique composition and structure absorb light efficiently, generating reactive oxygen species that break down dye molecules. Green TiO2 reduces harmful by-products, making it a sustainable textile wastewater treatment solution. The photocatalytic process of dye molecules by an anatase TiO2 catalyst generates highly reactive species like hydroxyl radicals (OH), holes (h+), superoxide (O2•−), and per hydroxy radicals. These species are formed in the solution when the catalyst absorbs radiation with a wavelength of 388 nm. These radicals react with chromophores and auxochromes to mineralize the dye molecule into carbon dioxide, water, and metal complexes [98]. The combined effects of surface area optimization, band structure modification, and crystallinity control, mediated by plant-derived biomolecules, give green-synthesized TiO2 nanoparticles (NPs) superior photocatalytic activity when compared to their chemically synthesized counterparts. Because of the chelation of the Ti4+ center, plant extracts high in polyphenols (–OH groups) or carboxylic acids (–COOH), like Cannabis sativa (bhang), promote anatase production, slowing hydrolysis and facilitating orderly crystallization. The metastable structure of anatase (band gap ~3.2 eV) improves photocatalytic activity and charge separation. High-temperature calcination (>500 °C) or extracts containing amines encourage rutile (band gap ~3.0 eV), which absorbs more visible light. In contrast to chemically synthesized anatase, jasmine flower extract produced rutile-phase TiO2 NPs with a 92% methylene blue degradation efficiency [61]. A high surface area-to-volume ratio and more reactive sites for dye adsorption and degradation (98.2% methylene blue elimination) were provided by the 12.5 ± 1.5 nm particle size of the TiO2 NPs made from Cannabis sativa [99]. In microwave-assisted synthesis, L-ascorbic acid produced oxygen vacancies (Vo) and Ti3+ defects, which reduced the band gap and extended light absorption into the visible and near-infrared regions. Under visible light, this faulty TiO2 produced 162 μmol g−1 h−1 of hydrogen [100]. Moreover, the decomposition of –OH, –COOH surface groups, or adsorbed bulky biomolecules upon illumination may be the source of decomposition products that either enhance photoactivity by highly active radicals or hinder photoactivity by free radical scavenging properties in further photocatalytic reactions. Another factor enhancing the degradation on green TiO2 may be better adsorption of dyes on the surface of photocatalyst. Finally, the adsorbed molecules may behave as photosensibilization agents, improving the photocatalytic process.
A flower extract produced TiO2 NPs with the highest absorbance and lowest band gap of 2.89 eV, according to Rathi [52]. Floral extract-generated TiO2 NPs likewise exhibited a maximum degradation of 86% against MB dye. Hiremath et al. [101] produced Taramindus TiO2 NPs with a band gap of 3.2 eV and a high surface area of 75 m2/g. The anatase material underwent testing to assess its ability to degrade Titan Yellow (TY) dye, with a focus on examining different variables. A maximum degradation of 95% was achieved under neutral conditions (pH = 7) with a dye concentration of 10 ppm and 75 mg of TiO2 NPs in less than 2 h. Pavithra et al. [102] synthesized TiO2 NPs using Colotropis gigantea for the purpose of degrading MB dye. A reduction of 96% was observed after one hour of utilizing Acorus calamus extract to create spherical and interconnected clusters of TiO2 with a particle size of 39 nm, achieving 96% degradation of RhB from its initial concentration [103]. The degradation of Congo Red (CR) dye and MB was also examined using Myristica fragrans TiO2 NPs. The anatase structured material, which exhibited a blue shift, achieved a maximum degradation of 99% and 97% against CR and MB, respectively. The optimal efficiency was achieved due to the formation of O2•− and holes. The authors of [104] investigated the impact of pH, time, dosage, and concentration on the degradation of MB utilizing Monsonia burkeana TiO2 NPs. They also examined the potential mechanism of degradation. The maximum degradation of 85% for MB occurred at a pH of 10 after 120 min, using a dosage of 60 mg TiO2 NPs and a concentration of MB of 20 ppm. However, when testing the material’s capacity to be utilized again, it was found that the TiO2 NPs only lasted for one cycle. This suggests that either the catalyst was lost, or the active sites were blocked, leading to a significant decrease in the material’s ability to degrade.

14. Mechanism of Photocatalysis

Instead of changing the photocatalyst, light energy is converted into chemical energy to synthesize or break down molecules. When light is applied, electron–hole pairs are created, which starts photocatalytic activity. Holes in the VB will appear when a semiconductor photocatalyst absorbs photons equal to or higher than its Eg while the electrons in the VB will be stimulated to the CB. On the other hand, photogenerated holes and electrons are captured by defect sites within the catalyst’s surface and/or in bulk, where they recombine to produce heat or photons [105]. The rate of photogenerated carrier recombination determines the carrier lifespan; rapid recombination results in short lifetimes, which in turn restricts the catalyst’s ability to convert photochemical energy into a useful form [106].
The transfer of electronic charges to the semiconductor catalyst’s outer surface helps to promote interfacial substrates and chemical reactions. When photogenerated electrons in CB interact with the dissolved oxygen at the catalytic surface, they produce superoxide radicals (O2•−), which then react with the holes on the VB to produce singlet oxygen (1O2). Photogenerated holes in the VB are effective oxidizers (+1.0 to +3.5 V vs. NHE) and interact with surface moisture (H2O or OH) to create H2O2 and hydroxyl radicals. Reaction between hydroxyl radicals (HO) and hydrogen peroxide (H2O2) produces protonated superoxide radicals (HOO) [107,108].
Ijms 26 05454 i018
Electrons and holes may recombine within the semiconductor rather than transfer to the catalytic outer surface. One disadvantage of this light-driven method is that recombination of photogenerated electrons and holes might lose most of the energy as heat. In this regard, a greater gap between electrons and holes is required to produce a quick shift in photocatalytic activity and redox chemical processes in the semiconductor.
Adjusting the intrinsic band structure can help with pollutant adsorption efficiency, photocatalytic efficiency, redox potential, and charge carrier recombination/separation. The key to understanding the photocatalytic process and increasing photocatalytic activity is having a solid understanding of electronic transition [109]. Figure 11 illustrates the fundamental principles of semiconductor photocatalysis.

15. Conclusions

This publication presents a review of methods for the synthesis of titanium dioxide nanoparticles using various plant extracts. Numerous possibilities for the synthesis of this nanomaterial are presented, paying special attention to the limitations. Physicochemical properties of nanomaterials obtained by green methods are characterized in terms of their morphology, optical properties, crystallographic structure, and the presence of functional groups. Finally, directions for the application of TiO2 nanomaterials are presented, especially as antimicrobial nanomaterials and as photocatalysts. The numerous significant benefits that are associated with these methods are the primary reason for the tremendous support that is being shown for green synthesis. In a wide variety of applications, green synthesis helps to reduce the negative effects that the generated nanoparticles have. In addition to this, it is safe, economically viable, efficient, and environmentally friendly, and it can be easily improved. Due to the reduced utilization of precursors during the green synthesis process, the green technique exhibits a higher level of efficiency in comparison with chemically synthesized nanoparticles techniques. For example, Psidium guajava-derived TiO2 NPs demonstrated the maximum zone of inhibition at 25 mm against Staphylococcus aureus. TiO2 NPs obtained with the use of Calotropis gigantea showed a reduction of 96% of MB dye after one hour. Because fewer plants are currently being utilized for the environmentally friendly synthesis of TiO2 NPs, there is still a great deal of work to be done in this field. On the other hand, these nanoparticles from green synthesis can be used in biomedical applications with low risks, and they are just as compatible with chemically synthesized nanoparticles in other potential applications. TiO2 NPs have the potential to be an effective treatment for polluted wastewater as well as soils that have been contaminated with metals. However, the efficacy of the treatment is contingent upon several factors, including the type of metal, the duration of exposure, the dosage, and the form of the nanoparticles. Green synthesis can decrease material expenses by roughly 30–50%. The removal of dangerous substances such as titanium tetrachloride might result in a reduction in raw material expenses by as much as 40%. TiO2 NPs synthesized via green methods might exhibit up to 25% more photocatalytic efficiency and 30% enhanced antibacterial activity relative to those produced through chemical synthesis. It is necessary to conduct additional research to improve one’s understanding of the production and utilization of green TiO2 NPs.

Author Contributions

N.S. and I.P. helped in conceptualization, methodology, investigation, writing—original draft, writing—review and editing, and supervision. P.I. and A.K. contributed to methodology, investigation, and writing—original draft. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data relevant to this work is presented in the manuscript.

Conflicts of Interest

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

References

  1. Xue, J.; Liu, J.; Liu, Y.; Li, H.; Wang, Y.; Sun, D.; Wang, W.; Huang, L.; Tang, J. Recent advances in synthetic methods and applications of Ag2S-based heterostructure photocatalysts. J. Mater. Chem. C 2019, 7, 3988–4003. [Google Scholar] [CrossRef]
  2. Priyadharsini, P.; Sundar, R.P.; Grace, P.K.; Naveen, S.; Sanjay, K.S.; Gnanaprakash, D.; Arun, J.; Pugazhendhi, A. Nanohybrid photocatalysts in dye (Colorants) wastewater treatment: Recent trends in simultaneous dye degradation, hydrogen production, storage and transport feasibility. J. Clean. Prod. 2023, 426, 139180. [Google Scholar] [CrossRef]
  3. Bak, T.; Nowotny, J.; Rekas, M.; Sorrell, C.C. Photo-electrochemical hydrogen generation from water using solar energy. Materials-related aspects. Int. J. Hydrogen Energy 2002, 27, 991–1022. [Google Scholar] [CrossRef]
  4. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  5. Nagaveni, K.; Sivalingam, G.; Hedge, M.S.; Madras, G. Solar photocatalytic degradation of dyes: High activity of combustion synthesized nano TiO2. Appl. Catal. B 2004, 48, 83–93. [Google Scholar] [CrossRef]
  6. Cong, Y.; Zhang, J.; Chen, F.; Anpo, M.; He, D. Preparation, photocatalytic activity, and mechanism of nano-TiO2 co-doped with nitrogen and iron (III). J. Phys. Chem. C 2007, 111, 10618–10623. [Google Scholar] [CrossRef]
  7. Macwan, D.P.; Dave, P.N.; Chaturvedi, S. A review on nano-TiO2 sol–gel type syntheses and its applications. J. Mater. Sci. 2011, 46, 3669–3686. [Google Scholar] [CrossRef]
  8. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z. Semiconductor heterojunction photocatalysts: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234–5244. [Google Scholar] [CrossRef]
  9. Kumar, S.G.; Devi, L.G. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. J. Phys. Chem. 2011, 115, 13211–13241. [Google Scholar] [CrossRef]
  10. Chawla, H.; Chandra, A.; Ingole, P.P.; Garg, S. Recent advancements in enhancement of photocatalytic activity using bismuth-based metal oxides Bi2MO6 (M = W, Mo, Cr) for environmental remediation and clean energy production. J. Ind. Eng. Chem. 2021, 95, 1–15. [Google Scholar] [CrossRef]
  11. Singhal, N.; Kumar, U. Noble metal modified TiO2: Selective photoreduction of CO2 to hydrocarbons. Mol. Catal. 2017, 439, 91–99. [Google Scholar] [CrossRef]
  12. Ohno, T.; Akiyoshi, M.; Umebayashi, T.; Asai, K.; Mitsui, T.; Matsumura, M. Preparation of S-doped TiO2 photocatalysts and their photocatalytic activities under visible light. Water Sci. Technol. 2004, 265, 115–121. [Google Scholar] [CrossRef]
  13. Li, D.; Haneda, H.; Hishita, S.; Ohashi, N. Visible-light-driven N−F−codoped TiO2 photocatalysts. 1. Synthesis by spray pyrolysis and surface characterization. Chem. Mater. 2005, 17, 2588–2595. [Google Scholar] [CrossRef]
  14. Wang, C.Y.; Bttcher, C.; Bahnemann, D.W.; Dohrmann, J.K. A comparative study of nanometer sized Fe(III)-doped TiO2 photocatalysts: Synthesis, characterization and activity. J. Mater. Chem. A 2003, 5, 11143. [Google Scholar]
  15. Zhou, Q.; Fang, Z.; Li, J.; Wang, M. Applications of TiO2 nanotube arrays in environmental and energy fields: A review. Microporous Mater. 2015, 202, 22–35. [Google Scholar] [CrossRef]
  16. Wang, Y.; He, Y.; Lai, Q.; Fan, M. Review of the progress in preparing nano TiO2: An important environmental engineering material. J. Environ. Sci. 2014, 26, 2139–2177. [Google Scholar] [CrossRef]
  17. Park, H.; Park, Y.; Kim, W.; Choi, W. Surface modification of TiO2 photocatalyst for environmental applications. J. Photochem. Photobiol. C 2013, 15, 1–20. [Google Scholar] [CrossRef]
  18. Phattalung, S.; Harding, D.J.; Pattanasattayavong, P.; Kim, H.; Lee, J.; Hwang, D.W.; Chung, T.D.; Yu, J. Band gap narrowing of TiO2 nanoparticles: A passivated Co-doping approach for enhanced photocatalytic activity. J. Phys. Chem. Solids. 2022, 162, 110503. [Google Scholar] [CrossRef]
  19. Ye, L.; Liang, Y. First principles study on band gap modulation of TiO2 (112) surface for enhancing optical properties. Phys. B 2024, 674, 415579. [Google Scholar] [CrossRef]
  20. Ghareeb, A.; Fouda, A.; Kishk, R.M.; El Kazzaz, W.M. Unlocking the potential of titanium dioxide nanoparticles: An insight into green synthesis, optimizations, characterizations, and multifunctional applications. MCFs 2024, 23, 341. [Google Scholar] [CrossRef]
  21. Gohari, G.; Mohammadi, A.; Akbari, A.; Panahirad, S.; Dadpour, M.R.; Fotopoulos, V.; Kimura, S. Titanium dioxide nanoparticles (TiO2 NPs) promote growth and ameliorate salinity stress effects on essential oil profile and biochemical attributes of Dracocephalum moldavica. Sci. Rep. 2020, 10, 912. [Google Scholar] [CrossRef] [PubMed]
  22. Liza, T.Z.; Tusher, M.M.; Anwar, F.; Monika, M.F.; Amin, K.F.; Asrafuzzaman, F.N. Effect of Ag-doping on morphology, structure, band gap and photocatalytic activity of bio-mediated TiO2 nanoparticles. Res. Mater. 2024, 22, 100559. [Google Scholar] [CrossRef]
  23. Lin, C.J.; Kao, L.-C.; Huang, Y.; Bañares, M.A.; Liou, S.Y.-H. Uniform deposition of coupled CdS and CdSe quantum dots on ZnO nanorod arrays as electrodes for photoelectrochemical solar water splitting. Int. J. Hydrogen Energy 2015, 40, 1388–1393. [Google Scholar] [CrossRef]
  24. Iravani, S. Green synthesis of metal nanoparticles using plants. Green Chem. 2011, 13, 2638–2650. [Google Scholar] [CrossRef]
  25. Pravallika, P.L.; Mohan, G.K.; Rao, K.V.; Shanker, K. Biosynthesis, characterization and acute oral toxicity studies of synthesized iron oxide nanoparticles using ethanolic extract of Centella asiatica plant. Mater. Lett. 2019, 236, 256–259. [Google Scholar] [CrossRef]
  26. Tran, Q.H.; Le, A.T. Silver nanoparticles: Synthesis, properties, toxicology, applications and perspectives. Adv. Nat. Sci. Nanosci. Nanotechnol. 2013, 14, 033001. [Google Scholar] [CrossRef]
  27. 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, 156, 111221. [Google Scholar]
  28. Yitagesu, G.B.; Leku, D.T.; Workneh, G.A. Green synthesis of TiO2 using Impatiens rothii hook. f. leaf extract for efficient removal of methylene blue dye. ACS Omega 2023, 46, 43999–44012. [Google Scholar] [CrossRef]
  29. Verma, V.; Al-Dossari, M.; Singh, J.; Rawat, M.; Kordy, M.G.; Shaban, M. A review on green synthesis of TiO2 NPs: Photocatalysis and antimicrobial applications. Polymers 2022, 7, 1444. [Google Scholar] [CrossRef]
  30. Ansari, F.S.; Daneshjou, S. Optimizing the green synthesis of antibacterial TiO2-anatase phase nanoparticles derived from spinach leaf extract. Sci. Rep. 2024, 14, 22440. [Google Scholar] [CrossRef]
  31. Zhang, X.; Huo, K.; Wang, H.; Zhang, W.; Chu, P.K. Influence of structure parameters and crystalline phase on the photocatalytic activity of TiO2 nanotube arrays. J. Nanosci. Nanotechnol. 2011, 12, 11200–11205. [Google Scholar] [CrossRef] [PubMed]
  32. Balaji, S.; Pandian, M.S.; Ramasamy, G.; Karchiyappan, T. Green synthesis of metal oxide nanoparticles using plant extracts: A sustainable approach to combat antimicrobial resistance. Environ. Nanotechnol. Monit. Manag. 2025, 23, 101066. [Google Scholar] [CrossRef]
  33. Ahmad, M.; Khan, S.U.; Ahmad, A.; Ahmad, I. Green synthesis of TiO2 nanoparticles using Rhizophora apiculata leaves extract and their photocatalytic activity. Appl. Surf. Sci. 2016, 392, 1229–1235. [Google Scholar]
  34. Anand, S.; Sahu, M.; Rath, R.N. Green synthesis of titanium dioxide nanoparticles using Ocimum sanctum extract and their catalytic and antibacterial applications. J. Environ. Chem. Eng. 2019, 3, 103284. [Google Scholar]
  35. Bhattacharyya, T.; Mondal, S.; Jana, N.R. Green synthesis of titanium dioxide nanoparticles using seed extract of Butea monosperma and their photocatalytic activity. J. Photochem. Photobiol. B Biol. 2018, 186, 181–187. [Google Scholar]
  36. Nasrollahzadeh, M.; Sajjadi, M.; Iravani, S.; Varma, R.S. Plant-mediated green synthesis of nanostructures: Mechanisms, characterization, and applications. Chem. Rev. 2021, 12, 7468–7533. [Google Scholar]
  37. Hua, D.; Gao, S.; Zhang, M.; Ma, W.; Huang, C. A novel Xanthan gum-based conductive hydrogel with excellent mechanical, biocompatible, and self-healing performances. Carbohydr. Polym. 2020, 247, 116743. [Google Scholar] [CrossRef]
  38. Munir, H.; Yaqoob, S.; Awan, K.A.; Imtiaz, A.; Naveed, H.; Ahmad, N.; Naeem, M.; Sultan, W.; Ma, Y. Unveiling the chemistry of citrus peel: Insights into nutraceutical potential and therapeutic applications. Foods 2024, 13, 1681. [Google Scholar] [CrossRef] [PubMed]
  39. Perron, N.R.; Brumaghim, J.L. A Review of the Antioxidant Mechanisms of Polyphenol Compounds Related to Iron Binding. Cell Biochem. Biophys. 2009, 53, 75–100. [Google Scholar] [CrossRef]
  40. Kuppusamy, P.; Yusoff, M.M.; Maniam, G.P.; Govindan, N. Biosynthesis of metallic nanoparticles using plant derivatives and their new avenues in pharmacological applications—An updated report. Saudi Pharm. J. 2016, 24, 473–484. [Google Scholar] [CrossRef]
  41. Chen, X.; Mao, S.S. Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  42. Diebold, U. The surface science of titanium dioxide. Surf. Sci. Rep. 2003, 48, 53–229. [Google Scholar] [CrossRef]
  43. Chen, D.; Burda, C. The electronic origin of the visible-light absorption properties of C-, N- and S-doped TiO2 nanomaterials. J. Am. Chem. Soc. 2008, 130, 5018–5019. [Google Scholar] [CrossRef] [PubMed]
  44. Fujishima, A.; Zhang, X. Titanium dioxide photocatalysis: Present situation and future approaches. Comptes Rendus Chim. 2006, 9, 750–760. [Google Scholar] [CrossRef]
  45. Lu, H.; Hou, L.; Zhang, Y.; Cao, X.; Xu, X.; Shang, Y. Pilot-scale and large-scale Fenton-like applications with nano-metal catalysts: From catalytic modules to scale-up applications. Water Res. 2024, 266, 122425. [Google Scholar] [CrossRef]
  46. Rajaram, P.; Jeice, A.R.; Jayakumar, K. Review of green synthesized TiO2 nanoparticles for diverse applications. Surf. Interfaces 2023, 39, 102912. [Google Scholar] [CrossRef]
  47. Gupta, D.; Boora, A.; Thakur, A.; Gupta, T.K. Green and sustainable synthesis of nanomaterials: Recent advancements and limitations. Environ. Res. 2023, 231, 116316. [Google Scholar] [CrossRef]
  48. Szymanski, N.J.; Rendy, B.; Fei, Y.; Kumar, R.E.; He, T.; Milsted, D.; McDermott, M.J.; Gallant, M.; Cubuk, E.D.; Merchant, A.; et al. An autonomous laboratory for the accelerated synthesis of novel materials. Nature 2023, 624, 86–91. [Google Scholar] [CrossRef]
  49. Miller, A.; Thompson, L. Energy consumption in TiO2 synthesis: A comparative study of various methods. J. Clean. Prod. 2017, 165, 1273–1280. [Google Scholar]
  50. Garcia, M.; Fernandez, R. Green synthesis approaches for TiO2 nanoparticles: Challenges and prospects. ACS Sustain. Chem. Eng. 2021, 35, 45–60. [Google Scholar]
  51. Gatou, M.A.; Syrrakou, A.; Lagopati, N.; Pavlatou, E.A. Photocatalytic TiO2-based nanostructures as a promising material for diverse environmental applications: A review. Reactions 2024, 1, 135–194. [Google Scholar] [CrossRef]
  52. Rathi, V.H.; Jeice, A.R. Green fabrication of titanium dioxide nanoparticles and their applications in photocatalytic dye degradation and microbial activities. Chem. Phys. 2023, 6, 100197. [Google Scholar] [CrossRef]
  53. Alberti, S.; Villa, S.; Singh, G.; Seland, F.; Martinelli, A.; Ferretti, M.; Canepa, F.; Caratto, V. Systematic Study on TiO2 Crystallization via Hydrothermal Synthesis in the Presence of Different Ferrite Nanoparticles as Nucleation Seeds. J. Nanosci. Nanotechnol. 2019, 8, 4994–4999. [Google Scholar] [CrossRef] [PubMed]
  54. Singh, J.; Kumar, S.; Alok, A.; Upadhyay, S.K.; Rawat, M.; Tsang, D.C. The potential of green synthesized zinc oxide nanoparticles as nutrient source for plant growth. J. Clean. Prod. 2019, 214, 1061–1070. [Google Scholar] [CrossRef]
  55. Nadeem, M.; Tungmunnithum, D.; Hano, C.; Abbasi, B.H.; Hashmi, S.S.; Ahmad, W. The current trends in the green syntheses of titanium oxide nanoparticles and their applications. Green Chem. Lett. Rev. 2018, 11, 492–502. [Google Scholar] [CrossRef]
  56. Subhapriya, S.; Gomathipriya, P. Green synthesis of titanium dioxide (TiO2) nanoparticles by Trigonella foenum-graecum extract and its antimicrobial properties. Microb. Pathog. 2018, 116, 215–220. [Google Scholar] [CrossRef]
  57. Bopape, D.A.; Tetana, Z.N.; Mabuba, N.; Motaung, D.E.; Hintsho-Mbita, N.C. Biosynthesis of TiO2 nanoparticles using Commelina benghanlensis for the photodegradation of methylene blue dye and antibiotics: Effect of plant concentration. Res. Chem. 2023, 5, 100825. [Google Scholar] [CrossRef]
  58. Mbenga, Y.; Adeyemi, J.; Mthiyane, D.M.N.; Singh, M.; Onwudiwe, D.C. Green synthesis, antioxidant and anticancer activities of TiO2 nanoparticles using aqueous extract of Tulbhagia violacea. Results. Chem. 2023, 6, 101007. [Google Scholar] [CrossRef]
  59. Shimi, A.K.; Wabaidur, S.M.; Siddiqui, M.R.; Islam, M.A.; Rane, K.P.; Jeevan, T.S. Photocatalytic Activity of Green Construction TiO2 Nanoparticles from Phyllanthus niruri Leaf Extract. J. Nanomater. 2022, 2022, 7011539. [Google Scholar] [CrossRef]
  60. Kaur, H.; Kaur, S.; Kumar, S. Eco-friendly approach: Synthesis of novel green TiO2 nanoparticles for degradation of reactive green 19 dye and replacement of chemical synthesized TiO2. J. Clust. Sci. 2021, 32, 1191–1204. [Google Scholar] [CrossRef]
  61. Aravind, M.; Amalanathan, M.; Mary, M. Synthesis of TiO2 nanoparticles by chemical and green synthesis methods and their multifaceted properties. SN Appl. Sci. 2021, 3, 409. [Google Scholar] [CrossRef]
  62. Sethy, N.K.; Arif, Z.; Mishra, P.K.; 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, 9, 171–181. [Google Scholar] [CrossRef]
  63. Arabi, N.; Kianvas, A.; Hajalilou, A.; Abouzarati-Lotf, E.; Abbasi-Chianeh, V.A. A facile and green synthetic approach toward fabrication of Alcea-and Thyme-stabilized TiO2 nanoparticles for photocatalytic applications. Arab. J. Chem. 2020, 13, 2132–2141. [Google Scholar] [CrossRef]
  64. Thakur, B.K.; Kumar, D.; Kumar, A.S. Green synthesis of titanium dioxide nanoparticles using Azadirachta indica leaf extract and evaluation of their antibacterial activity. Afr. J. Bot. 2019, 124, 223–227. [Google Scholar] [CrossRef]
  65. Abu-Dalo, M.; Jaradat, A.; Albiss, B.A.; Al-Rawashdeh, N.A. Green synthesis of TiO2 NPs/pristine pomegranate peel extract nanocomposite and its antimicrobial activity for water disinfection. J. Environ. Chem. Eng. 2019, 7, 103370. [Google Scholar] [CrossRef]
  66. Saif, S.; Tahir, A.; Chen, Y. Green synthesis of iron nanoparticles and their environmental applications and implications. Nanomaterials 2016, 11, 209. [Google Scholar] [CrossRef] [PubMed]
  67. Nasrollahzadeh, M.; Sajadi, S.M. Synthesis and characterization of titanium dioxide nanoparticles using Euphorbia heteradena Jaub root extract and evaluation of their stability. Ceram. Int. 2015, 10, 14435–14439. [Google Scholar] [CrossRef]
  68. Rajkumari, J.; Magdalane, C.M.; Siddhardha, B.; Madhavan, J.; Ramalingam, G.; Al-Dhabi, N.A.; Arasu, M.V.; Ghilan, A.K.; Duraipandiayan, V.; Kaviyarasu, K. Synthesis of titanium oxide nanoparticles using Aloe barbadensis mill and evaluation of its antibiofilm potential against Pseudomonas aeruginosa PAO1. J. Photochem. Photobiol. B 2019, 201, 111667. [Google Scholar] [CrossRef]
  69. Goutam, S.P.; Saxena, G.; Singh, V.; Yadav, A.K.; Bharagava, R.N.; Thapa, K.B. Green synthesis of TiO2 nanoparticles using leaf extract of Jatropha curcas L. for photocatalytic degradation of tannery wastewater. J. Chem. Eng. 2018, 336, 386–396. [Google Scholar] [CrossRef]
  70. Saratale, R.G.; Saratale, G.D.; Shin, H.S.; Jacob, J.M.; Pugazhendhi, A.; Bhaisare, M.; Kumar, G. New insights on the green synthesis of metallic nanoparticles using plant and waste biomaterials: Current knowledge, their agricultural and environmental applications. Environ. Sci. Pollut. Res. 2018, 25, 10164–10183. [Google Scholar] [CrossRef]
  71. Bhattarai, B.; Zaker, Y.; Bigioni, T.P. Green synthesis of gold and silver nanoparticles: Challenges and opportunities. Curr. Opin. Green Sustain Chem. 2018, 12, 91–100. [Google Scholar] [CrossRef]
  72. Guan, Z.; Ying, S.; Ofoegbu, P.C.; Clubb, P.; Rico, C.; He, F.; Hong, J. Green synthesis of nanoparticles: Current developments and limitations. Environ. Technol. Innov. 2022, 26, 102336. [Google Scholar]
  73. Parveen, K.; Banse, V.; Ledwani, L. Green synthesis of nanoparticles: Their advantages and disadvantages. AIP Conf. Proc. 2016, 1, 020048. [Google Scholar]
  74. Woźniak, A.; Malankowska, A.; Nowaczyk, G.; Grześkowiak, B.F.; Tuśnio, K.; Słomski, R.; Zaleska-Medynska, A.; Jurga, S. Size and shape-dependent cytotoxicity profile of gold nanoparticles for biomedical applications. J. Mater. Sci. Mater. Med. 2017, 28, 92. [Google Scholar] [CrossRef]
  75. Baudot, C.; Tan, C.M.; Kong, J.C. FTIR spectroscopy as a tool for nano-material characterization. Infrared Phys. Technol. 2010, 6, 434–438. [Google Scholar] [CrossRef]
  76. Irshad, M.A.; Nawaz, R.; Rehman, M.Z.; Imran, M.; Ahmad, M.J.; Ahmad, S.; Ali, S. Synthesis and characterization of titanium dioxide nanoparticles by chemical and green methods and their antifungal activities against wheat rust. Chemosphere 2020, 258, 127352. [Google Scholar] [CrossRef]
  77. Mikkelsen, R. Nanofertilizer and nanotechnology: A quick look. Better Crops 2018, 3, 18–19. [Google Scholar] [CrossRef]
  78. Isah, T. Stress and defense responses in plant secondary metabolites production. Biol. Res. 2019, 52, 39. [Google Scholar] [CrossRef] [PubMed]
  79. Usman, M.; Farooq, M.; Wakeel, A.; Nawaz, A.; Cheema, S.A.; Rehman, H.; Ashraf, I.; Sanaullah, M. Nanotechnology in agriculture: Current status, challenges and future opportunities. Sci. Total Environ. 2020, 15, 137721–137778. [Google Scholar] [CrossRef]
  80. Chandoliya, R.; Sharma, S.; Sharma, V.; Joshi, R.; Sivanesan, I. Titanium Dioxide Nanoparticle: A Comprehensive Review on Synthesis, Applications and Toxicity. Plants 2024, 21, 2964. [Google Scholar] [CrossRef]
  81. Campbell, J.; Burkitt, S.; Dong, N.; Zavaleta, C. Nanoparticle characterization techniques. In Nanoparticles for Biomedical Applications; Elsevier: Amsterdam, The Netherlands, 2020; Volume 1, pp. 129–144. [Google Scholar]
  82. Shimpi, N.G.; Mishra, S.; Shirole, S.D. Efficient green synthesis of TiO2 nanoparticles using Murraya koenigii leaf extract. J. Nanosci. Nanoeng. Appl. 2016, 1, 29–38. [Google Scholar]
  83. Solano, R.A.; Herrera, A.P.; Maestre, D.; Cremades, A. Fe-TiO2 nanoparticles synthesized by green chemistry for potential application in wastewater photocatalytic treatment. J. Nanotechnol. 2019, 2019, 4571848. [Google Scholar] [CrossRef]
  84. 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, 9, 095034. [Google Scholar] [CrossRef]
  85. Esfahani, R.N.; Khaghani, S.; Azizi, A.; Mortazaeinezhad, F.; Gomarian, M. Facile and eco-friendly synthesis of TiO2 NPs using extracts of Verbascum thapsus plant: An efficient photocatalyst for reduction of Cr (VI) ions in the aqueous solution. J. Iran. Chem. Soc. 2020, 17, 205–213. [Google Scholar] [CrossRef]
  86. Santhoshkumar, T.; Rahuman, A.A.; Jayaseelan, C.; Rajakumar, G.; Marimuthu, S.; Kirthi, A.V.; Velayutham, K.; Thomas, J.; Venkatesan, J.; Kim, S.K. Green synthesis of titanium dioxide nanoparticles using Psidium guajava extract and its antibacterial and antioxidant properties. Asian Pac. J. Trop. Med. 2014, 12, 968–976. [Google Scholar] [CrossRef]
  87. Ahmad, W.; Jaiswal, K.K.; Soni, S. Green synthesis of titanium dioxide (TiO2) nanoparticles by using Mentha arvensis leaves extract and its antimicrobial properties. Inorg. Nano-Met. Chem. 2020, 50, 1032–1038. [Google Scholar] [CrossRef]
  88. Farag, S.; Amr, A.; El-Shafei, A.; Asker, M.S.; Ibrahim, H.M. Green synthesis of titanium dioxide nanoparticles via bacterial cellulose (BC) produced from agricultural wastes. Cellulose 2021, 28, 7619–7632. [Google Scholar] [CrossRef]
  89. Amanulla, A.M.; Sundaram, R.J. Green synthesis of TiO2 nanoparticles using orange peel extract for antibacterial, cytotoxicity and humidity sensor applications. Mater. Today Proc. 2019, 8, 323–331. [Google Scholar] [CrossRef]
  90. Bekele, E.T.; Gonfa, B.A.; Zelekew, O.A.; Belay, H.H.; Sabir, F.K. Synthesis of titanium oxide nanoparticles using root extract of Kniphofia foliosa as a template, characterization, and its application on drug resistance bacteria. J. Nanomater. 2020, 2020, 2817037. [Google Scholar] [CrossRef]
  91. Hariharan, D.; Thangamuniyandi, P.; Selvakumar, P.; Devan, U.; Pugazhendhi, A.; Vasantharaja, R.; Nehru, L.C. Green approach synthesis of Pd@TiO2 nanoparticles: Characterization, visible light active picric acid degradation and anticancer activity. Process Biochem. 2019, 87, 83–88. [Google Scholar] [CrossRef]
  92. Swathi, N.; Sandhiya, D.; Rajeshkumar, S.; Lakshmi, T. Green synthesis of titanium dioxide nanoparticles using Cassia fistula and its antibacterial activity. Int. J. Res. Pharm. Sci. 2019, 10, 856–860. [Google Scholar] [CrossRef]
  93. Collin, F. Chemical basis of reactive oxygen species reactivity and involvement in neurodegenerative diseases. Int. J. Mol. Sci. 2019, 20, 2407. [Google Scholar] [CrossRef] [PubMed]
  94. Liao, C.; Li, Y.; Tjong, S.C. Bactericidal and cytotoxic properties of silver nanoparticles. Int. J. Mol. Sci. 2019, 20, 449. [Google Scholar] [CrossRef] [PubMed]
  95. Regmi, C.; Joshi, B.; Ray, S.K.; Gyawali, G.; Pandey, R.P. Understanding the mechanism of photocatalytic microbial decontamination of environmental wastewater. Front. Chem. 2018, 6, 33. [Google Scholar] [CrossRef] [PubMed]
  96. Matsunaga, T.; Tomoda, R.; Nakajima, T.; Wake, H. Photoelectrochemical sterilization of microbial cells by semiconductor powders. FEMS Microbiol. Lett. 1985, 29, 211–214. [Google Scholar] [CrossRef]
  97. Liou, J.W.; Chang, H.H. Bactericidal effects and mechanisms of visible light-responsive titanium dioxide photocatalysts on pathogenic bacteria. Arch. Immunol. Ther. Exp. 2012, 60, 267–275. [Google Scholar] [CrossRef]
  98. Ahmad, R.; Srivastava, S.; Ghosh, S.; Khare, S.K. Phytochemical delivery through nanocarriers: A review. Colloids Surf. B Biointerfaces 2021, 197, 111389. [Google Scholar] [CrossRef]
  99. BinSabt, M.; Sagar, V.; Singh, J.; Rawat, M.; Shaban, M. Green synthesis of CS-TiO2 NPs for efficient photocatalytic degradation of methylene blue dye. Polymers 2022, 13, 2677. [Google Scholar] [CrossRef]
  100. Kunthakudee, N.; Puangpetch, T.; Ramakul, P.; Serivalsatit, K.; Ponchio, C.; Hunsom, M. Ultra-fast green synthesis of a defective TiO2 photocatalyst towards hydrogen production. RSC Adv. 2024, 14, 24213–24225. [Google Scholar] [CrossRef]
  101. Hiremath, S.; MAL, A.R.; Prabha, C.; Vidya, C. Tamarindus indica mediated biosynthesis of nano TiO2 and its application in photocatalytic degradation of Titan yellow. J. Environ. Chem. Eng. 2018, 6, 7338–7346. [Google Scholar] [CrossRef]
  102. 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. [Google Scholar] [CrossRef]
  103. Ansari, A.; Siddiqui, V.U.; Rehman, W.U.; Akram, M.K.; Siddiqi, W.A.; Alosaimi, A.M.; Hussein, M.A.; Rafatullah, M. Green synthesis of TiO2 nanoparticles using Acorus calamus leaf extract and evaluating its photocatalytic and in vitro antimicrobial activity. Catalysts 2022, 12, 181. [Google Scholar] [CrossRef]
  104. Ngoepe, N.M.; Mathipa, M.M.; Hintsho-Mbita, N.C. Biosynthesis of titanium dioxide nanoparticles for the photodegradation of dyes and removal of bacteria. Optik 2020, 224, 65728. [Google Scholar] [CrossRef]
  105. Carp, O.; Huisman, C.L.; Reller, A. Photoinduced reactivity of titanium dioxide. Prog. Solid State Chem. 2004, 32, 33–177. [Google Scholar] [CrossRef]
  106. Yang, H.; Beibei, Y.; Wei, C.; Junjiao, Y. Preparation and photocatalytic activities of TiO2-based composite catalysts. Catalysts 2022, 10, 1263. [Google Scholar] [CrossRef]
  107. Srikanth, B.; Goutham, R.; Badri Narayan, R.; Ramprasath, A.; Gopinath, K.P.; Sankaranarayanan, A.R. Recent advancements in supporting materials for immobilised photocatalytic applications in wastewater treatment. J. Environ. Manag. 2017, 200, 60–78. [Google Scholar] [CrossRef]
  108. Guo, Q.; Zhou, C.; Ma, Z.; Yang, X. Fundamentals of TiO2 photocatalysis: Concepts, mechanisms, and challenges. Adv. Mater. 2019, 31, 1901997. [Google Scholar] [CrossRef]
  109. Arora, I.; Chawla, H.; Chandra, A.; Sagadevan, S.; Garg, S. Advances in the strategies for enhancing the photocatalytic activity of TiO2: Conversion from UV-light active to visible-light active photocatalyst. Inorg. Chem. Commun. 2022, 1, 143. [Google Scholar] [CrossRef]
Figure 1. Potential applications of TiO2 nanoparticles in various fields of life.
Figure 1. Potential applications of TiO2 nanoparticles in various fields of life.
Ijms 26 05454 g001
Figure 2. Synthesis approaches, such as chemical, physical, and green methods.
Figure 2. Synthesis approaches, such as chemical, physical, and green methods.
Ijms 26 05454 g002
Figure 3. Illustration of TiO2 green synthesis and its applications.
Figure 3. Illustration of TiO2 green synthesis and its applications.
Ijms 26 05454 g003
Figure 5. FT-IR absorption spectra of chemically synthesized TiO2-NPs (A)—(T1), (B)—(T2), and (C)—(T3) are synthesized by the plant extracts of T. portulacastrum and C. quinoa, respectively. Reprinted with permission from ref. [76] with permission of Elsevier.
Figure 5. FT-IR absorption spectra of chemically synthesized TiO2-NPs (A)—(T1), (B)—(T2), and (C)—(T3) are synthesized by the plant extracts of T. portulacastrum and C. quinoa, respectively. Reprinted with permission from ref. [76] with permission of Elsevier.
Ijms 26 05454 g005aIjms 26 05454 g005b
Figure 6. SEM and EDX spectrum of (a) Ceaspina pulcherrima flower, (b) Nervila aragona leaf, and (c) Manihot esculante peel extract TiO2 NPs. Reprinted from ref. [52].
Figure 6. SEM and EDX spectrum of (a) Ceaspina pulcherrima flower, (b) Nervila aragona leaf, and (c) Manihot esculante peel extract TiO2 NPs. Reprinted from ref. [52].
Ijms 26 05454 g006
Figure 7. XRD pattern of green synthesized TiO2 NPs with JCPDS card. Reprinted with permission from [82], with permission of IOP publisher.
Figure 7. XRD pattern of green synthesized TiO2 NPs with JCPDS card. Reprinted with permission from [82], with permission of IOP publisher.
Ijms 26 05454 g007
Figure 8. XRD of TiO2(1), TiO2(2), and TiO2(3) at different T. violacea bulb extract volumes of 10, 20, and 40 mL, respectively. Reprinted with permission from ref. [58] with permission of Elsevier.
Figure 8. XRD of TiO2(1), TiO2(2), and TiO2(3) at different T. violacea bulb extract volumes of 10, 20, and 40 mL, respectively. Reprinted with permission from ref. [58] with permission of Elsevier.
Ijms 26 05454 g008
Figure 9. Antimicrobial activity of Trigonella foenum-graecum derived TiO2 NPs against different microorganisms depicting zones of inhibition of (a) positive control, (b) TF-TiO2 NPs, and (c) dimethyl sulfoxide control. Reprinted with permission from ref. [56], with permission of Elsevier.
Figure 9. Antimicrobial activity of Trigonella foenum-graecum derived TiO2 NPs against different microorganisms depicting zones of inhibition of (a) positive control, (b) TF-TiO2 NPs, and (c) dimethyl sulfoxide control. Reprinted with permission from ref. [56], with permission of Elsevier.
Ijms 26 05454 g009
Figure 10. Antibacterial action mechanism of TiO2 NPs.
Figure 10. Antibacterial action mechanism of TiO2 NPs.
Ijms 26 05454 g010
Figure 11. Simplified diagram illustrating the fundamental principles of semiconductor photocatalysis.
Figure 11. Simplified diagram illustrating the fundamental principles of semiconductor photocatalysis.
Ijms 26 05454 g011
Table 1. Circular economy aspect of chemical synthesis versus specific green synthesis [32,33,34,35].
Table 1. Circular economy aspect of chemical synthesis versus specific green synthesis [32,33,34,35].
FeaturesTraditional Chemical SynthesisGreen-Synthesized TiO2 NPs
Material Cost ReductionBaseline30–50% lower
Photocatalytic EfficiencyLow25% higher
Antimicrobial ActivityLow30% improved
ReusabilityTypically, single useUp to 10 cycles
Waste UtilizationSignificant hazardous wasteReduces plants waste by 90%
ScalabilityEasily scalable but less sustainablePotential for 1 ton/month
Economic Value from WasteNo value recovery from wasteUSD 1000 per ton of plants by-products
Table 2. Comparison between green synthesized titania nanomaterials and with other green synthesized metal nanomaterials.
Table 2. Comparison between green synthesized titania nanomaterials and with other green synthesized metal nanomaterials.
PropertyTiO2
NPs
Silver
NPs
Zinc Oxide
NPs
Gold
NPs
Iron Oxide
NPs
Plant ExtractsAloe vera, neem, green teaCinnamon, neem, green teaHibiscus, aloe vera, turmericMangosteen, green tea,Green tea, eucalyptus, moringa
Photocatalytic
Activity
High (under UV/visible light)ModerateHigh
(UV light)
LowModerate (good under visible light)
Antimicrobial
Properties
ModerateHighModerateHighHigh
Cost of SynthesisLowModerateLowHighLow
Effectiveness in Wastewater TreatmentHigh (removes dyes, heavy metals, organic pollutants)High (antimicrobial, removes organic pollutants)High (removes dyes, heavy metals, and pathogens)Moderate (mainly used for sensing contaminants)High (effective in removing dyes and heavy metals)
Table 3. TiO2 NPs: Synthesis, Characterization and Application by using different plant extract.
Table 3. TiO2 NPs: Synthesis, Characterization and Application by using different plant extract.
Sr. No.PlantsMorphologySize [nm]CharacterizationApplicationReferences
1Kniphofia foliosaIjms 26 05454 i001Spherical----UV–VIS, XRD, SEM TEM, TGA and DTAAntibacterial against Staphylococcus aureus, Escherichia coli, Klebsiella pneumonia, and Streptococcus pyogenes[90]
2Syzygium cuminiIjms 26 05454 i002Irregular structure18UV–vis,
HRTEM,
TEM and
XRD
Picric Acid Degradation
and Anticancer activity
[91]
3Cassia fistulaIjms 26 05454 i003Spherical UV–VIS, XRD, SEM, TEM and TGAAntibacterial against Escherichia coli and Staphylococcus aureus[92]
4Averrhoa bilimbi fruitIjms 26 05454 i004Spherical1–8EDX,
FTIR,
SEM, and
TGA
Antibacterial against S. aureus P. aeruginosa, and Candida albicans fungi Degradation of methylene blue (MB)[88]
5Carica papayaIjms 26 05454 i005Spherical and cages like15.6UV–VIS, XRD, SEM and TEMDegradation of dye[84]
6Trianthema portulacastrumIjms 26 05454 i006Porous crystallites with small round particles10–12XRD, FTIR, SEM and EDXAntifungal against U. tritici[76]
7Jatropha curcas L.Ijms 26 05454 i007Spherical, tetragonal13.0UV–Visible, FESEM, EDS, FT-IR, XRD, DLS, BET and BJHremoval of chemical oxygen demand (COD) and chromium (Cr) from secondary treated TWW[69]
8Azadirachta indicaIjms 26 05454 i008Spherical124XRD, SEM, TEM and FT-IREscherichia coli, Staphylococcus aureus, Bacillus subtilis, Salmonella typhi and Klebsiella pneumonia[64]
9Alcea and ThymeIjms 26 05454 i009Irregularly shaped and spherical10.1XRD, FTIR, FESE and EDXDegradation of methylene blue (MB)[63]
10Jasmine FlowersIjms 26 05454 i010Spherical31–42XRD,
UV−vis and
SEM
Antibacterial against Escherichia coli and Staphylococcus aureus and Degradation of methylene blue (MB)[61]
11Lagenaria siceraria leafIjms 26 05454 i011Irregular303XRD, FTIR, FESEM, EDX,
HRTEM
and UV−vis
photo-degradation of RG-19 dye[60]
12Phyllanthus niruri leafIjms 26 05454 i012Crystalline23UV-DRS,
EDX, FTIR, FESEM
Degradation of methylene blue (MB)[59]
13Syzygium cuminiIjms 26 05454 i013Spherical round shaped22HRSEM,
HRTEM,
EDS,
FTIR,
XRD,
DLS and
BET
lead removal from wastewater[58]
14Tulbhagia violacelaIjms 26 05454 i014Rectangular and Irregular shaped3.0–3.4XRD, SEM, EDX, TEM, and UV−visantioxidant assay and Anticancer activity[58]
15Commelina benghalensisIjms 26 05454 i015Spherical and agglomerated30–200XRD, SEM, EDX, BET, TGA and UV−visDegradation of methylene blue (MB) and SSX[57]
16Wrightia tinctoria leafIjms 26 05454 i016Spherical22FTIR, UV, XRD, HR-TEM and HR-SEMAntifungal against S. aureus, S. faecalis, E. coli, P. vulgaris, E. faecalis, P. aeruginosa, Y. enterocolitica, B. subtilis and fungus C. albicans[56]
17Nervilaaragona leaf,
Ceaspinapulcherrima flower,
Manihotesculante plant extract
Ijms 26 05454 i017Spherical15–28XRD, FTIR, SEM, EDX, TEM and UV−visAntibacterial against E. coli and S. aureus and Degradation of methylene blue (MB)[52]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Shakeel, N.; Piwoński, I.; Iqbal, P.; Kisielewska, A. Green Synthesis of Titanium Dioxide Nanoparticles: Physicochemical Characterization and Applications: A Review. Int. J. Mol. Sci. 2025, 26, 5454. https://doi.org/10.3390/ijms26125454

AMA Style

Shakeel N, Piwoński I, Iqbal P, Kisielewska A. Green Synthesis of Titanium Dioxide Nanoparticles: Physicochemical Characterization and Applications: A Review. International Journal of Molecular Sciences. 2025; 26(12):5454. https://doi.org/10.3390/ijms26125454

Chicago/Turabian Style

Shakeel, Nasir, Ireneusz Piwoński, Parvaz Iqbal, and Aneta Kisielewska. 2025. "Green Synthesis of Titanium Dioxide Nanoparticles: Physicochemical Characterization and Applications: A Review" International Journal of Molecular Sciences 26, no. 12: 5454. https://doi.org/10.3390/ijms26125454

APA Style

Shakeel, N., Piwoński, I., Iqbal, P., & Kisielewska, A. (2025). Green Synthesis of Titanium Dioxide Nanoparticles: Physicochemical Characterization and Applications: A Review. International Journal of Molecular Sciences, 26(12), 5454. https://doi.org/10.3390/ijms26125454

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

Article Metrics

Back to TopTop