Next Article in Journal
Seasonal Changes in the Acceleration–Speed Profile of Elite Soccer Players: A Longitudinal Study
Previous Article in Journal
A Multi-Segment Expanded Anchor for Landslide Emergency Management
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Investigation of Biofabricated Iron Oxide Nanoparticles for Antimicrobial and Anticancer Efficiencies

by
Nilavukkarasi Mohandoss
1,
Sangeetha Renganathan
2,
Vijayakumar Subramaniyan
1,*,
Punitha Nagarajan
1,
Vidhya Elavarasan
1,
Prathipkumar Subramaniyan
3 and
Sekar Vijayakumar
4,*
1
PG and Research Department of Botany, A.V.V.M. Sri Pushpam College, Affiliated to Bharathidasan University, Poondi 613503, India
2
PG and Research Department of Mathematics, A.V.V.M. Sri Pushpam College, Bharathidasan University, Poondi 613503, India
3
National Institute of Technology, Tiruchirappalli 620015, India
4
Marine College, Shandong University, Weihai 264209, China
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2022, 12(24), 12986; https://doi.org/10.3390/app122412986
Submission received: 26 November 2022 / Revised: 13 December 2022 / Accepted: 15 December 2022 / Published: 17 December 2022

Abstract

:
Capparis zeylanica leaf extract was employed in this work to create iron oxide nanoparticles (α-Fe2O3) using anhydrous ferric chloride. The UV spectrum, XRD, FT-IR, and FE-SEM with EDX methods were used to characterize the fabricated nanoparticles. The iron oxide nanoparticles obtained were spherical in form, with an average crystallite size of 28.17 nm determined by XRD. The agar well diffusion method was used to assess the antimicrobial activity of the α-Fe2O3 nanoparticles created in this study against pathogenic organisms, Gram-negative bacteria (Escherichia coli and Pseudomonas aeroginosa), Gram-positive bacteria (Staphylococcus aureus and Streptococcus pyogenes), and fungi (Candida albicans and Aspergillus niger). Among the pathogens tested, S. pyogenes had the highest zones of inhibition (25 ± 1.26 mm), followed by S. aureus (23 ± 0.8 mm), E. coli (23 ± 2.46 mm), P. aeroginosa (22 ± 1.86 mm), C. albicans (19 ± 2.34 mm) and A. niger (17 ± 3.2 mm). The substance was further tested for anticancer activity against A549 (lung cancer) cells using the MTT assay. The cytotoxic reaction was found to be concentration-dependent. The present study, therefore, came to the conclusion that the bio-effectiveness of the manufactured α-Fe2O3 nanoparticles may result in applications in biomedical domains.

1. Introduction

Nanotechnology has developed a reputation in nanoscience during the past twenty years. Researchers and scientists are actively working to create diverse nanoparticles from innovative materials for various uses [1]. One of the most significant transition metals for biological purposes is iron oxide. Hematite (α-Fe2O3) is the most stable polymorph of iron oxide and is ideal for biomedical applications due to its adaptable magnetic and optical characteristics, great chemical stability, and biocompatibility. There are a number of ways to make iron oxide nanoparticles, including mechanochemical methods, wire electrical explosion, electrochemical methods, and green synthesis.
Plants are more favoured than microorganisms as biological resources because of their low cost and simple accessibility in nature. In addition to lowering the metal salt, the biomolecules used in the synthesis also functionalize the nanoparticles’ surfaces, leading to synergistic benefits in applications such as cancer therapy or the development of antimicrobial agents [2]. A promising alternative method for creating nanoparticles is through green synthesis. Compared to the currently available, traditional physical and chemical approaches, green synthesis provides a number of benefits [3]. Iron oxides are one of the most biocompatible metal oxide nanoparticles due to their exceptional minuscule physical properties, including superparamagnetism, firmness in liquid solutions, low susceptibility to oxidation, long blood half-lives, and flexible surface chemistry. They have a wide range of applications in environmental regulation, including as antibacterial drugs, in the adsorption of dyes, in food-related processes, and in the biomedical sector (drug delivery, magnetic cells, immunoassays, and hyperthermia treatment of cancer) [4,5,6,7,8,9,10,11]. In the current work, anhydrous ferric chloride was converted into α-Fe2O3 nanoparticles using a Capparis zeylanica leaf extract and the nanoparticles were examined for antimicrobial and antiproliferative activities.

2. Materials and Methods

2.1. Materials

Fecl3.6H2O of analytical grade was the sole substance employed in the manufacture of the hematite nanoparticles. We bought the bacterial culture medium from HiMedia (Mumbai, India). Sigma Aldrich in the USA provided the ethyl acetate, antibiotic solution, and Dulbecco’s Modified Eagle’s Medium (DMEM).

2.2. Biosynthesis of α-Fe2O3 Nanoparticles

The procedure recommended in prior research [12] was slightly modified to prepare the C. zeylanica leaf extract. C. zeylanica leaf extract was added to 50 mL of a 0.1 M aqueous solution of ferric chloride hexahydrate along with leaf extract. The orange-coloured ferric chloride hexahydrate solution changed to a brownish-black colour, precipitating a black substance. After being sonicated for 30 min, the mixture was filtered, rinsed with distilled water, and dried. To produce iron oxide nanoparticles, the dried precipitate was heated for three hours at 600 °C in a muffle furnace.

2.3. Characterization of α-Fe2O3 Nanoparticles

The X-ray diffraction (XRD) method was used to study the crystallinity of α-Fe2O3 nanoparticles. Using a UV–Vis NIR double-beam spectrophotometer, the optical properties of α-Fe2O3 was examined at different wavelengths between 300 and 700 nm. The substituents of the nanoparticles were examined using Fourier-transform infrared (FT-IR) spectroscopy (Perkin Elmer RX-I Spectrophotometer). A field emission scanning electron microscope (FE-SEM) was employed to examine the morphology of the α-Fe2O3 nanoparticles. The energy-dispersion X-ray (EDX) method was used to analyse the composition of the α-Fe2O3 nanoparticles.

2.4. Antimicrobial Activity

The antibacterial activity of the artificial α-Fe2O3 nanoparticles was investigated using ethyl acetate-based leaf extracts from C. zeylanica. The pathogenic organisms were collected from the Microbial Type Culture Collection (MTCC), Chandigarh, India, and included Streptococcus pyogenes, Staphylococcus aureus, Escherichia coli, Pseudomonas aeruginosa, Candida albicans, and Aspergillus niger. The aforementioned cultures were then cultivated and kept alive at 35 °C for bacterial species and 30 °C for fungal species on Mullar-Hinton agar (MHA) and Sabouraud Dextrose Agar (SDA) media, respectively. In MHA and SDA plates, a 6 mm well was pounded. A micropipette was used to inject 20 µL of sample into each plate’s well. Gentamycin (50 µg) served as a positive control for bacteria and fungi. The bacterial and fungal cultures were incubated for 24 h at 30 °C and 72 h at 37 °C, correspondingly. Each experiment was executed three times, with measurements taken in millimetres.

2.5. Anticancer Activity

The National Center for Cell Science (NCCS), Pune, India, provided the lung cancer cell line A549. The cells were cultured in DMEM media supplemented with 10% FBS, Penicillium (100 g/mL), streptomycin (100 g/mL), and amphotericin B (5 g/mL) and incubated overnight. Utilizing the established 3-(4,5-dimethylthiazole-2-yl)-2, 5-diphenyltetrazolium bromide (MTT) tests as described by Chelliah and Oh [13], the cytotoxic impact of synthesized α-Fe2O3 nanoparticles on the A549 cell line was assessed. A 96 well plate with 2 × 104 cells per well and a volume of 100 µL was used for cell incubation. After 24 h, the cell layer was treated with various doses (1.45, 3.9, 7.8, 15.6, 31.2, 62.5, 125, 250, 500, and 1000 g/mL) of Fe2O3 nanoparticles. Cells were incubated at 37 °C for 4 h after addition of 50 µL of MTT (1 mg/mL) to each well. The development of Formosan crystals was then disrupted in 200 µL of dimethyl sulphoxide after the medium had been withdrawn. A microplate reader was used to measure the absorption over the course of 15 min at a wavelength of 540 nm

2.6. Statistical Analysis

Each test was run in triplicate. One-way variant analysis was performed, and the results are presented as means ± standard errors.

3. Result and Discussion

3.1. UV–Visible Spectrum of α-Fe2O3 Nanoparticles

The spectrum of the α-Fe2O3 nanoparticles’ diffuse reflectance in the UV–visible range is displayed in Figure 1a. The prominent absorption peak at 354 nm corresponds to α-Fe2O3 nanoparticles [14].

3.2. XRD of α-Fe2O3 Nanoparticles

Figure 1b shows the XRD pattern of the α-Fe2O3 nanoparticles. The (012), (104), (110), (113), (024), (116), (018), (214), and (300) planes were, respectively, assigned to the XRD peaks at 24.76, 33.52, 35.12, 41.67, 49.79, 54.02, 57.22, 62.52, and 63.82 2h. The wide, sharp peaks further demonstrated the well-crystallized hexagonal structure of the α-Fe2O3 nanoparticles (corresponding to JCPDS file no. 33-0664).

3.3. FTIR Spectra of α-Fe2O3

The FTIR spectra of α-Fe2O3 are shown in Figure 1c. The synthetic α-Fe2O3 nanoparticles’ infrared spectra (FTIR), which were recorded in the 400–4000 cm−1 wavenumber region, were used to identify the compound’s chemical bonds and functional groups. The existence of an absorbed water molecule was indicated by the wide band at 3407 cm−1, which was attributed to the O-H stretching vibration [15]. The peaks at 3216 cm−1 in the fluoride-treated ferrous oxide-hydroxide nanoparticles were somewhat displaced to lower frequencies, which may have been the result of fluoride adsorption, ion exchange, or both [16]. The bending vibrations of the C-H bond in the organic compounds were thought to be responsible for the absorption peaks at 2967 cm−1 [17]. These adsorption peaks at 2076 cm−1 for the N=C=S (iso-thiocyanate) group provided evidence that proteins and other bioactive substances were present on the surface of the biosynthesized iron nanoparticles [18]. A small peak at 1028 cm−1 was caused by the stretching vibration of C-O H [19]. The FeO6 octahedron was present in the samples, as seen by the prominent absorption bands at 548 cm−1, which were attributed to the stretching of the Fe-O bond [20].

3.4. FE-SM with EDAX

In Figure 1d, pictures of the synthesized α-Fe2O3 nanoparticles obtained using FE-SEM are displayed. The α-Fe2O3 nanoparticles in the FE-SEM image are distributed uniformly and have a spherical shape. The α-Fe2O3 nanoparticles had an average particle size of 28.17 nm. The fabrication of α-Fe2O3 nanoparticles is depicted in Figure 1e as a histogram of the particle size distribution. Figure 1f displays the EDAX pattern. The components that made up the resulting α-Fe2O3 nanoparticles were iron (Fe) and oxygen (O).

3.5. Antimicrobial Activity

The substance was tested for its antimicrobial effectiveness against harmful organisms. Streptococcus pyogenes exhibited the largest inhibitory zones (25 mm), followed by Staphylococcus aureus (23 mm), Escherichia coli (23 mm), Pseudomonas aeroginosa (22 mm), Candida albicans (19 mm), and Aspergillus niger (17 mm). With the use of the agar well diffusion technique, a considerable zone of inhibition (ZOI) was discovered, as shown in Figure 2a. Table 1 shows that the rate of microbiological growth decreased as the α-Fe2O3 nanoparticle concentration was increased. The results for the synthesized nanoparticles’ antimicrobial efficacy against our chosen microorganisms showed that they possessed strong antimicrobial activity.
Similarly, Leisha et al. [21] demonstrated that biosynthesized iron-oxide nanoparticles can inhibit the growth of Staphylococcus aureus, Escherichia coli, P. aeruginosa, and Streptococcus mutans and have potential bactericidal activity. The strong antimicrobial activity of the synthesized nanoparticles was likely brought about by the presence of the plant extract on their surfaces. Even though these were extremely positive results, additional research is required to determine whether iron-oxide nanoparticles are a viable alternative to silver nanoparticles for the treatment of bacterial infections. Although the precise mechanism of the antimicrobial action of the α-Fe3O4 nanoparticles that causes damage to bacterial proteins and DNA has not yet been identified, it was speculated that it may involve oxidative stress brought about by reactive oxygen species, such superoxide radicals (O2-), singlet oxygen (1O2), hydroxyl radicals (-OH), or hydrogen peroxide (H2O2). The process that produces ROS can readily diffuse into the cytoplasm of the cell, causing ROS-induced release in the mitochondria and causing death [22].

3.6. Anticancer Activity

The α-Fe2O3 nanoparticles were tested for cytotoxicity against human A549 cancer cells at a range of doses (1–100 mg/mL). A phase-contrast inverted microscope was used to examine the cell growth. After receiving treatment with α-Fe2O3 nanoparticles, A549 cells shrank and their nuclei became fractured and compacted, as shown in the microscopy images in Figure 2b. At lower hematite concentrations (1–10 mg/mL), significant morphological alterations were not seen. According to our research, the substance has a concentration-dependent cytotoxic effect on A549 lung cancer cells (Figure 2c). At a concentration of 15.6 µg/mL, which is regarded as the IC50 value, α-Fe2O3 nanoparticles demonstrated 50% cell killing. According to the designated activity levels of the US National Cancer Institute’s Center For Cancer Research, the produced nanoparticles are moderately active [23].

4. Conclusions

Biofabricated α-Fe2O3 nanoparticles were successfully synthesized using C. zeylanica leaf extract. The conformations of the nanoparticles were characterized using various techniques. The α-Fe2O3 nanoparticles possess admirable antimicrobial and antiproliferative activities.

Author Contributions

N.M., Methodology; S.R. and P.S., Data curation; P.N., Formal analysis; V.E., Resources; V.S., Writing—Original draft; S.V., Writing—Review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by DST-FIST (SR/FST/College-222/2014) and DBT-STAR (HRD-11011/18/2022-HRD-DBT).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

On behalf of all the authors, the corresponding author states that our data are available upon reasonable request.

Acknowledgments

We express our gratitude to the management of A.V.V.M. Sri Pushpam College (Autonomous), Poondi, for providing us with the necessary support and research facilities to complete this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yusefi, M.; Shameli, K.; Ali, R.R.; Pang, S.W.; Teow, S.Y. Evaluating Anticancer Activity of Plant-Mediated Synthesized Iron Oxide Nanoparticles Using Punica Granatum Fruit Peel Extract. J. Mol. Struct. 2020, 1204, 127539. [Google Scholar] [CrossRef]
  2. Yoonus, J.; Resmi, R.; Beena, B. Evaluation of antibacterial and anticancer activity of green synthesized iron oxide (a-Fe2O3) nanoparticles. Mater. Today Proc. 2021, 46, 2969–2974. [Google Scholar] [CrossRef]
  3. Ahmad, W.; Kumar, J.K.; Amjad, M. Euphorbia herita leaf extract as a reducing agent in a facile green synthesis of iron oxide nanoparticles and antimicrobial activity evaluation. Inorg. Nano-Met. Chem. 2020, 51, 1147–1154. [Google Scholar] [CrossRef]
  4. Nilavukkarasi, M.; Vijayakumar, S.; Prathip Kumar, S. Biological synthesis and characterization of silver nanoparticles with Capparis zeylanica L. leaf extract for potent antimicrobial and anti proliferation efficiency. Mater. Sci. Energy Technol. 2020, 3, 371–376. [Google Scholar] [CrossRef]
  5. Zununi Vahed, S.; Fathi, N.; Samiei, M.; Maleki Dizaj, S.; Sharifi, S. Targeted Cancer Drug Delivery with Aptamer-Functionalized Polymeric Nanoparticles. J. Drug Target. 2019, 27, 292–299. [Google Scholar] [CrossRef] [PubMed]
  6. Vasantharaj, S.; Sathiyavimal, S.; Senthilkumar, P.; LewisOscar, F.; Pugazhendhi, A. Biosynthesis of Iron Oxide Nanoparticles Using Leaf Extract of Ruellia Tuberosa: Antimicrobial Properties and Their Applications in Photocatalytic Degradation. J. Photochem. Photobiol. B Biol. 2019, 192, 74–82. [Google Scholar] [CrossRef]
  7. Vinci, G.; Rapa, M. Noble Metal Nanoparticles Applications: Recent Trends in Food Control. J. Bioeng. 2019, 6, 10. [Google Scholar] [CrossRef] [Green Version]
  8. Yusof, H.M.; Mohamad, R.; Zaidan, U.H. Microbial Synthesis of Zinc Oxide Nanoparticles and Their Potential Application as an Antimicrobial Agent and a Feed Supplement in Animal Industry: A Review. J. Anim. Sci. Biotechnol. 2019, 10, 57. [Google Scholar] [CrossRef] [Green Version]
  9. Borah, D.; Das, N.; Das, N.; Bhattacharjee, A.; Sarmah, P.; Ghosh, K. Alga Mediated Facile green Synthesis of Silver Nanoparticles: Photophysical, Catalytic and Antibacterial Activity. Appl. Organomet. Chem. 2020, 34, e5597. [Google Scholar] [CrossRef]
  10. Hernández, A.A.; Aguirre, G.; Cariño-Cortés, R.; MendozaHuizar, L.H.; Jiménez-Alvarado, R. Iron Oxide Nanoparticles: Synthesis, Functionalization, and Applications in Diagnosis and Treatment of Cancer. Chem. Pap. 2020, 74, 3809–3824. [Google Scholar] [CrossRef]
  11. Vahidi, H.; Barabadi, H.; Saravanan, M. Emerging Selenium Nanoparticles to Combat Cancer: A Systematic Review. J. Clust. Sci. 2020, 31, 301–309. [Google Scholar] [CrossRef]
  12. Mughal, B.; Zaidi, S.Z.J.; Zhang, X.; Hassan, S.U. Biogenic Nanoparticles: Synthesis, Characterisation and Applications. Appl. Sci. 2021, 11, 2598. [Google Scholar] [CrossRef]
  13. Parthasarathy, V.; Selvi, J.; Senthil Kumar, K.P. Evaluation of mechanical, optical and thermal properties of PVA nanocomposites embedded with Fe2O3 nanofillers and the investigation of their thermal decomposition characteristics under non-isothermal heating condition. Polym. Bull. 2021, 78, 2191–2210. [Google Scholar] [CrossRef]
  14. Chelliah, R.; Oh, D.H. Screening for anticancer activity: 3-(4,5-Dimethylthiazol-2-yl)-2,5-diphenyltetrazolium (MTT) assay. In Methods in Actinobacteriology; Springer Protocols Handbooks; Dharumadurai, D., Ed.; Humana: New York, NY, USA, 2022; pp. 423–425. [Google Scholar] [CrossRef]
  15. Veeramani, S.; Shanmugam, K.; Sahadevan, R. Folate Targeted Galactomannan Coated Iron Oxide Nanoparticles as a Nanocarrier for Targeted Drug Delivery of Capecitabine. Int. J. Med. Nano Res. 2018, 5, 25. [Google Scholar] [CrossRef] [Green Version]
  16. Raul, P.K.; Devi, R.R.; Umlong, I.M.; Banerjee, S.; Singh, L.; Purkait, M. Removal of Fluoride from Water Using Iron Oxide-Hydroxide Nanoparticles. J. Nanosci. Nanotechnol. 2012, 12, 3922–3930. [Google Scholar] [CrossRef]
  17. Si, J.C.; Xing, Y.; Peng, M.L.; Zhang, C.; Buske, N.; Chenbc, C.; Cui, Y.L. Solvothermal synthesis of tunable iron oxide nanorods and their transfer from organic phase to water phase. CrystEngComm 2014, 16, 512–516. [Google Scholar] [CrossRef]
  18. Gouda, A.R.; Sidkey, N.M.; Shawky, H.A.; Yasser, A.A.H. Biosynthesis, Characterization and Antimicrobial Activity Of Iron Oxide Nanoparticles Synthesized By Fungi. Al-Azhar J. Pharm. Sci. 2020, 62, 164–179. [Google Scholar] [CrossRef]
  19. Hassan, S.S.M.; Shafy, H.I.A.; Mansour, S.M.M. Removal of pyrene and benzo (a) pyrene micropollutant from water via adsorption by green synthesized iron oxide nanoparticles. Adv. Nat. Sci. Nanosci. Nanotechnol. 2018, 9, 015006. [Google Scholar] [CrossRef]
  20. Mani, A.D.; Soibam, I. Effect of the Size of the Dopants on the Dielectric and Magnetic Properties of Bi (La,Gd) FeO3 Multiferroic Materials. J. Nano Res. 2018, 51, 61–68. [Google Scholar] [CrossRef]
  21. Armijo, L.M.; Wawrzyniec, S.J.; Kopciuch, M. Antibacterial activity of iron oxide, iron nitride, and tobramycin conjugated nanoparticles against Pseudomonas aeruginosa biofilms. J. Nanobiotechnol. 2020, 18, 35. [Google Scholar] [CrossRef] [Green Version]
  22. Rudramurthy, G.R.; Swamy, M.K.; Sinniah, U.R.; Ghasemzadeh, A. Nanoparticles: Alternatives Against Drug-Resistant Pathogenic Microbes. Molecules 2016, 21, 836. [Google Scholar] [CrossRef] [PubMed]
  23. Chothiphirat, A.; Nittayaboon, K.; Kanokwiroon, K.; Srisawat, T.; Navakanitworaku, R. Anticancer potential of fruit extracts from vatica diospyroides symington type SS and their effect on program cell death of cervical cancer cell lines. Sci. World J. 2019, 9, 5491904. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Bio-fabricated α-Fe2O3 nanoparticles: (a) UV spectrum, (b) XRD, (c) FT-IR, (d) FE-SEM, (e) histogram distribution, (f) EDAX.
Figure 1. Bio-fabricated α-Fe2O3 nanoparticles: (a) UV spectrum, (b) XRD, (c) FT-IR, (d) FE-SEM, (e) histogram distribution, (f) EDAX.
Applsci 12 12986 g001
Figure 2. Bio-fabricated α-Fe2O3 nanoparticles: (a) antimicrobial activity, (b) MTT assay, (c) anticancer activity.
Figure 2. Bio-fabricated α-Fe2O3 nanoparticles: (a) antimicrobial activity, (b) MTT assay, (c) anticancer activity.
Applsci 12 12986 g002
Table 1. Antimicrobial activity of biosynthesised α-Fe2O3 nanoparticles against microorganisms.
Table 1. Antimicrobial activity of biosynthesised α-Fe2O3 nanoparticles against microorganisms.
S. No.Name of the OrganismZone of Inhibition (µg/mL−1) a
PEFecl3.6H2Oα-Fe2O3Positive ControlNegative Control
1.Streptococcus pyogenes9 ± 1.3714 ± 1.2625 ± 2.5620 ± 1.320
2.Staphylococcus aureus7 ± 1.24 12 ± 1.4323 ± 2.2317 ± 1.550
4.Escherichia coli6 ± 0.767 ± 0.2023 ± 1.3413 ± 1.220
3.Pseudomonas aeruginosa8 ± 1.2310 ± 1.2122 ± 1.2315 ± 1.350
5.Candida albicans7 ± 1.1911 ± 0.1719 ± 0.9811 ± 1.120
6.Aspergillus niger5 ± 0.996 ± 1.2717 ± 1.629 ± 0.750
a Mean values of three triplicates ± standard deviation. PE—plant extract, Fecl3.6H2O—ferric chloride hexahydrate, α-Fe2O3—biosynthesized iron oxide nanoparticles, Control—ethyl acetate.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mohandoss, N.; Renganathan, S.; Subramaniyan, V.; Nagarajan, P.; Elavarasan, V.; Subramaniyan, P.; Vijayakumar, S. Investigation of Biofabricated Iron Oxide Nanoparticles for Antimicrobial and Anticancer Efficiencies. Appl. Sci. 2022, 12, 12986. https://doi.org/10.3390/app122412986

AMA Style

Mohandoss N, Renganathan S, Subramaniyan V, Nagarajan P, Elavarasan V, Subramaniyan P, Vijayakumar S. Investigation of Biofabricated Iron Oxide Nanoparticles for Antimicrobial and Anticancer Efficiencies. Applied Sciences. 2022; 12(24):12986. https://doi.org/10.3390/app122412986

Chicago/Turabian Style

Mohandoss, Nilavukkarasi, Sangeetha Renganathan, Vijayakumar Subramaniyan, Punitha Nagarajan, Vidhya Elavarasan, Prathipkumar Subramaniyan, and Sekar Vijayakumar. 2022. "Investigation of Biofabricated Iron Oxide Nanoparticles for Antimicrobial and Anticancer Efficiencies" Applied Sciences 12, no. 24: 12986. https://doi.org/10.3390/app122412986

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