Open Access Article
Tan Van
Lam
ab,
Giang Thanh
Tran
ab,
Ngoan Thi Thao
Nguyen
ab,
Ngoc-Kim-Ngan
Phan
ab and
Duyen Thi Cam
Nguyen
*ab
aCenter for Hi-Tech Development, Nguyen Tat Thanh University, Saigon Hi-Tech Park, Ho Chi Minh City, Vietnam
bInstitute of Applied Technology and Sustainable Development, Nguyen Tat Thanh University, Ho Chi Minh City, Vietnam
First published on 4th November 2025
Cerium oxide nanoparticles (CeO2-NPs) have garnered significant interest due to their unique properties as well as a wide range of applications in the environmental and biomedical fields. This work examines the recent advances in the production of CeO2-NPs using various biological agents, including plant extracts and biosources, as sustainable and eco-friendly alternatives to conventional synthetic routes. The formation mechanisms, structural characteristics, and influence of factors on the synthesis of CeO2-NPs were discussed. We found that most CeO2-NPs possessed spherical shapes, small particle sizes (10–100 nm), and large surface areas (12–100 m2 g−1). Bio-mediated CeO2-NPs and their composites had diverse biomedical applications such as antibacterial, antifungal, antioxidant, anticancer, neuroprotective, enzymatic, biosensing, and seed germination activities. Furthermore, CeO2-based composites, such as ZnO–CeO2, Ag-doped CeO2, cellulose/CeO2, CeO2/biochar, and Cu/CeO2, acted as photocatalysts and adsorbents for wastewater treatment applications. The performance of CeO2 and its composites in the removal of organic pollutants such as dyes, antibiotics, nonsteroidal anti-inflammatory drugs, pesticides, and other organic compounds was evaluated. Remarkably, CeO2-NPs and their composites exhibited good adsorption capacities of 46–201 mg g−1 and removal efficiency of up to 99% against heavy metals, dyes, antibiotics, and inorganic contaminants. Finally, this review analyzed the limitations and proposed future research directions for biosynthesized CeO2-NPs and their composites in biomedical and environmental technologies.
CeO2 nanoparticles can be synthesized chemically, physically, or biologically, and among these methods, green synthesis using biological agents has attracted great attention. The bio-mediated green synthesis approach uses biological entities such as plant extracts, microorganisms, and biomolecules to moderate the synthesis of the nanoparticles under mild conditions.13,14 Compared to chemical methods, biosynthesis offers merits such as reduced toxicity, enhanced biocompatibility, and environmental friendliness.15 The phytochemicals, e.g., flavonoids, alkaloids, isoflavones, phenolic compounds, anthocyanins, indoles, and glucosinolates, in plant extracts and proteins in bacterial culture filtrates can function as capping, complexing, and redox agents, which enhance the functionality of green nanomaterials for specific applications.16–19 Despite the above-mentioned advantages, the bioactive compounds present in plant extracts and microbial metabolites significantly affect the morphology, size, and surface properties of the nanoparticles.20 Also, understanding the precise mechanism and controlling the underlying biosynthesis process remain a challenge; therefore, further investigations into the reaction pathways and influencing factors are needed.
Biosynthesized CeO2 nanoparticles have attracted great attention in environmental remediation and biomedical performance.1,21 In the biomedical field, these particles have exhibited strong antibacterial, antifungal, antioxidant, and anticancer activity.22 Hence, the effect of green CeO2 nanoparticles on drug delivery, wound healing, and neuroprotection should be significant. Green CeO2 nanoparticles also exhibit catalytic activity, allowing enzyme-like functions and hence increasing their potential in therapeutic uses.23 In environmental chemistry, they have been extensively studied for pollution control, demonstrating high efficiency in the photocatalytic degradation of toxic organic contaminants in water, such as dyes, antibiotics, pesticides, and nitroaromatic compounds.24–27 Additionally, the strong adsorption capability of green CeO2 nanoparticles enables the removal of heavy metals, e.g., Cr6+, Cd2+, and Pd2+,28 inorganic ions including F− ions,29 and many hazardous pollutants such as synthetic dyes from wastewater.30,31 The versatility of CeO2 nanoparticles also extends to sensing technologies, where green CeO2 nanoparticles can improve the detection of biological substances and environmental toxins.32–34
Recently, several reviews have examined the biosynthesis of CeO2 nanoparticles using biological sources for various applications, suggesting rapid advancements in this field.22,35–37 For instance, Naidi et al.22 and Pansambal et al.36 evaluated green synthesis strategies employing diverse bio-sources, including plants, microbes, and microbial biomass. Both studies also assessed the impact of fabrication factors on the properties of plant extract-mediated CeO2 nanoparticles and multiple applications, such as photocatalysis, antimicrobial activity, antioxidant activity, anticancer, and cytotoxicity. However, these reviews lack a comprehensive analysis and in-depth discussion on their mechanisms for biological applications. Very recently, Vinitha et al.35 reviewed different synthesis approaches from top-down to bottom-up, and the use of various bio-sources such as plants, microbes, and algae for the bio-mediated synthesis of CeO2 nanoparticles. However, their work only provided a general overview of synthesis methods and did not critically discuss many recent advancements.
Herein, the present review critically examined the biosynthesis methods and exhibited the wide-ranging potential applications of CeO2 nanoparticles. We profoundly assessed their state of the art biological applications, including neuroprotection, enzymatic activity, seed germination, and sensing technologies of green CeO2 nanoparticles. The role of CeO2 nanoparticles as photocatalysts and adsorbents for environmental remediation was also discussed. This review is expected to bridge the knowledge gap and give insights into the limitations and future prospects of biosynthesized CeO2 nanoparticles.
The synthesis of CeO2 nanoparticles using phytochemicals from plant extracts is detailed as follows. Firstly, plant sources (leaf, flowers, etc.) are extracted using water or alcohol as a single or mixed solvent. The as-prepared phytochemicals are mixed with cerium salt precursors to form a complex between the cerium ions and bioactive compounds. During this stage, ammonia solution or sodium hydroxide is often added to maintain an alkaline medium (pH > 7), which is essential for the oxidation of Ce3+ to Ce4+. In the plant extract-mediated synthesis of CeO2 nanoparticles via the co-precipitation method, particularly, an alkaline medium is required to ensure the easy formation of CeO2 nanoparticles.41,42
Once the cerium ions form complexes with the bio-compounds, this process leads to the formation of Ce–O–Ce bridges.43 The resulting precipitate is collected and prepared for calcination. Upon thermal decomposition at temperatures above 300 °C, as shown in Table 1, residual organic matter and water evaporate, leaving behind an amorphous cerium oxide network that crystallizes into CeO2. Bioactive substances, such as tannins, terpenoids, and phenolic acids, promote the nucleation and growth of nanoparticles. Indeed, natural capping agents, including phytochemicals, e.g., flavonoids, polyphenols, and alkaloids, and polysaccharides, e.g., starch, cellulose, and chitosan, help to control the particle size and prevent aggregation, enhancing the stability of the nanoparticles.44,45 Thus, bio-compounds not only drive the redox process but also regulate nanoparticle growth, leading to the formation of eco-friendly ceria nanoparticles with desirable properties.
| Biosource | Synthesis method | Synthesis temperature (°C) | Time (h) | Surface area (m2 g−1) | Pore volume (cm3 g−1) | Pore diameter (nm) | Particle size (nm) | Morphology | Band gap (eV) | Ref. |
|---|---|---|---|---|---|---|---|---|---|---|
| Cow urine | Reflux | 150 | 1.5 | — | — | — | 100 | Flake-like shape | 4.53 | 32 |
| Honey | Reflux | 150 | 1.5 | — | — | — | 100 | Flake-like shape | 4.65 | 32 |
| Moringa oleifera seeds | Calcination | 400 | 3 | 57 | — | 3.4 | 12 | Porous foam-like shape | 2.5 | 49 |
| Equisetum ramosissimum Desf. | Calcination | 400 | 2 | — | — | — | 25 | Spherical shape | 3.14 | 50 |
| Portulaca oleracea | Calcination | 500 | 2 | — | — | — | 16 | Cubic | — | 51 |
| Azadirachta indica leaf | Calcination | 400 | 5 | — | — | — | 40 | Distorted spherical | — | 25 |
| Centella asiatica leaf | Solution combustion | 570 | 20 min | — | — | — | 45 | Spherical shape | 3.7 | 52 |
| Sapindus mukorossi seed extract | Calcination | 500 | — | 40 | — | 1.2 | 60–100 | Irregular shape | 2.8 | 53 |
| Echinacea purpurea leaf | Calcination | 450 | 1 | 100 | 0.03 | 20 | 12 | Irregular spherical and spherical | — | 54 |
| Melastoma sp. leaf | Calcination | 500 | 2 | — | — | — | 20 | Spherical shape | — | 55 |
| Annona reticulata leaf | Combustion method | 450 | 30 min | 62 | 0.31 | 27 | 7 | — | — | 56 |
| Stevia rebaudiana leaf | Calcination | 700 | 5 | — | — | — | 50 | Sphere-like | — | 57 |
| Withania somnifera roots | Combustion method | 350 | 15 min | — | — | — | 8–12 | Spherical shape | 3.5 | 58 |
| Punica granatum peel | Reflux | 80 | 24 | 62 | 0.04 | — | 28 | Nanorods | 3.1 | 59 |
| Banana peel | Calcination | 400 | 2 | 12 | 0.02 | 1.2 | 21 | Spherical shape | — | 60 |
| Vigna radiate extract | Calcination | 400 | 2 | — | — | — | 24 | Flake-like shape | 3.2 | 61 |
| Syzygium aromaticum (clove bud) | Calcination | 600 | 4 | — | — | — | 30–60 | — | — | 62 |
| Morinda tinctoria | Precipitation | Room | 10 min | — | — | — | 21 | Spherical shape | — | 47 |
| Matricaria recutita extract | Calcination | 600 | 3 | 30 | 0.004 | 8.4 | 12 | Spherical shape | 1.9 | 63 |
| Chenopodium quinoa leaf extract | Precipitation | 80 | 6 | — | — | — | 7–10 | Spherical shape | — | 64 |
| Azadirachta indica leaf extract | Precipitation | 80 | 6 | — | — | — | 7–10 | — | — | 65 |
| Azadirachta indica extract plant | Calcination | 500 | 3 | — | — | — | 13 | Irregular shape | 2.4 | 66 |
| Plectranthus barbatus leaf extract | Calcination | 500 | 3 | — | — | — | 11 | Spherical shape | 3.0 | 67 |
| Rosa damascena extract | Calcination | 500 | 1.7 | — | — | — | 29 | Sponge-like shape | — | 68 |
| Deinococcus radiodurans culture filtrate | Calcination | 200 | 5 | — | — | — | 2.4 | Spherical shape | — | 23 |
On the one hand, the bio-mediated synthesis of CeO2 nanoparticles is completed without the addition of alkaline solution. This means that the synthesis process can be carried out using only two components including cerium precursors and plant extracts. For instance, the green synthesis of CeO2 nanoparticles and CeO2-based nanocomposites from plant extracts of Pelargonium hortorum,46Morinda tinctoria,47 and Ocimum sanctum48 has been reported. After complexation is finished, the complex product is dried and calcined at 400–500 °C to form CeO2 or CeO2-based composites. Even without alkaline agents, the synthesized products still exhibit the characteristic crystal structure of CeO2 according to XRD analysis, where diffraction peaks corresponding to the (111), (200), (220), and (311) planes were found.46–48 These findings indicate the presence of a fluorite cubic structure in the bio-mediated synthesized CeO2 nanoparticles.
Avoiding the use of alkaline media during complexation offers both advantages and disadvantages. In terms of advantages, a key benefit is the elimination of strong alkaline reagents, thereby reducing the risk of unwanted side reactions such as excessive oxidation and uncontrolled precipitation. This approach aligns with green chemistry principles, minimizing chemical waste and enhancing the eco-friendliness and biocompatibility. However, the absence of alkaline agents may slow the redox kinetics, requiring longer reaction times. Therefore, although this method improves the sustainability, optimizing the reaction conditions remains essential for producing high-quality CeO2 nanoparticles.
Despite the growing interest in plant-based nanoparticle synthesis, the precise identification of key phytochemicals responsible for reduction, capping, and stabilization remains under investigation.69 The complex interactions between bioactive compounds and metal ions in solution are not yet fully understood. Additionally, the variability in phytochemical composition across plant species, nutrient and growth conditions, and extraction methods further complicates the identification of specific compounds involved in the synthesis process. Thus, more investigations are needed to elucidate the role of individual phytochemicals and the underlying formation mechanisms of green CeO2 nanoparticles from bio-mediated synthesis methods.
Table 1 shows that the synthesis of green CeO2 nanoparticle primarily uses bio-sources from various parts of plants and a few products from animals including cow urine, honey, and bacterial culture filtrates. These bio-natural materials function as complexing and capping agents and are significantly involved in the formation of CeO2 nanoparticles. Variations in the ratio of bioactive compounds in the precursor solution can lead to considerable variations in the properties of the synthesized CeO2 nanoparticles. For instance, Mohammed et al.50 indicated that employing various amounts of Equisetum ramosissimum Desf. extract in the process for the synthesis of CeO2 nanoparticles had a profound effect on their crystallite sizes. Specifically, with an increase in extract volume from 5 mL to 15 mL, the average crystallite size decreased from 8 nm to 6 nm. Nevertheless, with a subsequent increase in extract volume to 20 mL, the average crystallite size increased to 12 nm. It was explained that a larger amount of extract led to better complexation between Ce3+/Ce4+ and phytochemicals, thereby reducing the aggregation during calcination. When the amount of extract added was too large, the pH of the solution might change, affecting the formation of CeO2 crystals.
Similarly, Mamatha et al.52 observed that increasing the volume of Centella asiatica leaf extract from 10 mL to 30 mL in the synthesis of CeO2 reduced its average crystallite size from 18 nm to 8 nm. This fluctuation can be attributed to the dual role of bioactive compounds as both capping and complexation agents. When a sufficient amount of extract (5–15 mL) was used, more bioactive molecules are available to keep the growing CeO2 nuclei stable. This prevented the CeO2 nuclei from clumping together and helped form smaller crystallites. The authors also interpreted that if more extract was added, residual bio-agents could slow down nucleation. Instead of forming new nuclei, the available cerium ions continue to accumulate on existing nuclei, resulting in fewer but larger particles. This shift in growth dynamics leads to an overall increase in crystallite size at higher extract concentrations.
Changes in the crystallite size and band gap of CeO2 nanoparticles were also reported. Indeed, Mohammed et al.50 indicated the influence of crystallite size on the band gap energy of CeO2 nanoparticles due to the quantum confinement effect. The results showed that a decrease in the average crystallite size of green-synthesized CeO2 nanoparticles from 8 nm to 6 nm resulted in an increase in their band gap energy from 3.14 eV to 3.37 eV. Conversely, Mamatha et al.52 reported that with a decrease in the mean crystallite size of Centella asiatica leaf extract-synthesized ceria nanoparticles from 18 nm to 8 nm, their band gap energy was reduced from 3.70 eV to 3.60 eV. This change happens because of the quantum confinement effect, where smaller crystallites trap more charge carriers. This trapping limits the electron wavelength and requires more energy for electronic transitions.70 However, variations in the observed trends may arise due to differences in the synthesis conditions, defect density, and presence of organic residues affecting the electronic structure of the nanoparticles.
Another key factor that influences the properties of bio-mediated CeO2 nanoparticles is their synthesis conditions. As shown in Table 1, various conditions are employed, i.e., high-temperature calcination (200–700 °C) for extended durations (1–5 h). Precipitation processes are also performed at room temperature in a shorter time frame (6–10 min) under alkaline conditions. Other methods, such as solution combustion at 350–450 °C for 15–30 min and reflux techniques, also contribute to variations in the properties of CeO2.
Table 1 indicates that different synthesis methods lead to variations in the internal properties of CeO2 nanoparticles, which depend on the synthesis temperature and synthesis duration. For example, Manojkumar et al.60 synthesized CeO2 nanoflowers using banana peel extract via a calcination method at 400 °C for 2 h. The resulting nanoparticles had a surface area of 12 m2 g−1 and average particle size of 21 nm. In another report, Surendhiran et al.71 synthesized CeO2 nanoparticles using Moringa oleifera seed extract via calcination at 400 °C for 3 h, obtaining a surface area of 57 m2 g−1 and average particle size of 12 nm. These results can be explained by the long synthesis period, which facilitates the formation of larger nuclei before significant particle growth, producing smaller-sized particles. Also, a longer duration results in better interaction between the nanoparticles and the capping agents, which can promote the dispersion and prevent the aggregation of the particles, resulting ultimately in a smaller particle size and a higher surface area. However, opposite results have been reported; a longer synthesis time resulted in a larger particle size. For instance, Rani et al.25 synthesized CeO2 nanoparticles using Azadirachta indica leaf extract via calcination at 400 °C for 5 h, obtaining an average particle size of 40 nm. This indicates that the synthesis time does not have a consistent effect on the structural properties, although the effects of synthesis temperature and time on the characteristics of CeO2 nanoparticles are undeniable.
In another study, Anand et al.72 investigated the effect of synthesis temperature on the particle size of Nyctanthes arbor-tristis extract-based cerium oxide nanoparticles. The outputs showed a lack of systematic correlation between temperature and particle size. Calcination of the sample at 100 °C, 200 °C, and 300 °C resulted in average particle diameters of 2 nm, 14 nm, and 4 nm, respectively. This difference can be attributed to the interaction between nucleation and growth dynamics. At 100 °C, slow nucleation and limited crystal growth promoted smaller particle sizes (2 nm). However, at higher temperatures, the decomposition of organic components in the plant extract alters the capping and complexation behavior of the bioactive compounds with cerium ions, influencing the aggregation and growth of the nanoparticles.
Metal and metal oxide-based CeO2 composites significantly decreased the aggregation of metal/CeO2 composites and improved their band gap energy (Table S1 in the SI document). For example, Ahmad et al.47 successfully fabricated Ag/CeO2 composites using Morinda tinctoria plant extract. The results indicated that the green Ag/CeO2 composites had a mean crystal size of approximately 17 nm compared to that of 21 nm for the cerium oxide nanoparticles synthesized using the same plant extract. Additionally, the band gap of Ag/CeO2 composites also exhibited a lower value of 2.50 eV compared to that of 2.74 eV for CeO2 nanoparticles. A similar trend was also observed in the synthesis of CeO2/ZnO composites with the addition of Sapindus mukorossi extract.53 The green CeO2/ZnO composites had a smaller average particle size of 45 nm compared to that of 60 and 80 nm for ZnO and CeO2, respectively. Also, the band gap of the CeO2/ZnO composites had a lower value of 2.6 eV compared with that of 2.8 and 3.1 eV for ZnO and CeO2, respectively. This is due to the formation of a heterojunction between ZnO and CeO2, which possesses a lower band gap energy. When these two components integrate, they facilitate charge transfer, leading to a reduction in the overall band gap and improved photocatalytic efficiency by enhancing charge separation and minimizing electron–hole recombination. However, more analyses, such as photoluminescence spectroscopy, should be conducted to evaluate the electron–hole recombination rate. Electrochemical impedance spectroscopy should be also conducted to investigate the photogenerated charge transfer performance of CeO2-based catalysts.
Notable improvements in the band gap energy and particle size of biopolymer/CeO2 composites were recently reported (Table S1). For example, Shalini et al.77 used the Oldenlandia umbellata extract for the bio-fabrication of a gelatin/alginate/CeO2 composite. The outcomes showed that the band gap and average particle size of this composite were 2.65 eV and 73 nm, respectively. Meanwhile, the green CeO2 nanoparticles synthesized using the same protocol had a band gap of 3.1 eV and an average particle size of 94 nm. The authors clarified that with the addition of gelatin/alginate, the interactions of the oxygen atoms of CeO2 with gelatin/alginate are enhanced throughout the synthesis. The chemical bonding generates electron confinement in the O 2p orbitals, leading to a reduced bandgap in the gelatin/alginate/CeO2 composite. At the same time, the polymer matrix could serve as a template, controlling the nucleation and growth of the CeO2 nanoparticles and inhibiting uncontrolled particle agglomeration.
| Materials | Plant source | Main applications | Ref. |
|---|---|---|---|
| CeO2 | Gomutra and honey | Antioxidant | 32 |
| CeO2 | Moringa oleifera | Antioxidant; antibacterial: S. aureus and E. coli | 49 |
| CeO2 | Equisetum ramosissimum | Antimicrobial: S. mutans | 50 |
| CeO2 | Cannabis sativa | Antibacterial: Salmonella enteritidis, Escherichia coli, anticancer: breast cancer | 80 |
| CeO2 | Centella asiatica | Neuroprotective: neuroblastoma (SH-SY5Y) | 52 |
| CeO2 | Echinacea purpurea | Anticancer: MCF7 and HCT116 | 54 |
| Gelatin–sodium alginate–CeO2 | Oldenlandia umbellata | Anticancer: breast cancer | 77 |
| CeO2 | Melastoma | Antibacterial: E. coli and K. pneumonia | 55 |
| ZnO–CeO2 | Syzygium aromaticum | Antibacterial: E. coli and S. aureus | 81 |
| Ni-doped CeO2 | Pedalium murex | Antifungal: Candida albicans and Candida krusei | 74 |
| CeO2 | Stevia rebaudiana | Antibacterial: P. aeruginosa and E. faecalis | 57 |
| CeO2 | Aspergillus terreus | Neuroprotective: SH-SY5Y cells | 58 |
| Ag-doped CeO2/rGO | Punica granatum | Antibacterial: E. coli and S. aureus | 59 |
| CeO2 | Musa paradisiaca | Antimicrobial: E. coli, B. cereus, and S. aureus | 60 |
| Ag-decorated CeO2 | Morinda tinctoria | Antimicrobial: E.coli and S. aureus | 47 |
| CeO2 | Jacaranda mimosifolia | Antimicrobial: E. coli and B. subtilis | 82 |
| CeO2 | Carica papaya | Antimicrobial: S. mutans, P. aeruginosa, B. subtilis, E. coli, A. fumigatus, and A. niger | 83 |
| CeO2 | Matricaria recutita | Antimicrobial: E. coli, S. aureus and K. pneumoniae | 63 |
| CeO2 | Chenopodium quinoa | Antifungal: Ustilago tritici | 64 |
| Zn–Ni dual-doped CeO2 | Cucurbita pepo | Anticancer: Huh-7 cell line | 75 |
| CeO2 | Colocasia esculenta | Seed germination of mung bean | 84 |
| CeO2 | Morinda citrifolia | Antioxidant defense system in Vigna mungo | 85 |
| Cellulose/CeO2 | Lycopersicon esculentum | Antibacterial: E. coli and S. aureus | 79 |
| Chitosan–CeO2 | Tea leaf | Antimicrobial: E. coli, S. aureus, and B. cinerea | 78 |
| CeO2 | Osmium Sanctum | Antibacterial | 86 |
| CeO2 | Scoparia dulcis | Anticancer: adenocarcinomic lung | 87 |
| CeO2 | Acacia concinna | Antibacterial: S. pneumoniae and E.coli | 31 |
| CeO2 | Cannabis sativa | Enzymatic: acetylcholinesterase inhibition | 80 |
![]() | ||
| Fig. 1 (a) Minimum inhibitory concentration (MIC) and minimum bactericidal concentration (MBC) of CeO2 nanoparticles synthesized using the Cannabis sativa leaf extract against four types of bacteria. Adapted with permission from ref. 80 Copyright (2023), Elsevier. (b) The zone of inhibition for E. coli and S. aureus using CeO2 and Ag/CeO2 nanoparticles. Reproduced with permission from ref. 47 Copyright (2023), Elsevier. (c) The effect of CeO2 nanomaterials on bacteria cells. Reproduced with permission from ref. 88 Copyright (2024), Elsevier. | ||
CeO2 nanoparticles show antibacterial activity because they can release antibacterial agents such as reactive oxygen species that disrupt bacterial cells. For example, Surendhiran et al.71 successfully synthesized CeO2 using Moringa oleifera extract, which exhibited antibacterial activity against E. coli and S. aureus. Specifically, the authors suggested that CeO2 disrupts the enzymes in bacteria by generating hydrogen peroxide, which eventually leads to the death of bacterial cells. Regarding the outcomes, the synthesized CeO2 nanoparticles are more effective against Gram-negative bacteria than Gram-positive bacteria due to their thinner peptidoglycan. Additionally, the surface structures of CeO2 nanoparticles can physically damage or inhibit bacteria growth. Moreover, the oxygen vacancies in metallic oxide materials are beneficial for the generation of reactive oxygen species. It was reported that the doping of Ag nanoparticles into CeO2 materials synthesized from Punica granatum leaf extract can enhance the formation of oxygen vacancies.59 As evidenced by UV-vis adsorption, the shift in the peak at 385 nm to 450 nm indicates the presence of oxygen vacancy defects due to the substitution of Ag+ ions in the lattice and self-doping of Ce3+.
Moreover, the formation of oxygen vacancies is known to enhance the antifungal action of CeO2-based nanocomposites. Indeed, Lohitha and Albert74 indicated that the doping of Ni into the crystal lattice of CeO2 facilitated the formation of oxygen vacancies (Fig. 2). Specifically, based on XPS results, the formation of defects related to oxygen vacancies between the 4f and 2p levels of cerium and oxygen ions appeared through the emission peaks at 450 and 467 nm, respectively. Additionally, a high oxygen vacancy concentration can facilitate the flow of oxygen atoms, and hence enhance the redox reactions on the surface of Ni/CeO2 materials, which exhibit better antifungal properties. As a result, Ni-3%/CeO2 showed a large zone of inhibition of 13 nm and 15 nm for Candida albicans and Candida krusei, respectively.
![]() | ||
| Fig. 2 (a) Process for the synthesis of CeO2 nanoparticles using the Pedalium murex leaf extract. (b) Energy level diagram of the photocatalytic process on Ni-doped CeO2 nanoparticles. Reproduced with permission from ref. 74 Copyright (2025), Springer. (c) UV-Vis of CeO2, CeO2/reduced graphene oxide and Ag-doped CeO2/reduced graphene oxide. Reproduced with permission from ref. 59 Copyright (2024), Elsevier. | ||
![]() | ||
| Fig. 3 (a) The steps in the DPPH experiments, and (b and c) the antioxidant activity percentages of CeO2 nanoparticles. Reproduced with permission from ref. 32 Copyright (2024), Elsevier. | ||
![]() | ||
| Fig. 4 (a) The cell viability of CeO2 nanoparticles at different concentrations and compared to cisplatin-treated and untreated samples. CeSD5, CeSD15 and CeSD25 denote CeO2 synthesized using 5 mL, 15 mL and 25 mL of the Scoparia dulcis extract, respectively. Reproduced with permission from ref. 87 Copyright (2022), Springer. (b) The anticancer mechanism of CeO2 nanoparticles through ROS production. Reproduced with permission from ref. 89 Copyright (2022), MPDI. (c) The selective anticancer mechanism of CeO2 for cancer and healthy cells. Reproduced with permission from ref. 90 Copyright (2022), Springer. | ||
Eco-friendly synthesis methods utilizing plant extracts enhances the biocompatibility of CeO2, making it a safer alternative for potential cancer therapies by minimizing damage to healthy cells. For example, CeO2 nanomaterials bio-mediated by Cannabis sativa extract demonstrated good anticancer activity by selectively targeting breast cancer cells (MCF-7) while preserving normal cells (mammary epithelial cells).80 As a result, green CeO2 was more toxic to cancer cells than normal cells, as evidenced by its lower IC50 value (25 μg mL−1) for MCF-7 compared to that (75 μg mL−1) for normal cells. These outcomes proved the potential of green CeO2 nanoparticles for targeted anticancer therapy. The main selective mechanism depending on the surrounding environment is explained in Fig. 5c. Accordingly, green CeO2 nanoparticles exhibit anticancer activity due to coexistence of Ce3+ and Ce4+, along with oxygen vacancies on their surface.90 The nanoceria forms help to scavenge ROS at the neutral or weakly basic pH (7.2–7.5) of healthy cells, while acting as ROS producers in the weakly acidic tumor cell environment (5.6–6.2). This disparity endows CeO2 nanoparticles with activity for targeted anticancer therapy.
![]() | ||
| Fig. 5 (a) The bio-mediated synthesis of the CeO2, Cd/CeO2 and Ru/CeO2 nanomaterials using Vigna radiata, (b) XRD spectra of the CeO2, Cd/CeO2 and Ru/CeO2 nanomaterials, and (c) cyclic voltammogram response of the pure glassy carbon electrode (1), glucose (19 μM)/GCE (2), the glucose/CeO2-modified electrode (3), the glucose/(Cd/CeO2)-modified electrode (4) and the glucose/(Ru/CeO2)-modified electrode (5) in glucose solution. Reproduced with permission from ref. 61 Copyright (2024), Elsevier. | ||
Mamatha et al.52 successfully synthesized CeO2 material using Centella asiatica leaf extract for protecting SH-SY5Y neuronal cells from oxidative stress caused by H2O2. In this study, the SH-SY5Y cells were pretreated with H2O2 for 24 h, and then treated with CeO2 to investigate their activity. The treatment of CeO2 nanoparticles at 40 μg mL−1 significantly reduced the cytotoxicity caused by oxidative stress from H2O2 by 74.44% compared to the control cells. In addition, the leakage test of lactate dehydrogenase (LDH), an important enzyme in most living cells, was also performed to determine the integrity of the cell membrane and cell death. Accordingly, the bio-mediated CeO2 nanoparticles reduced LDH leakage by two-times compared to the cells treated with only H2O2. Similarly, Sultana et al.58 synthesized CeO2 nanoparticles using Aspergillus terreus extract to counteract rotenone-induced toxicity in SH-SY5Y cells. The outcomes revealed that the ROS level was reduced by 59% at 80 μg per mL CeO2 nanoparticles for SH-SY5Y cells compared to the rotenone-treated sample. Although these results might not be comparable with that of CeO2 synthesized using Centella asiatica,52 the Aspergillus terreus extract-mediated CeO2-treated cells still showed a significant 79% reduction in LDH release compared to the cells exposed to rotenone alone.
| Materials | Plant source | Main applications | Ref. |
|---|---|---|---|
| CeO2 | Gomutra and honey | MO dye degradation | 32 |
| CeO2 | Moringa oleifera | CR and MG dye degradation | 49 |
| CeO2 | Portulaca oleracea | MO and MB dye degradation | 51 |
| Ag-doped CeO2@SnO2 | Azadirachta indica | MNZ photocatalytic degradation | 25 |
| CeO2@ZnO | Sapindus mukorossi | Eriochrome black T dye and endosulfan pesticide photocatalytic degradation | 53 |
| Gelatin sodium alginate CeO2 hydrogel | Oldenlandia umbellata | MR photocatalytic degradation | 77 |
| CeO2 | Spirulina platensis | MB dye photocatalytic degradation | 94 |
| CeO2 | Annona reticulata | Catalyst for the production of biodiesel using A. reticulata seeds oil | 56 |
| ZnO–CeO2 | Syzygium aromaticum | MR photocatalytic degradation | 81 |
| CeO2 | Stevia rebaudiana | Tetracycline antibiotics photocatalytic degradation | 57 |
| ZnO–CeO2 | Parkia speciosa Hassk. | Nitroaromatic compound photocatalytic reduction | 27 |
| Ag-doped CeO2/rGO | Punica granatum | MB photocatalytic degradation | 59 |
| CeO2 | Musa paradisiaca | MO photocatalytic degradation | 60 |
| Ag-decorated CeO2 | Morinda tinctoria | Bromophenol blue photocatalytic degradation | 47 |
| CeO2 | Jacaranda mimosifolia | MB photocatalytic degradation | 82 |
| CeO2 | Matricaria recutita | MB and MO photocatalytic degradation | 63 |
| CeO2 | Cucurbita pepo | MB photocatalytic degradation | 95 |
| CeO2 | Morinda citrifolia | Methyl red and Acid Red 87 photocatalytic degradation | 85 |
| Cellulose/CeO2 | Lycopersicon esculentum | RhB photocatalytic degradation | 79 |
| CeO2 | Azadirachta indica | Naproxen and piroxicam drug photocatalytic degradation | 66 |
Based on these mechanisms, some researches focused on the synthesis of CeO2 nanoparticles mediated by green resources (Fig. 6). For example, the photocatalytic performance of CeO2 nanoparticles synthesized using Spirulina platensis extract was evaluated for degrading methylene blue dye.94 The CeO2 nanoparticles exhibited a superior degradation efficiency of 98% for 5 ppm MB dye using a concentration of 0.3 g L−1. Additionally, the photocatalytic activity of green CeO2 synthesized using Spirulina platensis also was outstanding compared to pure CeO2 synthesized without plant extract. In contrast, the CeO2 nanoparticles synthesized using Jacaranda mimosifolia extract achieved a slightly lower degradation efficiency of 95% at the same MB dye concentration but required a higher nanoparticle dosage of 0.5 g L−1.82 These results indicated that the Spirulina platensis-mediated CeO2 nanoparticles are a more effective catalyst for the degradation of MB dye at a lower dosage.
![]() | ||
Fig. 6 (a) The dye degradation mechanism of CeO2 synthesized using the Moringa oleifera seed extract. Reproduced with permission from ref. 71 Copyright (2024), Elsevier. (b) MB degradation percentage of CeO2 and Ag@CeO2-doped reduced graphene oxide after 5 cycles. Reproduced with permission from ref. 59 Copyright (2024), Elsevier. (c) The comparison between chemical CeO2 and biological CeO2 at different ratios of 2 : 1 and 1 : 1 for MB removal. Reproduced with permission from ref. 94 Copyright (2024), Springer. | ||
In addition to pure CeO2 nanoparticles, doping CeO2 with other materials significantly enhanced its photocatalytic activity. For instance, although the CeO2 nanoparticles synthesized using Punica granatum extract achieved a degradation efficiency of 65% for MB dye, Ag-doped CeO2 could further boost the efficiency to 100% within 60 min.59 This can be explained by the fact that Ag doping and surface modification with plant extract might promote charge separation and reduce electron–hole recombination. In another study, the CeO2 nanoparticles synthesized using Syzygium aromaticum extract achieved a 89% degradation efficiency.81 However, the optimal results were recorded using ZnO-doped CeO2, resulting in an improved photocatalytic performance of 94% degradation efficiency. This synergistic effect makes the application of CeO2-based doped materials a promising approach for more effective environmental remediation through photocatalysis pathways.
![]() | ||
| Fig. 7 (a) The degradation percentage of CeO2, SnO2, Ag–CeO2@SnO2 synthesized using the Azadirachta indica extract. Reproduced with permission from ref. 25 Copyright (2024), Elsevier. (b) Scavenger testing for the photocatalytic degradation of antibiotics by CeO2 synthesized via the Stevia rebaudiana leaf extract. Reproduced with permission from ref. 57 Copyright (2024), Elsevier. (c) The degradation path of tetracycline and metronidazole in the photocatalytic process by GC-MS analysis. Reproduced with permission from ref. 25 Copyright (2024), Elsevier. | ||
889 ng L−1.96 The contamination of NSAIDs in wastewater has seriously impacts on aquatic organisms and even human related to some diseases such as endocrine disruption, locomotive disorders, and body deformations.97 Owing to their outstanding performance in the photocatalytic removal of antibiotics, CeO2 nanomaterials synthesized using plant extracts are also applied to remove nonsteroidal anti-inflammatory drugs (NSAIDs) from wastewater. For example, Quddus et al.98 carried out a study using CeO2 nanomaterials synthesized using neem leaf extract for the removal of NSAIDs, i.e. piroxicam and naproxen. In this study, the degradation efficiency of piroxicam and naproxen was 97% and 89% in 60 min and 80 min, respectively. These results were regarded competitive with other chemically and complicated synthesized materials. For example, Ag–Fe3O4@Ca–Al synthesized through many steps exhibited the piroxicam degradation percentage of 97% in 180 min.99 Eventually, the MIL-53(Al)@TiO2 porous materials showed the photocatalytic degradation of 80.3% naproxen in 240 min.100 Accordingly, CeO2 nanoparticles synthesized using plant extracts may be more efficient than other chemically synthesized materials in enhancing the photocatalytic performance, eco-friendliness, and sustainability for various environmental applications.
![]() | ||
| Fig. 8 (a) The type II heterojunction of ZnO/CeO2 for photocatalytic degradation. (b) The type Z-scheme heterojunction of ZnO/CeO2 for photocatalytic degradation. (c) The endosulfan removal using ZnO, CeO2, and ZnO/CeO2. (d) The degradation pathway of endosulfan. Reproduced with permission from ref. 53 Copyright (2023), the Royal Society of Chemistry. | ||
| Materials | Biosource | Pollutants | Adsorption conditions | Main findings | Reusability | Ref. |
|---|---|---|---|---|---|---|
| CeO2 nanoparticles | Azadirachta indica leaf extracts | Cr6+ ions | pH: 4, Cr6+ concentration: 30 mg L−1, dosage: 0.7 g L−1, 5 h | Adsorption capacity: 48 mg g−1, efficiency: 93% | — | 65 |
| CeO2 nanoparticles | Azadirachta indica leaf extracts | Cd2+ ions | pH: 4, Cd2+ concentration: 30 mg L−1, dosage: 0.7 g L−1, 5 h | Adsorption capacity: 46 mg g−1, efficiency: 90% | — | 65 |
| CeO2 nanoparticles | Pisum sativum pod extract | Pb2+ ions | pH: 6, Pb2+ concentration: 100 mg L−1, dosage: 0.2 g L−1, 3 h | Efficiency: 87%, maximum adsorption capacity: 100 mg g−1 | — | 28 |
| CeO2 nanoparticles | Pisum sativum pod extract | Cr6+ ions | pH: 3, Cr6+ concentration: 100 mg L−1, dosage: 0.2 g L−1, 3 h | Efficiency: 94%, maximum adsorption capacity: 125 mg g−1 | — | 28 |
| CeO2 nanoparticles | Pisum sativum pod extract | Cd2+ ions | pH: 6, Cd2+ concentration: 100 mg L−1, dosage: 0.2 g L−1, 3 h | Efficiency: 88%, maximum adsorption capacity: 100 mg g−1 | — | 28 |
| CeO2 nanoparticles | Citrus limon peel extract | U6+ ions | pH: 4, U6+ concentration: 0.1 mg L−1, dosage: 0.3 g L−1, 1 h | Efficiency: 95%, maximum adsorption capacity: 46 mg g−1 | — | 103 |
| CeO2 nanoparticles | Acacia concinna fruit extract | Reactive blue | pH: 3, dye concentration: 50 mg L−1, dosage: 5 mg, 3 h | Adsorption capacity: 189 mg g−1, efficiency: 99% | 5 cycles: removal efficiency: 98% (1st cycle) to 74% (5th cycle) | 31 |
| CeO2/biochar composite | Saccharum officinarum extract | Methylene blue | pH: 10, dye concentration: 50 mg L−1, dosage: 0.5 g L−1, 1.25 h | Adsorption capacity: 96 mg g−1 | 5 cycles: removal efficiency: 96% (1st cycle) to 94% (5th cycle) | 104 |
| CeO2 nanoparticles | Saccharum officinarum extract | Methylene blue | pH: 10, dye concentration: 50 mg L−1, dosage: 0.5 g L−1, 1.25 h | Adsorption capacity: 50 mg g−1 | — | 104 |
| Cu/CeO2 composite | Murraya koenigii extract | Congo red | pH: 6.5, dye concentration: 10 mg L−1, dosage: 0.2 g L−1, 35 min | Efficiency: 98% | — | 73 |
| CeO2 nanoparticles | Echinacea purpurea extract | Cefoperazone | Antibiotic concentration: 100 mg L−1, dosage: 0.5 g L−1, 3 h | Adsorption capacity: 201 mg g−1 | — | 54 |
| CeO2 nanoparticles | Litchi chinensis | Fluoride | pH: 7, fluoride concentration: 10 mg L−1, dosage: 7 mg L−1, 35 min | Maximum adsorption capacity: 167 mg g−1 | — | 29 |
![]() | ||
| Fig. 9 (a) Process for the synthesis of CeO2 nanoparticles using the green peapod extract. (b) SEM images of green ceria nanoparticles. (c) FTIR spectrum of green ceria. (d and e) Langmuir and Freundlich isotherm models for Pb(II), Cr(VI), and Cd(II) adsorption by CeO2 nanoparticles, respectively. Adapted with permission from ref. 28 Copyright (2024), Springer. | ||
Other studies have also confirmed the high adsorption capacity of green ceria nanoparticles for heavy metal adsorption. Masood et al.65 evaluated the removal efficiency of green CeO2 nanoparticles synthesized using Azadirachta indica leaf extract for the adsorption of Cr6+ and Cd2+ ions. Their findings showed that under the optimal conditions of pH 4, Cr6+ and Cd2+ concentration of 30 mg L−1, CeO2 dosage of 0.7 g L−1, and contact time of 5 h, the CeO2 nanoparticles achieved an adsorption capacity of 48 mg g−1 and 93% efficiency for Cr6+ ions, and 46 mg g−1 and 90% efficiency for Cd2+ ions. Interestingly, green CeO2 nanoparticles also demonstrated great potential for adsorbing radioactive metals such as uranium. Kashyap et al.103 reported that Citrus limon peel extract-based CeO2 nanoparticles achieved a uranium removal efficiency of 95% and a maximum adsorption capacity of 46 mg g−1 under the conditions of pH 4, uranium concentration of 0.1 mg L−1, dosage of 0.3 g L−1, and contact time of 1 h.
Because of their favorable adsorption characteristics, CeO2 nanoparticles were modified with carbon-based materials such as biochar to increase the adsorption efficiency of the resulting composites. For instance, Gupta et al.104 prepared a CeO2/biochar composite using Saccharum officinarum extract for methylene blue adsorption. The synthesis involved fabricating green CeO2via a precipitation method, followed by mixing the green cerium oxide with biochar derived from Areca nut shells, and then using a hydrothermal process to form the composite (Fig. 10a). The phytochemicals in the plant extracts and biochar scaffolds significantly influenced the surface chemistry of the composites. In the XPS C 1s spectra, the authors found primary peaks at 285 eV (C–H and C–C) and 286 eV (C–O). Moreover, FTIR analysis confirmed the presence of –OH, –NH–, C
C, C–C, and C–O stretching, as shown in Fig. 10b. The CeO2/biochar composite also showed a very high efficiency of 98%, corresponding to the adsorption capacity of 492 mg g−1 (Fig. 10c). The researchers ascribed its excellent removal efficiency to the combination of physisorption and chemisorption mechanisms (Fig. 10d). Besides, they verified that the CeO2/biochar possessed a pHpzc of 6.3, implying that its surface was negatively charged at an optimum pH of 8. The composite surface attracted methylene blue through electrostatic attractions. Biochar contains functional groups, including –COOH and –OH, which act as active sites for the adsorption of methylene blue via electrostatic interactions, H-bonding, and pore-filling. The aromatic ring structure of the biochar on CeO2/biochar enables π–π stacking interactions, which enhance the adsorption of methylene blue. These interactions served to increase the adsorption capacity of CeO2/biochar for methylene blue.
![]() | ||
| Fig. 10 (a) Procedure for the synthesis of CeO2/biochar nanocomposite using sugarcane extract and Areca nut shell. (b) FTIR spectrum of composite. (c) Adsorption capacity of composite for methylene blue versus time. (d) Proposed adsorption mechanism. Reproduced with permission from ref. 104 Copyright (2024), Elsevier. | ||
In another study, Norbert et al.73 demonstrated that the copper content in green Cu-doped CeO2 nanocomposites played a key role in their Congo Red removal efficiency. Under the optimum conditions (pH 6.5, dye concentration of 10 mg L−1, nanoparticle dosage of 0.2 g L−1, and contact time of 35 min), the nanocomposite containing 15% copper exhibited 98% removal efficiency. However, the ceria nanoparticles and nanocomposites with 5% and 10% Cu content exhibited lower adsorption efficiencies of 15%, 50%, and 93%, respectively. These authors assigned this effect to the surface changes due to the formation of the CuO phase and the enhancement in oxygen vacancies that occurs with the reduction of Ce4+ to Ce3+. This observation was corroborated by the XPS study because the Oβ peak (530 eV) was found to have a greater intensity with an increase in copper doping, indicating the enhancement in oxygen vacancies. The kinetic analysis revealed that the pseudo-second-order model provided the best fit for Cu-doped CeO2, which suggested that chemisorption controlled the adsorption process. The adsorption isotherm analysis demonstrated that the Langmuir model accurately described its adsorption behavior, suggesting a monolayer adsorption mechanism. Moreover, the values of 0 < RL < 1 also confirmed the presence of favorable adsorption. Copper doping enhanced the capacity of CeO2 to adsorb and bind dye molecules, as shown by the higher Qmax values (0.9 to 2 mg g−1), where the 15% copper nanocomposite showed greater adsorption activity than pure CeO2.
The biocompound-mediated synthesis of ceria nanoparticles often lacks surface functionalization, limiting the interaction of CeO2 nanoparticles with specific pollutants or biological targets for adsorption. Cerium oxide nanoparticle-based composites only showed the integration of ceria with biochar and metal components.73,104 Meanwhile, the use of other potential reinforcement agents, such as activated carbon (AC), metal–organic frameworks (MOFs), and MXenes, can boost their adsorption performance. AC has a high surface area and many functional groups, which assist in removing contaminants.115 MOFs have large porosity and adsorption sites; therefore, MOFs may selectively interact with pollutants.116,117 In the case of MXenes, these materials have hydrophilic property, high conductivity, and active sites on their surface, which allows strong adsorption properties.118 Thus, incorporating these materials in ceria-based composites could greatly increase their adsorption capacity and broaden their potential uses.
These nanoparticles show lower colloidal stability than their chemically synthesized counterparts, increasing their agglomeration during storage. For example, the green CeO2 nanoparticles synthesized using Acacia concinna plant extract had an efficiency of 189 mg g−1 for reactive blue but the stability test showed a loss of efficiency from about 98% to 73% over 5 cycles.31 Conversely, the CeO2/Fe3O4/g-C3N4 composite synthesized using a chemical method retained approximately 99% adsorption efficiency for Rose Bengal after five cycles.119 Future approaches could investigate surface modification, such as the functionalization of cerium oxide nanoparticles with amine, carboxyl or hydroxyl groups, to amplify their selectivity and reusability. Therefore, addressing these limitations will enhance the viability of green-synthesized CeO2 nanoparticles for wastewater treatment and environmental remediation.
Although the adsorption, photocatalytic, and biological activities of biosynthesized CeO2 nanoparticles have been widely investigated, the precise mechanisms underlying the interactions of green ceria with pollutants and biological systems are poorly understood. For instance, the contribution of oxygen vacancies to their excellent abilities for dye and heavy metal adsorption has been widely researched, but the mechanism by which certain plant-derived capping agents manipulate these defects is not clear. Thus, these mechanisms should be further investigated using in situ XPS and high-resolution TEM, both of which will provide greater insight into these mechanisms.
Moreover, the high oxygen storage capacity of ceria makes it an ideal candidate for the fabrication of electrocatalysts for hydrogen evolution reactions.126,127 Green-synthesized CeO2-based catalysts could replace expensive noble metals in hydrogen fuel production. Recent studies suggest that modifying biosynthesized CeO2 with transition metals or carbon-based materials (Cr, TiO2, and MoxC) can significantly elevate its catalytic activity, stability, and electron transfer efficiency.128–130 Therefore, green ceria nanoparticles may be promising materials for next-generation green hydrogen production.
Although CeO2 has been investigated for cancer treatment, the role of ceria responding to stimuli as targeted drug carriers has not been comprehensively investigated. In conventional drug delivery systems, premature drug release or low site specificity leads to a reduction in therapeutic efficacy and a higher rate of side effects. By functionalizing green CeO2 with pH- or temperature-sensitive polymers, intelligent nanocarriers responsive to the characteristics of the tumor microenvironment can be achieved as smart nanodrugs, allowing specific, on-demand drug release at the treated site.131–133 In addition, the implementation of other stimuli-responsive moieties, including enzyme- or redox-sensitive coatings, can also boost the specificity and therapeutic efficacy of ceria-based drug carriers.134,135 This offers the potential of a much more effective and sustainable alternative to conventional chemotherapy, a direction in which further research could transform the treatment of cancer.
| This journal is © The Royal Society of Chemistry 2025 |