Open access peer-reviewed chapter

The Potential Application of Nanoparticles on Grains during Storage: Part 2 – An Overview of Inhibition against Fungi and Mycotoxin Biosynthesis

Written By

Daniel Nsengumuremyi, Parise Adadi, Gavers K. Oppong, Nadezhda V. Barakova and Elena F. Krivoshapkina

Reviewed: 17 June 2020 Published: 15 September 2020

DOI: 10.5772/intechopen.93213

From the Edited Volume

Mycotoxins and Food Safety

Edited by Suna Sabuncuoğlu

Chapter metrics overview

831 Chapter Downloads

View Full Metrics

Abstract

Mycotoxins are secondary metabolites synthesized by filamentous fungi. They are common food contaminants that cause mycotoxicosis in humans and animals. Due to the severity of health risk pose by these mycotoxins, many countries have enacted strict measures to curb this menace. One promising measure is the use of nanoparticles. Herein, we present an overview of the application of titanium dioxide, chitosan, ultradisperse humic sapropel suspension, and carbon-based nanoparticles, a novel and innovative method of reducing mycotoxin production and the subsequent contamination of grains. All nanoparticles considered enhanced cell permeability by disrupting the membrane, resulting in the outflow of cellular materials. However, concentration, volume, type, and illumination (sunlight) influenced the fungicidal potential of NPs.

Keywords

  • filamentous fungi
  • mycotoxins
  • nanoparticles
  • fungicide
  • reactive oxygen species

1. Introduction

Microorganisms, including fungi, contaminate grains during storage. These fungi do not only reduce grain quality, but also produce mycotoxins which pose health risks to consumers [1, 2]. According to Kady et al. [3], Aspergillus, Fusarium, Penicillium, and Rhizopus are the most common genera in barley, wheat, maize, and sorghum. These grains serve as staple food worldwide. Nowadays, nanotechnology is advancing in many fields, namely biotechnology, analytical chemistry, agriculture, and others. However, its application in crop protection is still in its early stages [4, 5].

The biocidal activity of nanoparticles is well documented. Herein, we proposed the utilization of nanoparticles to inhibit fungal growth and the production/synthesis of mycotoxins. Therefore, the second part of this chapter aims to discuss other promising nanoparticles (titanium dioxide nanoparticles, chitosan nanoparticles, ultradisperse humic sapropel suspension (UDHSS) nanoparticles, and carbon-based nanoparticles/nanomaterials) of interest which could be applied during grain storage. The toxicological aspects, as well as the proposed modes of application are discussed.

Advertisement

2. Titanium dioxide nanoparticles

Titanium dioxide (TiO2) nanoparticles (TiO2-NPs), or ultrafine TiO2, are particles of TiO2 with diameters 1–100 nm. The TiO2-NPs activity is exciting to researchers because of its specific characteristics which include; size, shape, crystal structure, surface stability among others [6]. They are among top five NPs used in consumer items such as cosmetics, food products, paints, and medicines [7]. TiO2 received USFDA approval hence regarded as safe. It is widely used as food colorant in candies, sweets, chewing gums, etc. Anatase (used in printing inks and photocatalysts), rutile (used in colorants and sunscreens), and brookite are the three primary forms of TiO2-NPs [8, 9, 10, 11, 12]. In 1985, Matsunaga et al. [13] first documented the antimicrobial activity of TiO2. They observed that microbial cells were dead when exposed to a TiO2-Pt catalyst illuminated with UV light.

The biocidal activity of TiO2 has been reported [14, 15, 16, 17, 18, 19]. Table 1 shows the fungicidal activity of TiO2-NPs against fungi species known to contaminate grains with the mycotoxins they synthesize.

OrganismReference
C. albicans, S. cerevisiae[31]
A. niger AS3315[32]
F. verticillioides[33]
A. niger spores[34]
A. niger, S. cerevisiae[35]
F. oxysporum f. sp. lycopersici[36]
C. albicans ATCC 10231, F. solani ATCC 36031[37]
C. albicans[27]
C. famata[38]
C. vini, Hansenula anomala CCY-138-30[39]
Cladobotryum varium, Trichoderma harzianum, Spicellum roseum[40]
Cladosporium cladospoiroides, Epicoccum nigrum, F. mucor, Penicillium oxalicum, Trichoderma asperellum, Pestaotiopsis maculans[41]
Diaporthe actinidae[25]
Erysiphe cichoracearum, Peronophythora litchii[42]
Molds and yeasts (not specified)[43]
Fusarium spp. (equisetii, oxypartan, anthophilum, verticillioides, solani)[44, 45]
P. citrinum[46, 47]
P. expansum[26]
S. cerevisiae[13, 48]

Table 1.

Fungicidal activities of TiO2-NPs on mycotoxins-producing fungi.

Modified with permission from Ref 4498160008350.

TiO2-NPs have been widely applied as antimicrobial agents in recent years due to their unique properties such as resistance to high temperatures, low solubility, high surface area, cost-effectiveness, hydrophilicity, and strong oxidizing properties [20].

TiO2-incorporated polyethylene (PE) film inhibited growth of E. coli and S. aureus. UV light significantly enhanced the biocidal activity within 60 minutes of illumination [20]. Several studies [21, 22, 23, 24, 25, 26] have documented the biocidal efficacy of TiO2 against E. coli, S. aureus, P. aeruginosa, and P. expansum.

The photocatalytic oxidation of surfaces coated with TiO2 and ultraviolet A (UVA) was effective against E. coli, P. aeruginosa, S. aureus, and E. faecium than the control [27]. A collaborated research [28] assessed the biocidal activity of the crude and annealed TiO2-NPs. The results revealed that doped Ag-TiO2 (7%) NPs killed 100%, 95%, and 96% of P. aeruginosa, S. aureus, and E. coli, respectively, at 40 mg/30 mL.

Assessing ecotoxicity of TiO2-NPs against bioluminescent bacterium (Aliivibrio fischeri), algae (Pseudokirchneriella subcapitata, Scenedesmus subspicatus, and Chlorella vulgaris), protozoon (Tetrahymena pyriformis), water flea (Daphnia magna), and an aquatic macrophyte, Lemna minor [29] revealed these organisms showed significant behavioral and physiological changes when exposed to low TiO2-NP concentrations (0.1 and 0.05 μg/L), thus demonstrated the ability of TiO2-NPs to alter molecular pathways via which these organisms obtained vital nutrition for growth and synthesis of compounds (i.e., chlorophyll, etc.).

Maneerat and Hayata [26] tested the fungicidal activity of TiO2 photocatalysts against P. expansum in the form of TiO2 powder and TiO2 coated on a plastic film. Both TiO2-NPs suppressed the conidial germination and growth of the fungi. The quantity of TiO2-NPs added correlated with the fungicidal activity.

Nitrogen-doped TiO2 [TiO2 (N)] exhibited potent biocidal activity with regards to reducing the number of surviving organisms than carbon-doped TiO2 [TiO2 (C)]. Therefore, TiO2 (N) NPs can inactivate spores of B. anthracis (hazardous microorganism) under illumination by conventional light sources such as incandescent lamps [30].

2.1 Mechanistic action of TiO2-NPs antimicrobial activity

TiO2-NPs are the photocatalysts used to destroy unwanted organic compounds in the air, water, soil, and, more recently, in food [21].

Photocatalysis can be defined as the catalyst-driven acceleration of a light-induced reaction [49, 50, 51, 52]. Homogeneous and heterogeneous photocatalytic processes utilize metal complexes (transition metal complexes like iron, copper, chromium, etc.) and semiconducting materials such as TiO2, ZnO, SnO2, and CeO2 as catalysts. In the presence of light and heat, metal complexes become excited and form metal ion complexes, in contrast, semiconducting materials become excited due to the combination of electronic structures which is characterized by a filled valence band, empty conduction band, and light absorption properties, resulting in the generation of reactive oxygen species (ROS) or hydroxyl radicals. These hydroxyl radicals inflict damage to microbial cells [49, 50, 51, 53, 54, 55]. The subsequent hole in the valence band could further react with H2O in the grains or hydroxide ions adsorbed on the surface of TiO2-NPs to generate hydroxyl radicals (OH•), with electron in the conduction band reduce O2 to superoxide ions (O2) [21]. Gogniat and Dukan [56] demonstrated that DNA was denatured by hydroxyl radicals generated via the Fenton reaction resulting in cell death.

Electron paramagnetic resonance (EPR) spectroscopy study confirmed the photoproduction of hydroxyl radicals (OH•) from different TiO2. The efficiency of hydroxyl radical generation depends on the source/origin of TiO2 [57].

Cells are negatively charged [58] under optimum physiological condition due to heparan sulfate proteoglycans [59]. However, disease could trigger the cells to synthesize certain compounds which cause cell surface to become positively charge. Microbial cell could act as a hole for electron transfer between organism and its components [60]. The iron cluster on cell surface, in the periplasmic space, or inside the cell (proteins (such as ferritin)), could act as a precursor for iron-catalyzed Haber-Weiss reaction, which generates additional hydroxyl radicals in the presence of H2O2 and the superoxide ion [61].

Different treatments (photocatalysis, water, TIO2, UV-A) applied to elucidate the effects of lipid peroxidation on S. cerevisiae revealed high malondialdehyde (MDA) in TiO2 -treated subjects with 2 hours. The results demostrated that TiO2 was sufficient to damage membrane, thus interfered with permeability of the cell which led to the leakage of vital intracellular molecules (Figure 1) [48]. Similarly, Draper and Hadley [62] found photocatalysis-induced cell wall damage on S. cerevisiae [48]. This may decrease intracellular enzymatic activity as well as leaking of amino acids and NH4+, suggesting a drastic impact on proteins [63].

Figure 1.

Schematic illustration of the solar photocatalytic process for microbial cell inactivation in the presence of an aqueous suspension of TiO2. Modified with permission from ref 4498160008350 [72]. Contact between the cells and TiO2-NPs affects membrane permeability; however, this is reversible. The availability of more NPs could enhance the damage to cell wall, thus allowing leakage of small molecules such as ions. Damage at this stage may be irreversible, and this accompanies cell death. Higher molecular weight components such as proteins could further be leaked followed by protrusion of the cytoplasmic membrane into the surrounding medium through degraded areas of the peptidoglycan and lysis of the cell. Intracellular components are then degraded progressively especially from the point of contact with photocatalyst, followed by complete mineralization.

Cellular respiratory enzymes lost their activity after been exposed to irradiated TiO2 (0.5 mg/mL), and the kinetics correlated with the losses of cell viability. Furthermore, when glucose was used instead of succinate as the electron donor, similar effects were observed. From this outcome, Li et al. [78] proposed that ROS generated from an irradiated TiO2 surface, interacted with the polyunsaturated phospholipids in E. coli. Moreover, cell membrane structure was perforated due to lipid peroxidation creating a hole for more TiO2-NPs to pass into interior of the cell, thus rendering respiratory proteinsinactive and subsequent cell death.

A progressive decrease in esterase activity was observed after exposing S. cerevisiae to irradiated TiO2 [63]. Other researchers documented overexpression and inhibition (expressed at lower levels, including those encoding six cbb3-type cytochrome C oxidase subunits, an electron transfer flavoprotein, and two oxidoreductases) of genes associated with energy production and conversion processes. TiO2-NPs exerted a stimulating effect on the respiratory chain and the electron transfer mechanism of the microorganism [64, 65].

Likewise, Matsunaga et al. [13] observed that incubating TiO2/Pt NPs under metal halide lamp irradiation with E. coli, Ch. vulgaris, L. acidophilus, and S. cerevisiae inhibited cell respiration mechanisms and subsequent cell death. However, the results were not consistent as Ch. vulgaris had a thick cell wall mainly composed of polysaccharides and pectin hence, had comparative advantages (protection) over the other microbes.

Kubacka et al. [65] examined genome/proteome-wide expression profiles of P. aeruginosa PAO1 cells treated with TiO2-based nanocomposite films. An increase and decrease in the levels of 165 and 151 transcripts were respectively reported in cells with TiO2-coated Ethylene vinyl alcohol (EVOH) particles. Few proteins were detected at a statistically significant level (p ≤ 0.1) in cells treated with TiO2-coated EVOH particles compared to the control. TiO2-UV treatment significantly suppressed (from 5.4- to 15.1-fold) the expression levels of genes essential for cell wall. However, 14 genes encoding for lipid metabolism essential for cell membrane were over-expressed (from 5.6- to 23.0-fold), unexpectedly, 2 were expressed at a lower level (from 5.5- to 7.4-fold).

In vivo and in vitro studies confirm that hydroxyl radicals inflict damage (breakage) on DNA strands. The extent of damage was minimized when dimethyl sulfoxide, catalase, or mannitol were incorporated in the reaction mixture [66]. However, the findings [66] contradicts previous studies [21, 67]. Exposing either purine or pyrimidine bases to TiO2 and light from a 100-W Hg lamp resulted in the detection of NO3 and NH4+ ion. However, when native DNA and RNA molecules were subject to the same conditions, unknown peroxide species, along with phosphate and carbon dioxide, were detected, suggesting the breakage and mineralization of sugar-phosphate backbone of DNA and RNA molecules, respectively [68].

Kikuchi et al. [67] demonstrated the role of ROS on photocatalytic bactericidal activity. They utilized a porous polytetrafluoroethylene (PTFE) membrane in their system to physically separate the E. coli suspension from the TiO2 thin. The results showed an impressive photokilling capability of the system with and without (control) PTFE - which was attributed to the generated H2O2. A group [69] demonstrated the stimulating effect of TiO2-NPs on lipolytic activity in A. niger. The results showed that TiO2-NPs significantly increased lipase biosynthesis (more than 1.5 times) compared to the control experiment. Treatment with TiO2-NPs (size: 40 nm, concentration: 10 mg/L) in all culture media, enhanced lipolytic activity by 78.57% and 57.49% on the 4th and 5th day of cultivation, respectively. This finding reaffirms that smaller NPs can penetrate the cell membrane easily than bigger NPs, thus easily interact with molecular proteins, resulting in stimulating effects.

Gomes et al. [70] assessed the effects TiO2-NMs (NM103, NM104, and NM105) and bulk TiO2 against Enchytraeus crypticus with and without UV radiation. Microarray analysis revealed 10431 differentially expressed genes (DEGs) (p < 0.01) triggered as a result of exposure to TiO2-NMs under no-UV. All samples under UV exposure registered an up-regulation of several transcripts, including caspase apoptosis-related cysteine peptidases, a signature of apoptosis activation, whereas under darkness the apoptotic signaling pathway was inhibited, suggesting that the oxi-radicals generated during the photoactivation of TiO2 might substantially contribute to the apoptotic response and damage to the cell membrane. DNA damage was triggered after exposing samples to bulk/nano TiO2 [71]. However, the findings of Gomes et al. [70] contradicted the [71] as reported that TiO2-NMs_under no-UV impaired DNA repair, while bulk_TiO2 under no-UV activated DNA repair mechanisms, suggesting that size of the TiO2-NPs contributes to biocidal activity.

Advertisement

3. Chitosan nanoparticles

Chitin and chitosan have been widely used in the fabrication of polymer scaffolds [73]. Chitosan is a linear polysaccharide, a nontoxic biopolymer derived from the deacetylation of chitin, and used in many fields, including agriculture, medicine, and in vinification due to its biocidal potential. In agriculture, chitosan is used as biopesticide; in medicine, it is used to stop bleeding, wound healing, and as an antibacterial agent. Biodegradability, high permeability, nontoxic to humans, and cost-effectiveness are the features which make chitosan NPs unique. Chitosan and its derivatives have attracted considerable attention due to their biocidal activities [74, 75]. Several authors have reported the beneficial application of chitosan and its oligosaccharides which includes antitumor [76], neuroprotective [77], antimicrobial [78, 79, 80, 81, 82, 83, 84, 85], and anti-inflammatory [86] agents. Table 2 summarizes the fungicidal activities of chitosan against important agricultural microorganisms contaminating stored grains.

ReferenceSources of chitosan (CTS)Deacetylation (%)MicroorganismsConcentrationForm applied
[104]Not reported (industrially made)71.5A. niger, A. parasiticus0 (control), 0.1, 0.5, 1.0, 2.0, 3.0, 4.0, and 5.0 mg/mLSolutions
[105]Not reported (industrially made)75–85A. flavus IMI242687, C. cladosporioides IMI 274019, M. racemosus IMI 017313, P. aurantiogriseum IMI 297953, Byssochlamys spp. BF, Byssochlamys spp. GCB, Byssochlamys spp. SB, S. cerevisiae 28, S. cerevisiae 3085, S. cerevisiae SD, Z. bailii906, Z. bailii HP, S. exiguus 391, S. pombe, S. ludwigii0, 1, 5, 10 g/L (fungi) and 5 mL (for yeast)Solutions
[106]Not reported85, 81, and 82 for low-, medium-, and high-molecular weight chitosan respectivelyA. alternata, B. fabae, F. oxysporum, P. digitatum, P. debrianum, R. solani250, 500, 1000, 1500, 2000, 2500, 3000, 3500 and 4000 mg/LSolutions
[97]Not reported79S. exiguus, S. ludwigii, T. delbrueckii0.05%, 0.005%Solutions
[107]Not reportedNot reportedC. neoformans strain B3501Different concentration (0, 0.625, 1.25, 2.5, and 5 mg/mL) was employedBiofilm
[108]Industrially prepared chitosan95Psychrophilic, mesophilic, Pseudomonad, yeasts and moldsNot reportedCoating
[87]Crab shell~90P. expansum (blue mold)Various concentrations were applied for in vivo (0, 0.1, 0.5 and 1.0% (w/v)) in vitro (0, 0.001, 0.01, 0.1, 0.2, 0.3, 0.4, 0.5, 0.6, 0.7, 0.8, 0.9, and 1.0% (w/v)) experimentCoating
[109]Crab shell82F. oxysporum f. sp. radicis-lycopersici 127, Candida krusei VKPM Y-2594, E. coli ATCC 5945, P. aureofaciensVKPMB-7542, E. agglomerans VKPM B-7541, B. subtilis VKPM B-75400.1, 0.5, or 1.0 mg/mL for the fungicidal test while 0.9 mL of LMW or DDC-LMW chitosan was used for the bacteriostatic testSolutions
[110]Not reported75-85R. stolonifera, E. coli DH5α strain1.0 mLCoatings and solutions
[111]Not reported (industrially made)100 (MMW) and 97 (LMW)R. oryzae CECT 2340, A. alternata CECT 20560, A. niger CECT 20886 mLFilms and solutions
[112]Not reported (industrial made)85–89Molds and total flora isolated from strawberries (R. Stolonifer and B. cinerea)Final concentration before the spraying was 0.02%, w/v
[113]Not reported80%R. solani Kuhn, F. oxysporum (Schl.) f. sp. Cucumernum owen, C. cucumerinum Ell. Et Arthur, B. cinerea Pers., C. orbiculare (Berk. & Mont.) Arx, P. asparagi (sacc.a) Bubak, A. Kikuchiama Tanaka, P. italicum Wehmer, Fusarium oxysporum Schl. F. Sp. Uasinfectum (Atk.) Snyd. & Hans, V. ctahliae Kleb., R. sclani Kuhn., B. berengeriana de Not. f. Sp., Piricola (Nose) Koganezaea
et Sakuma, Sclerotinia sclerotiorum (Lib.) de Bary, Venturia nashicola Tanaka et Yamamoto, Gibberella zeae (Schw.) Petch and Phytophthora infestabs (Mont.)
20, 30, 50, 100, and 150 mg/LSolutions
[114]Not reported90A. niger0.1% or 1% (w/v)Coatings, films, and liquid
[115]Shrimp shellNot reportedA. alternata f. sp. lycopersici100–6400 μg/mLSolutions

Table 2.

Some selected studies on fungicidal activities of chitosan NPs.

Fungal decay on pear fruit was suppressed by the combination of chitosan, yeast antagonist Cryptococcus laurentii, and CaCl2. The results showed that mixture of chitosan at 0.5% and C. laurentii exerted greater effects compared to chitosan or C. laurentii alone. CaCl2 showed little antifungicidal activity; however, it combination with chitosan and C. laurentii led to an effective and stable reduction of fungal decay [87], thus minimize or eradicate the menace of postharvest losses. Anthracnose in papaya caused by Colletotrichum gloeosporioides was controlled by the combination of Burkholderia cepacia, chitosan (0.75%) and CaCl2 [88]. Postharvest blue, green, and grey molds affecting apple, oranges, and lemons were effectively controlled by mixing glycol chitosan (0.2%) with Candida saitoana [89, 90, 91]. Ag/chitosan-NPs showed significant antifungal activity against A. flavus, A. alternata, and Rsolani hence could be used during grain storage [92, 93]. The synergistic effect (fungicidal activities) of hybrid copper(II) chitosan NPs to inhibit the growth of Fgraminearum, Verticillium dahlia 57, and F. solani 169 was reported. In both cases, the NPs exerted an excellent efficacy in repressing the growth of fungi [94, 95]. Other authors reported that certain strains of A. flavus, Cladosporium cladosporioides, P. aurantiogriseum, and Torulaspora delbrueckii were resistant to chitosan at levels as high as 1% [7, 96]. The application of chitosan (0.025 and 0.05%) was effective against Saccharomycodes ludwigii and Saccharomyces exiguous. A rapid reduction in the number of yeast colonies was observed 2–4 min after application [97].

According to an earlier report, the effectiveness of the biocidal activity of chitosan depends on the molecular weight, degree of acetylation, and concentration [98, 99]. The application of NPs coated with polyethylene glycol (PEG) and natural garlic oil against Tribolium castaneum, a vital storage pest showed high efficiency over an extended period (8 months) due to the slow and persistent release of the active components [100]. The study highlighted the potential application of PEG-NPs as capsules to encapsulate various natural bioactive ingredients (i.e., oil from Azadirachta indica, extracts of Khaya anthotheca, alkaloid extracts of Piper guineense [101], etc.) for controll release and subsequent killing of microorganisms and pests during grain storage. Furthermore, [102, 103] extensively reviewed the literature on the biocidal activities of natural compounds (i.e., herbs, species, etc.) and its potential application in postharvest control.

3.1 Mechanistic action of chitosan nanoparticle antimicrobial activity

According to literature [116, 117], chitosan is composed of polycationic copolymers, with glucosamine and N-acetylglucosamine as axillary units, which contributes to its antimicrobial activity. The difference in environmental pH, pKa of chitosan and its derivatives creates an electric field for an electrostatic interaction between the polycationic structure and the anionic components of the cell (i.e., lipopolysaccharide and cell surface proteins), thus altering cell permeability [118, 119, 120, 121, 122, 123]. High pH enhance rapid protonation, which increase the positive charge density (polycationic activity) of chitosan. A positive correlation was established between charge density and the biocidal activity of quaternized chitosan [124, 125, 126, 127]. The inhibition potential of chitosan could be incapacitated when the charge density is reduced [120] due to changes of pH values. A similar outcome was reported by Qin et al. [128]. The antimicrobial mechanism was associated with the interaction of the negatively charged cell membranes and the cationic NH3+groups of the chitosan derivative, which increase membrane permeability resulting in lysis [129] and leakage of macromolecules killing the cells. A carboxyfluorescein (CF)-loaded liposome study showed the effectiveness of lower molecular weight (LMW) chitosan on the cell membrane. The results showed that 0.75 μg/μL of LMW chitosan triggered moderate (≈7%) leakage of carboxyfluorescein found in the large unilamellar vesicles [130]. Similarly, Ing et al. [131] reported that chitosan NPs prepared from different concentrations of LMW and high molecular weight (HMW) showed efficient inhibitory activity against C. albicans (MICLMW = 0.25–0.86 mg/mL and MICHMW = 0.6–1.0 mg/mL) and F. solani (MICLMW= 0.86–1.2 mg/mL and MICHMW = 0.5–1.2 mg/mL) compared to the solution form (MIC = 3 mg/mL for both MWs and species). The authors established a statistical linear relationship between MW and particle size/zeta potential, thus provided an avenue for the manipulation of physicochemical properties of NPs to maximize its ability to penetrate the cells, trigger leakage of intracellular component, eventually killing the fungi and extend safety of the grains.

Researchers [132, 133, 134, 135] proposed the fundamental mechanism contributing to interaction of negatively charged surface components of fungi and bacteria with the positively charged NH3+ groups of glucosamine (chitosan), which alters cell surface, and trigger leaking of intracellular substances, resulting in the impairment of vital physiological activities thus killing the microorganism. The inability of the second amino groups on N-acetylation of chitosan oligomers to donate positive chargeresult in the inhibition of its fungistatic activity [136]. Therefore, the contribution of NH3+ groups to biocidal activity cannot be ignored and should carefully be considered to maximize the effects.

The outer membrane (OM), inner core of lipopolysaccharide (LPS) molecules, and lipid components of Gram(−) bacteria are composed of anionic groups like phosphate and carboxyl, which contribute to the hydrophilic nature of the cell wall, thus creatin interaction of charges (electrostatic) with divalent cations. The OM protects Gram (−) bacteria cells from macromolecules and hydrophobic compounds (antibiotics and toxic drugs), giving Gram(−) bacteria a comparative advantage over Gram(+) bacteria. Therefore, breaching the integrity of the OM by chitosan could enhance its biocidal activity toward Gram(−) bacteria [137, 138]. On the other hand peptidoglycan (PG) and teichoic acid (TA) on the cell wall of Gram(+) bacteria have polyanionic group, which facilitates interaction via covalent bond with N-acetylmuramic acid in the PG layer, or via glycolipid- which links outer leaflet of the cytoplasmic membrane [139]. As documented by Kong et al. [120], the poly(glycerol phosphate) anion groups aid the structural stability of cell wall in addition to some membrane-bound enzymes.

LMW chitosan showed higher efficiency perforate/penetrate the microbial cell compared to HMW chitosan, which interacts with DNA to change the translation and transcription profile of genes. Chitosan binds to DNA with accurate precision, denying the organism of normal DNA transcription and mRNA synthesis, resulting in cell death [140, 141, 142].

A decrease in the induction of β-galactosidase was observed when yeast cells were exposed to chitosan. A concentration of 0.35 mg/mL chitosan reduced β-galactosidase activity by 32%. An increased in concentration (1.25 mg/mL) further led to the reduction of enzyme activity. The control experiment did not follow the trend. Likewise, the treated cells showed that chitosan greatly influenced protein biosynthesis in the yeast [130]. Previous work [143] documented cell sensitivity to chitosan, which altered the deletions of genes involved in sphingolipid (e.g., ipt1Δ, skn1Δ, lcb3Δ) and ergosterol (e.g., erg3Δ, erg5Δ) biosynthesis. In 1981, Hadwiger et al. [144] detected chitosan within plant cytoplasm and nucleus within 15 min after application, which indicate that chitosan can efficiently penetrate the thicker cell wall (the reason for its detection) and potentially interfered with DNA transcription and translation. This study suggests that chitosan can easily penetrate microbial cells since plants have a thicker cell wall than microbes.

Moreover, looking at the time factor (15 min), it is evident that chitosan can quickly interact with fungi and bacteria cellular DNA with subsequent inhibition of DNA transcription, as well as RNA and protein synthesis [140, 145, 146], leading to cell death. Chitosan triggered transcriptional responses when introduced to S. cerevisiae strain X2180-1A (MATa SUC2 mal gal2 CUP1). T-Profiler analysis showed cis-regulatory motifs apart from the environmental stress response correlated positively with expression in the chitosan-treated sample. Cin5p, Crz1p, and Rlm1p were the transcription factors associated with identified binding sites. Genes participating in cell wall organization, biogenesis, and signal transduction were also triggered in the treated sample compared to the control [134]. Some factors influencing the antimicrobial activity of chitosan is discussed above; however, Kong et al. [120] and Hosseinnejad and Jafari [147] published an excellent reviews on these factors.

Advertisement

4. Ultradisperse humic sapropel suspension (UDHSS) nanoparticles (UDHSS-NPs)

Sapropel is benthos found in fresh water, formed under anaerobic conditions from dead organic matter of anhydrobiotic microflora and microfauna. It is principally composed of nutrients (i.e., sugars, minerals, lipids, etc.) and organic compounds known as humic substances [148, 149, 150, 151]. Sapropels and sapropel extracts showed antibacterial and antifungal properties in previous studies hence could used as an alternate and novel biocidal agent during grain storage. The antimicrobial properties of sapropels is attributed to the presence of humic substances [152, 153, 154, 155, 156]. Sapropel has become a popular raw material for therapeutic applications, production of sorbents, organic fertilizers, and food supplements [157]. UDHSS-NPs are organic NPs which exhibits potent biocidal activities due to the presence of humic substances [148]. Fulvic acids (FAs), humic acids (HAs), mumie, and humin are the principal constituents of humic substances (HSs) in sapropels [158, 159, 160, 161], and are reportedly attribute to their biocidal properties. Many studies [152, 153, 154, 155] have illustrated the inhibitory effects of sapropel on bacteria (S. aureus, E. coli, etc.) and yeasts (Candida, etc.). A micrograph of UDHSS-NPs is shown in Figure 2 however, its characteristics were not included in the present study.

Figure 2.

UDHSS nanoparticles under a scanning electron microscope (SEM).

In a series of tests performed by Barakova et al. [148], experiments 2 and 3 exhibited most significant fungicidal effects on A. niger, a species which poses a greater threat to grain/food industries due to the potent mycotoxins it produce. A report showed that hematite NPs (hematite-HA complexes) significantly inhibited the growth and gene expression of P. putida KT2440. The bactericidal activities were ascribed to the oxidative stress induced by generated ROS. It was also shown that the physicochemical properties of the NPs (e.g., surface charge and size) influenced the efficacy of the hematite-HA complexes [162]. Therefore, modification of UDHSS-NPs could improve its biocidal properties.

A group of researchers [163] assessed the fungicidal activity of HAs and FAs extracted from soils on phytopathogenic fungal species (Physalospora piricola (P.P), Botrytis cinerea (B.C), Rhizoctonia cerealis (R.C), Fusarium graminearum (F.G), Phytophthora infestans (P.I), Sclerotinia sclerotiorum (S.S), Rhizoctonia solani (R.S), Cercospora arachidicola Hori (C.H), and Bipolaris maydis (B.M)). The results showed that HA exhibited above 30% and 50% inhibition against B.C, R.C, F.G, P.I, and P.P, respectively. The inhibition exerted by HA on all the species was higher compared to FA except for B.C. Correlation analysis further revealed that the inhibition rates of HAs decreased significantly with time (years) (p < 0.05) against most tested fungi except P.I., whereas FAs showed a negative correlation with cultivation years (p < 0.05) against most of the tested fungi except F.G. and S.S.

Recently, Ong et al. [164] documented that HAs (10 mg L−1 HA) altered enzyme activity in zebrafish embryo. Physicochemical properties such as size, zeta potential, and particle dissolution influenced their actions. It was further shown that coupling HAs with NPs enhanced the activity of the composite NPs. The addition of HAs reduced the hydrodynamic diameters of all examined NP suspensions except cadmium selenide (CdSe) NPs. Ezhkov and colleagues [165] developed NP-sapropel composite with particle size 45.0–180.0 nm and investigated its effects on treated albino mice. The results showed scarring of organ walls and shedding/exfoliation of the superficial epithelial cells. Further histological analysis of the oesophagus wall showed a significant thinning of the horny substance and the removal of the stratified epithelium of the mucous membranes in areas in contact with the NPs.

4.1 Mechanistic action of UDHSS-NPs antimicrobial activity

Several studies have described the biological activity of sapropel on enzymes, which confirms its antimicrobial activity. Details of these studies are discussed below in a quest to put forward a proposed mechanism by which UDHSS-NPs kill microorganisms. Environmental factors such as temperature, pH, oxygen, and moisture play a vital role in the mechanistic action of UDHSS-NPs. According to Perdue [166], HSs is complex mixture containing aliphatic, aromatic carboxyl and hydroxyl functional groups, which binds with microbial cells either on grains or in the environment (i.e., water, soil, etc.), thus alter the membrane structural intergrity and its functions. According to literature, the fungi cell walls share similarities with plant and bacterial and indeed with the extracellular matrix material of mammalian cells. The anionic surface, β1,4- and β1,3-linked polystarch forms a ribbon-like or helical (β1,3-glucan) structures which interacts with opposite charges. The cross linking of glycans of in eubacterial walls with peptides as well as phenolics and polysaccharides in plats promotes hydrogen bonding [167, 168]. Furthermore, the fungal cell wall is uniquely composed of mannoproteins, chitins, α- and β-linked glucans which serves many functions including; metabolism, ion exchange as well as providing cell rigidity and shape [169]. With the latter interacting with the HS. The interactions between HS and microbial cells depend on the lipophilicity and electric potential of the HS and cell [170], coupled with the size of the UDHSS-NPs. Microbial cells are composed of cations such as H+, Na+, K+, Li2+, Al3+, Ca2+, Cu2+, or Pb2+ which interact with UDHSS-NPs thus penetrate the cell. As documented by Lofts et al. [171], cation-HS interactions exert control on the reactivity of cation, including its bioavailability for further reaction. Studies have shown the effects of binding metals with HS on water and soil ecosystems [172, 173, 174]. Natural and artificial HS got attracted to rice cells [175], macrophyte of Ceratophyllum demersum, crustaceans—Gammarus pulex, and vertebrates—tadpoles of Rana arvalis [176], which support the hypothesis that HS is charged and naturally interacts with microorganisms. When HS penetrates or is taken up by a cell, the electric potential of the cell is disrupted, denying the cell the ability to provide support in terms of rigidity, shape and metabolism, thus creating pores through which vital intracellular structures are leaked out.

In an in vivo experiment, Vigneault et al. [177] discovered that Suwannee River HA and FA enhanced the release/leakage of the fluorescent probe sulforhodamine-B (SRB) encapsulated within 1-palmitoyl-2-oleoyl-sn-glycerol-3-phosphatidylcholine (POPC) vesicles. With regards to HA, a pH from 7.6 to 5.7 enhanced its surfactant-like effect. In conclusion, HS can alter the permeability of microbial cell, to create pores via which intracellular components are leaked out, killing the microorganism. However, the concentration, functionalisation (acylation), and pH of HS could potentially influence the biocidal activity.

According to Almatov and Akhmerov [178], 0.2–0.8 mg/mL mumie activated mitochondrial respiration and inhibited cellular succinate-oxidase and NADH-oxidase activity (mitochondrion). Similarly, mumie triggered the outflow of Ca2+ [160].

Previous studies [179, 180, 181] reported that mumie induced a dose-dependent elevation of superoxide dismutase, catalase, and glutathione peroxidase in rats. These enzymes are involved in the generation of ROS in an HA-induced antimicrobial or biological effects, which killed microorganisms and other grain storage pest.

A small-molecular size humic (LMSH) extracted from the feces of Nicodrilus and Allolobophora rosea enhanced the uptake of nitrate by plant roots and the accumulation of anions in the leaves. Further molecular analysis showed that LMSH influenced gene transcription in roots and long-distance effects in shoots as observed for Mha2 and the ZmNrt2.1 gene, respectively [182], which indicate HS can interfere with protein synthesis in microbes. FA and HA extracted from a podzol stimulated respiration in rat liver mitochondria at concentrations between 40 and 360 mg/L. Depending on the duration of contact with mitochondria, uncoupled oxidative phosphorylation may occur subsequently affecting the growth of the microorganism [183].

A product of oxidative DNA damage, 8-hydroxy-2′-deoxyguanosine increased significantly after treatment with HA, indicating the ability of HA to inflict damage on DNA. The endonuclease activity of the viral RNA polymerase was inhibited when it came in contact with HA [184]. The concentrations (5, 10, and 15 mgL-1) of HA and its organic extract significantly increased luciferase reporter gene activity in H4IIE.luc cells in a dose-dependent manner, which affected various molecular processes [185], thus killing the cell. The addition of HA (300 mg kg−1) to soil stimulated the growth of bot laurel plants and rhizospheric bacteria and actinomycetes. However, high dose (3000 mg kg−1), exerted an inhibitory effects [186]. The effects of HS on the hormone of Caenorhabditis elegans [170, 187], the sex ratio of Xiphophorus helleri [188], and the change in biochemical parameters of amphipod [189] were reported. These studies reiterate the potential biological effects of HS on microorganisms at the molecular level thus making them vulnerably for UDHSS-NPs.

Advertisement

5. Carbon-based nanoparticles/nanomaterials

Recently, carbon-based nanomaterials/particles (CNPs), which include nanotubes (i.e., double- or single-walled carbon nanotubes (DWCNT/SWCNTs)), fullerenes, and graphene oxide (GO) (Figure 3), have gained attention due to their potent biocidal activities. According to literature, the biocidal potency of these novel NPs is influenced by their physical/chemical properties, high adsorptive potentials, size, large surface area, and colloidal stability under wide range of pH. Increasing the NPs’ surface area led to a decrease in size, with concomitant increase in adsorption and absorption (into fungi cell), which improved interaction [190, 191, 192, 193, 194, 195, 196] with subsequent inhibition of fungal growth.

Figure 3.

Various carbon-based NPs [208].

The mycelia biomass and aflatoxin biosynthesis in A. flavus NRRL 3251 was negatively influenced at 10 μg mL−1 of fullerene C60 (fullerols C60(OH)24). The effects (growth arrest) was concentration-dependent. However, the antioxidative activity of the furrerols declined over time [197]. Hao and colleagues [198] investigated the fungicidal potentials of metal (copper oxide (CuO), ferric oxide (Fe2O3), and TiO2NPs) and carbon-based NPs (multiwalled carbon nanotubes, fullerene, and reduced graphene oxide) against Botrytis cinerea. The results showed that all the six NPs exhibited biocidal activity with 50 mg/L of fullerene showing the strongest antifungal effects.

Reduced graphene oxide (rGO) nanosheets inhibited the mycelial growth of A. niger, A. oryzae, and F. oxysporum with half maximal inhibitory concentrations (IC50) of 500, 500, and 250 μg/mL, respectively. The fungicidal activity as ascribed to the sharp edge of the rGO [199] which inflict injury on the cells, resulting in leaking of the cell components. Another hypothesis is that the organic functional groups on the fungi cell wall chemically interact with the ROS in rGO [200], which halts the uptake of nutrient and excretion of waste metabolites eventually killing the fungi.

Among the six carbon nanomaterials (SWCNTs, MWCNTs, GO, rGO, C60, and activated carbon (AC)) assessed for their fungicidal activity against pathogenic fungi (i.e., F. graminearum and F. poae), SWCNTs (500 μg/mL) exhibited the most potent activity, followed by MWCNTs, GO, and rGO respectively. However, the other two CNPs (C60 and AC) showed minimal activity, probably due to insufficient contact with fungal spores [201]. Conclusively, increasing the concentration of CNPs (62.5 < 125 < 250 < 500 μg/mL) increased the fungicidal potency. In a similar study, Wang et al. [202] reported that modifying the surface of MWCNTs with ▬OH, ▬COOH, and ▬NH2 improved the fungicidal activity (inhibition in spore elongation and germination) than the unmodified CNTs. It is hypothesized that modified CNTs formed a stable dispersions, which favoured interaction with spores, as a results enhanced antifungal activity. The authors observed a reduction in F. graminearum spore from 68.5, 54.5, 28.3, 27.4, and 29.5 μm, when 500 μg/mL MWCNTs (control), MWCNTs-COOH, MWCNTs-OH, and MWCNTs-NH2 were applied, respectively. Moreover, previous works [203, 204] documented that biological activity of nanotubes improved upon addition of functionalized aliphatic amide (covalent) and polyethylene glycol (PEG) and/or polyoxyethylene(40)nonylphenyl ether (IGPAL) (non-covalent) chemical groups [205]. Zare-Zardini et al. [206] conducted a covalent functionalization of MWCNTs with lysine and arginine under radiation. The modified MWCNTs exhibited potent biocidal activity against all test fungi (A. niger, A. fumigatus, C. albicans, P. chrysogenum, S. cerevisiae, F. culmorum, Microsporum canis, Trichophyton mentagrophytes, Trichophyton rubrum, and P. lilacinum) compared unmodified MWCNTs. Surprisingly, the fungicidal activity of MWCNTs-arginine against all the test fungi was slightly higher than MWCNTs-lysine. The authors hypothesized that the positive charge on arginine might have enhanced the binding of NPs on the fungal membrane and altered the genetic makeup (DNA). Thus, lysine and arginine could be utilized to improve the fungicidal activity of CTNs. Recently, Katerine et al. [207] reported the fungicidal activity of cotton fabric silica-silver carbon-based hybrid NPs against A. sp., Cladosporium sp. and Chaetomium globosum. The fabrics with high number carbon exerted the most increased biocidal activity on C. globosum and Aspergillus sp.

5.1 Mechanistic action of CNPs fungicidal activity

The ability of CNPs to interact and integrate into fungi cells determines their fungicidal activities. Wang et al. [201] reported the importance of surface contact of CNPS to their biocidal functionality. A transmission electron microscopy study showed CNPs interacted and integrated into spores and form an aggregation. It was hypothesized that the van der Waals force in CNTs was strong enough formed a bond with spores, as a result, triggered plasmolysis. Similarly, Zare-Zardini et al. [206] reported a strong interaction in functionalized CNTs with arginine than lysine with fungi membrane. These interactions may result in CNPs been internalized into fungi. Transmission electron microscopy (TEM) analysis showed direct evidence of nanographene been internalized in Caco-2 cells [209], which support the above finding. Moreover, cells treated with CNPs showed evidence of plasmolysis [201].

They compared images of healthy and treated (incubation for 3 h with CNPs) cell membranes of F. graminearum spores. They observed an intact, slick, compact, inerratic, and well-positioned cytoplasm for the untreated cell; however, after treatment the latter cell were transfigured, contracted and gathered. This shows solute lost through CNPs point of contact. Interaction of CNPs with fungi’ membrane led to a decrease in membrane integrity by the stresses exerted by the electrostatic forces between the microbial outer surface and CNTs, resulting in membrane oxidation [210]. CNPs are reported to be a contributing factor to the over-generation of ROS, which could trigger fungal cells to enter oxidative stress, causing excessive impairment to cellular components and permanent DNA laddering that could potentially lead biocidal activity against the cells [211, 212]. In contrast, Saha et al. [209] found that all CNPs (C1, C2, C3, C4 C5) assessed did not contribute to ROS production in Caco-2 cells. A decrease in ATP was observed. Conclusively, CNTs applied was internalized and disrupt the functionality of mitochondria which explains the reason for the low ATP observed. However, the treatment did not influence the production of ROS.

Advertisement

6. Toxicological aspects of NPs

According to Higashisaka et al. [213], NPs with diameters ≤100 nm are presently been used in various applications, including food production (e.g., to improve texture). An orally ingested NP can cross the gastrointestinal barrier, absorbed into the blood, and alter normal physiological functions, thus causing adverse health is to consumer(s) [213, 214]. Ezhkov and colleagues reported acute catarrhal inflammation on esophagus, stomach, and duodenum of mice fed with sapropel-NPs at a dose of 1.8 g/kg. However, 0.3 g/kg and 1.5 g/kg dose did not manifest any toxic effects [165].

A positive correlation was established between residues of Ag-NPs in rat organs and the NP suspension applied. NPs migrated from the luminal side to the intestinal epithelial cells via endocytosis or transcytosis, which are accumulated in the mentioned organs. However, all treated rats were able to excrete the NPs from most organs except the brain and testes [214, 215, 216, 217]. Cellular uptake of NPs is similar to mechanism of the antimicrobial activity as its also depends on size, surface charge, and dispersion or aggregation state [218, 219]. Rhodamine B (RhB) labeled carboxymethyl chitosan grafted NPs (RhBCMCNP) and chitosan hydrochloride grafted NPs (RhB-CHNP) bearing positive or negative charges used as model chitosan to elucidate the effects of particle size and surface charge on the cellular uptake of NPs revealed that the surface charges were attracted to the macrophages, and could be attributed to the electrostatic interactions between particles and phagocytic cells. Besides, different cell lines, irrespective particle size, and surface charge difference influence the uptake of NPs [219]. Kim et al. [220] detected traces of Ag-NPs in blood, liver and other organs after they orally fed rats at a dose of 30, 125, or 500 mg/kg BW/day. A significant (p < 0.05) dose-related decrease in the bodyweight of high-dose male rats at the fourth, fifth, and seventh weeks was observed; however, no significant dose-dependent changes in the female rats. Further hematological assays showed a significant increase (p < 0.01) in cholesterol in the both high-dose male and female rats. A significant increase (p < 0.01) in alkine phosphotase (ALP) was also indicated for the high-dose female rats. The authors reported no-observed-adverse-effect level (NOAEL) and lowest-observed-adverse-effect level (LOAEL) as 30 mg/kg BW/day and 125 mg/kg BW/day, respectively. Treatment with NPs of diameters ranging from 25 to 80 nm at a dose of 5000 mg/kg body weight altered the levels of alanine transaminase, aspartate transaminase, blood urea nitrogen, and lactate dehydrogenase, along with lesions on the liver and kidneys of female mice. Myocardial damage associated with groups showing a notable changes in serum LDH and alpha-HBDH levels compared to the control experiment. Also, a biodistribution test disclosed that TiO2 was predominantly retained in the liver, spleen, kidneys, and lung tissues, indicating TiO2-NPs were transported via endocytosis to other tissues and organs after their uptake by the gastrointestinal tract [221]. Contradictory finding was reported by Warheit et al. [222] where no adverse effects were manifested after orally fed rats with TiO2-NPs. However, the NOAEL on rats exposed for 90 days was >1000 mg/kg BW/day. In a similar study, Sharma et al. [223] divided male Swiss albino mice into three groups (group 1—vehicle control (water); group 2—ZnO nanoparticles (300 mg/kg body weight); group 3—ZnO nanoparticles (50 mg/kg)) and fed them with 50 and 300 mg/kg b.wt. ZnO-NPs for 14 consecutive days. ZnO-NPs induced oxidative stress, which damage the DNA and apoptosis in the mouse liver. Additionally, elevated levels of ALT and ALP serum and subsequent pathological lesions were observed in the treated mice. Lastly, at a higher dose (300 mg/kg) of ZnO-NPs, a significant (p < 0.05) induction of lipid peroxidation was observed in the liver, brain, and kidney (Figure 4) of the treated mice in comparison with the control test. Cho et al. [224] discovered ZnO-NPs had a higher absorption efficiency than TiO2-NPs in rats. ZnO-NP concentrations in the liver and kidney were significantly higher compared to the control, whereas with TiO2-NPs, no dramatic increase was detected in the sampled organs. In the feces, very high and low concentrations of Ti and Zn were detected, respectively. The concentration of ZnO in the spleen and brain was minimally elevated. Similarly, Ti concentrations were not drastically increased in urine; in contrast, it was Zn levels, that remarkably changed. Therefore, the absorption of various NPs could be attributed to the higher dissolution rate in the acidic gastric fluid; however, this might not be applicable when NPs are utilized during grain storage. When a stored grain undergoes sun drying, milling, etc., the levels of NPs may decrease to a level that could not affect the consumer health. Moreover, many NPs have received approval for application in many fields. Nevertheless, rigorous studies are warranted to expound on any risks or the safety of NPs use in grain storage. According to Zare-Zardini et al. [206], CNPs appeared less toxic to humans and animals compared to metal NPs and are therefore the better alternative and a novel method for reducing mycotoxin biosynthesis in grains.

Figure 4.

Zinc content in selected tissue of the mice (n = 5) after oral administration of ZnO nanoparticles (NPs) (300 mg/kg) for 14 consecutive days. Data represent mean±S.E.M. of three animals. *p < 0.05, compared to control. Modified with permission from ref 4495441125809.

Advertisement

7. Proposed methods of applying NPs during grain storage

  1. The first method is direct processing of grains with solutions of the required concentration of NPs. To achieve this, biocompatible NPs in an aqueous dispersion medium with pH values close to neutral should be used.

  2. Treatment with aerosols NPs could also be used if the NPs are dispersed evenly over the granary or silos. The aerosols to apply should be modified to prevent aggregation on grains. Using aerosols saves time and labor since additional drying is not required.

  3. The use of packages made from NPs during storage, transportation, and sale will extend the shelf life of grains. Alternatively, NPs formulated cubes could be place in jute bags with grains; however, periodic mixing is required to distribute the NPs.

  4. In our opinion, one of the most inexpensive methods is the use NPs in the production of materials for granaries, as well as treating interior and exterior surfaces of the storage facilities.

Advertisement

8. Conclusions

The ability of NPs to suppress the synthesis of mycotoxins in fungi and other microorganisms could be a breakthrough to curbe the issue of aflatoxin prevalence worldwide. NPs displayed excellent antifungal activity against important fungal species which contaminate grains with toxins during storage. The concentration, volume, type, and illumination (sunlight) significantly influenced the biostatic activity of NPs. Hence, these factors should carefully be considered when applying NPs in grain storage. The proposed NPs are environmentally friendly and pose no threat to consumer compared to some conventional methods of grain preservation. Several ex vivo, in vivo, and in vitro studies supports these claims. Moreover, NPs are biocompatible to the human system hence their usage in the food industry. Despite safety of NPs guaranteed by international safety organizations such as the Food Safety Authority, routine testing is required to understand the impact it has on grain nutritional, sensory, and other physicochemical parameters.

Advertisement

Acknowledgments

This work was supported by the Ministry of Science and Higher Education of Russia (Project No.075-15-2019-1896).

Advertisement

Disclosure statement

The authors reported no potential conflict of interest.

References

  1. 1. Narvankar DS, Singh CB, Jayas DS, White NDG. Assessment of soft X-ray imaging for detection of fungal infection in wheat. Biosystems Engineering. 2009;103:49-56
  2. 2. Nawrocka A, Cieœla J. Influence of silver nanoparticles on food components in wheat. International Agrophysics. 2013;27:49-55
  3. 3. El-Kady IA, Abdel-Hafez SII, El-Maraghy SS. Contribution to the fungal flora of cereal grains in Egypt. Mycopathologia. 1982;77:103-109
  4. 4. Kitherian S. Nano and bio-nanoparticles for insect control. Research Journal of Nanoscience and Nanotechnology. 2017;7(1):1-9
  5. 5. Alif AA, Thangapandiyan S. Comparative bioassay of silver nanoparticles and malathion on infestation of red flour beetle, Tribolium castaneum. Journal of Basic and Applied Zoology. 2019;80:55
  6. 6. Zallen R, Moret M. The optical absorption edge of brookite TiO2. Solid State Communications. 2006;137(3):154-157
  7. 7. Shukla RK, Sharma V, Pandey AK, Singh S, Sultana S, Dhawan A. ROS-mediated genotoxicity induced by titanium dioxide nanoparticles in human epidermal cells. Toxicology In Vitro. 2011;25(1):231-241
  8. 8. Morishige T, Yoshioka Y, Tanabe A, Yao X, Tsunoda S-I, Tsutsumi Y, et al. Titanium dioxide induces different levels of IL-1β production dependent on its particle characteristics through caspase-1 activation mediated by reactive oxygen species and cathepsin B. Biochemical and Biophysical Research Communications. 2010;392(2):160-165
  9. 9. Macwan DP, Dave PN, Shalini C. A review on nano-TiO2 sol–gel type syntheses and its applications. Journal of Materials Science. 2011;46:3669-3686
  10. 10. Buettner KM, Valentine AM. Bioinorganic chemistry of titanium. Chemical Reviews. 2012;112:1863-1881
  11. 11. Brun E, Barreau F, Veronesi G, Fayard B, Sorieul S, Chanéac C, et al. Titanium dioxide nanoparticle impact and translocation through ex vivo, in vivo and in vitro gut epithelia. Particle and Fibre Toxicology. 2014;11(1):13
  12. 12. Banerjee K, Thiagarajan P. A review of titanium di oxide nanoparticles—Synthesis, applications and toxicity concerns. Nanoscience Nanotechnology-Asia. 2014;4:132-143
  13. 13. Matsunaga T, Tamoda R, Nakajima T, Wake H. Photoelectrochemical sterilization of microbial cells by semiconductor powders. FEMS Microbiology Letters. 1985;29:211-214
  14. 14. Rincón A, Pulgarin C. Photocatalytical inactivation of E. coli: Effect of (continuous–intermittent) light intensity and of (suspended–fixed) TiO2 concentration. Applied Catalysis B: Environmental. 2003;44(3):263-284
  15. 15. Rincon A. Effect of pH, inorganic ions, organic matter and H2O2 on E. coli K12 photocatalytic inactivation by TiO2 implications in solar water disinfection. Applied Catalysis B: Environmental. 2004;51(4):283-302
  16. 16. Benabbou A, Derriche Z, Felix C, Lejeune P, Guillard C. Photocatalytic inactivation of Escherischia coli: Effect of concentration of TiO2 and microorganism, nature, and intensity of UV irradiation. Applied Catalysis B: Environmental. 2007;76(3-4):257-263
  17. 17. Pigeot-Rémy S, Simonet F, Errazuriz-Cerda E, Lazzaroni J, Atlan D, Guillard C. Photocatalysis and disinfection of water: Identification of potential bacterial targets. Applied Catalysis B: Environmental. 2011;104(3-4):390-398
  18. 18. Pigeot-Rémy S, Simonet F, Atlan D, Lazzaroni J, Guillard C. Bactericidal efficiency and mode of action: A comparative study of photochemistry and photocatalysis. Water Research. 2012;46(10):3208-3218
  19. 19. Zimbone M, Buccheri M, Cacciato G, Sanz R, Rappazzo G, Boninelli S, et al. Photocatalytical and antibacterial activity of TiO2 nanoparticles obtained by laser ablation in water. Applied Catalysis B: Environmental. 2015;165:487-494
  20. 20. Xing Y, Li X, Zhang L, Xu Q , Che Z, Li W, et al. Effect of TiO2 nanoparticles on the antibacterial and physical properties of polyethylene-based film. Progress in Organic Coatings. 2012;73(2-3):219-224
  21. 21. Maness PC, Smolinski S, Blake DM, Huang Z, Wolfrum EJ, Jacoby WA. Bactericidal activity of photocatalytic TiO2 reaction: Toward an understanding of its killing mechanism. Applied and Environmental Microbiology. 1999;65:4094-4098
  22. 22. Choi YS, Kim BW. Lactic acid recovery from fermentation broth using one-stage electrodialysis. Journal of Chemical Technology and Biotechnology. 2000;75:1145-1150
  23. 23. Wist J, Sanabria J, Dierolf C, Torres W, Pulgarin C. Evaluation of photocatalytic disinfection of crude water for drinking-water production. Journal of Photochemical and Photobiology A: Chemistry. 2002;147:241-246
  24. 24. Kim B, Kim D, Cho D, Cho S. Bactericidal effect of TiO2 photocatalyst on selected food-borne pathogenic bacteria. Chemosphere. 2003;52:277-281
  25. 25. Hur JS, Oh SO, Lim KM, Jung JS, Kim JW, Koh YJ. Novel effects of TiO2 photocatalytic ozonation on control of postharvest fungal spoilage of kiwifruit. Postharvest Biology and Technology. 2005;35:109-113
  26. 26. Maneerat C, Hayata Y. Antifungal activity of TiO2 photocatalysis against Penicillium expansum in vitro and in fruit tests. International Journal of Food Microbiology. 2006;107:99-103
  27. 27. Kühn KP, Chaberny IF, Massholder K, Stickler M, Benz VW, Sonntag H-G, et al. Disinfection of surfaces by photocatalytic oxidation with titanium dioxide and UVA light. Chemosphere. 2003;53(1):71-77
  28. 28. Gupta K, Singh RP, Pandey A, Pandey A. Photocatalytic antibacterial performance of TiO2 and Ag-doped TiO2 against S. aureus, P. aeruginosa and E. coli. Beilstein Journal of Nanotechnology. 2013;4:345-351
  29. 29. Fekete-Kertész I, Maros G, Gruiz K, Molnár M. The effect of TiO2 nanoparticles on the aquatic ecosystem: A comparative ecotoxicity study with test organisms of different trophic levels. Periodica Polytechnica, Chemical Engineering. 2016;60(4):231-243
  30. 30. Kau JH, Sun DS, Huang HH, Wong MS, Lin HC, Chang HH. Role of visible light-activated photocatalyst on the reduction of anthrax spore-induced mortality in mice. PLoS One. 2009;4(1)
  31. 31. Seven O, Dindar B, Aydemir S, Metin D, Ozinel M, Icli S. Solar photocatalytic disinfection of a group of bacteria and fungi aqueous suspensions with TiO2, ZnO and Sahara Desert dust. Journal of Photochemistry and Photobiology A: Chemistry. 2004;165(1-3):103-107
  32. 32. Chen F, Yang X, Wu Q . Antifungal capability of TiO2 coated film on moist wood. Building and Environment. 2009;44(5):1088-1093
  33. 33. Pokhum C, Viboonratanasri D, Chawengkijwanich C. New insight into the disinfection mechanism of Fusarium monoliforme and Aspergillus niger by TiO2 photocatalyst under low intensity UVA light. Journal of Photochemistry and Photobiology B: Biology. 2017;176:17-24
  34. 34. Wolfrum EJ, Huang J, Blake DM, Maness P-C, Huang Z, Fiest J, et al. Photocatalytic oxidation of bacteria, bacterial and fungal spores, and model biofilm components to carbon dioxide on titanium dioxide-coated surfaces. Environmental Science & Technology. 2002;36(15):3412-3419
  35. 35. Erkan A, Bakir U, Karakas G. Photocatalytic microbial inactivation over Pd doped SnO2 and TiO2 thin films. Journal of Photochemistry and Photobiology A: Chemistry. 2006;184(3):313-321
  36. 36. Anbeaki RAA, Hameed FR, Gassim FAG. Antifungal activity of titanum dioxide photocatalysis against Fusarium oxysporum f.sp. lycopersici. Euphrates Journal of Agricultural Science. 2009;1:14-26
  37. 37. Lonnen J, Kilvington S, Kehoe SC, Touati FA, McGuigan KG. Solar and photocatalytic disinfection of protozoan, fungal and bacterial microbes in drinking water. Water Research. 2005;39(5):877-883. DOI: 10.1016/j.watres.2004.11.023
  38. 38. Yao Y, Ohko Y, Sekiguchi Y, Fujishima A, Kubota Y. Self-sterilization using silicone catheters coated with Ag and TiO2 nanocomposite thin film. Journal of Biomedical Materials Research. Part B: Applied Biomaterials. 2008;85B(2):453-460. DOI: 10.1002/jbm.b.30965
  39. 39. Veselá M, Veselý M, Chomoucká J, Lipenská M. Photocatalytic disinfection of water using Ag/TiO2. Chemicke Listy. 2008;102:507-508
  40. 40. Sawada D, Ohmasa M, Fukuda M, Masuno K, Koide H, Tsunoda S, et al. Disinfection of some pathogens of mushroom cultivation by photocatalytic treatment. Mycoscience. 2005;46(1):54-60. DOI: 10.1007/s10267-004-0211-y
  41. 41. Giannantonio DJ, Kurth JC, Kurtis KE, Sobecky PA. Effects of concrete properties and nutrients on fungal colonization and fouling. International Biodeterioration & Biodegradation. 2009;63(3):252-259. DOI: 10.1016/j.ibiod.2008.10.002
  42. 42. Lu JW, Li FB, Guo T, Lin LW, Hou MF, Liu TX. TiO2 photocatalytic antifungal technique for crops diseases control. Journal of Environmental Sciences. 2006;18:397-401
  43. 43. Koide S, Nonami T. Disinfecting efficacy of a plastic container covered with photocatalyst for postharvest. Food Control. 2007;18(1):1-4. DOI: 10.1016/j.foodcont.2005.08.001
  44. 44. Sichel C, de Cara M, Tello J, Blanco J, Fernández-Ibáñez P. Solar photocatalytic disinfection of agricultural pathogenic fungi: Fusarium species. Applied Catalysis B: Environmental. 2007;74(1-2):152-160. DOI: 10.1016/j.apcatb.2007.02.005
  45. 45. Sichel C, Tello J, de Cara M, Fernández-Ibáñez P. Effect of UV solar intensity and dose on the photocatalytic disinfection of bacteria and fungi. Catalysis Today. 2007;129(1-2):152-160. DOI: 10.1016/j.cattod.2007.06.061
  46. 46. Lin CY, Li CS. Effectiveness of titanium dioxide photocatalyst filters for controlling bioaerosols. Aerosol Science and Technology. 2003;37(2):162-170. DOI: 10.1080/02786820300951
  47. 47. Lin CY, Li CS. Inactivation of microorganisms on the photocatalytic surfaces in air. Aerosol Science and Technology. 2003;37(12):939-946. DOI: 10.1080/02786820300900
  48. 48. Thabet S, Simonet F, Lemaire M, Guillard C, Cotton P. Impact of photocatalysis on fungal cells: Depiction of cellular and molecular effects on Saccharomyces cerevisiae. Applied and Environmental Microbiology. 2014;80(24):7527-7535
  49. 49. Fujishima A, Rao TN, Tryk DA. Titanium dioxide photocatalysis. Journal of Photochemistry and Photobiology C. 2000;1:1-21
  50. 50. Rajeshwar K, Osugi M, Chanmanee W, Chenthamarakshan C, Zanoni M, Kajitvichyanukul P, et al. Heterogeneous photocatalytic treatment of organic dyes in air and aqueous media. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2008;9(4):171-192
  51. 51. Rehman S, Ullah R, Butt A, Gohar N. Strategies of making TiO2 and ZnO visible light active. Journal of Hazardous Materials. 2009;170(2-3):560-569
  52. 52. Saravanan R, Gracia F, Stephen A. Basic principles, mechanism, and challenges of photocatalysis. In: Khan MM, Pradhan D, Sohn Y, editors. Nanocomposites for Visible Light-induced Photocatalysis. Springer Series on Polymer and Composite Materials. Switzerland: Springer International Publishing; 2017. pp. 19-40
  53. 53. Khan MM, Adil SF, Al-Mayouf A. Metal oxides as photocatalysts. Journal of Saudi Chemical Society. 2015;19(5):462-464
  54. 54. Konstantinou IK, Albanis TA. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: kinetic and mechanistic investigations. Applied Catalysis B: Environmental. 2004;49(1):1-14
  55. 55. Nakata K, Fujishima A. TiO2 photocatalysis: Design and applications. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2012;13(3):169-189
  56. 56. Gogniat G, Dukan S. TiO2 photocatalysis causes DNA damage via Fenton reaction-generated hydroxyl radicals during the recovery period. Applied and Environmental Microbiology. 2007;73(23):7740-7743
  57. 57. Riegel G, Bolton JR. Photocatalytic efficiency variability in TiO2 particles. The Journal of Physical Chemistry. 1995;99(12):4215-4224
  58. 58. Chen B, Le W, Wang Y, Li Z, Wang D, Lin L, et al. Targeting negative surface charges of cancer cells by multifunctional nanoprobes. Theranostics. 2016;6(11):1887-1898. DOI: 10.7150/thno.16358
  59. 59. Bishop JR, Schuksz M, Esko JD. Heparan sulphate proteoglycans fine-tune mammalian physiology. Nature. 2007;446(7139):1030-1037
  60. 60. Blake DM, Maness PC, Huang Z, Wolfrum EJ, Huang J, Jacoby WA. Application of the photocatalytic chemistry of titanium dioxide to disinfection and the killing of cancer cells. Separation and Purification Methods. 1999;28(1):1-50
  61. 61. Youngman RJ. Oxygen activation: is the hydroxyl radical always biologically relevant? Trends in Biochemical Sciences. 1984;9(6):280-283
  62. 62. Draper HH, Hadley M. Malondialdehyde determination as index of lipid peroxidation. Methods in Enzymology. 1990;186:421-431
  63. 63. Thabet S, Weiss-Gayet M, Dappozze F, Cotton P, Guillard C. Photocatalysis on yeast cells: Toward targets and mechanisms. Applied Catalysis B: Environmental. 2013;140-141:169-178
  64. 64. Schobert M, Gorisch H. Cytochrome C550 is an essential component of the quinoprotein ethanol oxidation system in Pseudomonas aeruginosa: Cloning and sequencing of the genes encoding cytochrome C550 and an adjacent acetaldehyde dehydrogenase. Microbiology. 1999;145(2):471-481
  65. 65. Kubacka A, Diez MS, Rojo D, Bargiela R, Ciordia S, Zapico I, et al. Understanding the antimicrobial mechanism of TiO2-based nanocomposite films in a pathogenic bacterium. Scientific Reports. 2014;4(1):4134
  66. 66. Dunford R, Salinaro A, Cai L, Serpone N, Horikoshi S, Hidaka H, et al. Chemical oxidation and DNA damage catalysed by inorganic sunscreen ingredients. FEBS Letters. 1997;418(1-2):87-90
  67. 67. Kikuchi Y, Sunada K, Iyoda T, Hashimoto K, Fujishima A. Photocatalytic bactericidal effect of TiO2 thin films: Dynamic view of the active oxygen species responsible for the effect. Journal of Photochemistry and Photobiology A: Chemistry. 1997;106(1-3):51-56
  68. 68. Hidaka H, Horikoshi S, Serpone N, Knowland J. In vitro photochemical damage to DNA, RNA and their bases by an inorganic sunscreen agent on exposure to UVA and UVB radiation. Journal of Photochemistry and Photobiology A: Chemistry. 1997;111(1-3):205-213
  69. 69. Cezara B, Alexandra C, Janeta T, Svetlana L, Steliana C, Elena D, et al. Effect of nano-oxides TiO2 and Fe3O4 on lipase biosynthesis by Aspergillus niger CNMN-FD-01 micromycete. Journal of the Academy of Sciences of Moldova. Life Sciences. 2017;2:125-130
  70. 70. Gomes SIL, Roca CP, Kammer FVD, Scott-Fordsmand JJ, Amorim MJB. Mechanisms of (photo) toxicity of TiO2 nanomaterials (NM103, NM104, NM105): Using high-throughput gene expression in Enchytraeus crypticus. Nanoscale. 2018;10(46):21960-21970
  71. 71. Dagata A, Fasulo S, Dallas LJ, Fisher AS, Maisano M, Readman JW, et al. Enhanced toxicity of ‘bulk titanium dioxide compared to ‘fresh and ‘aged nano-TiO2 in marine mussels (Mytilus galloprovincialis). Nanotoxicology. 2013;8(5):549-558
  72. 72. Blanco-Galvez J, Fernández-Ibáñez P, Malato-Rodríguez S. Solar photocatalytic detoxification and disinfection of water: Recent overview. Journal of Solar Energy Engineering. 2007;129(1). DOI: 10.1115/1.2390948
  73. 73. Elieh-Ali-Komi D, Hamblin MR. Chitin and chitosan: Production and application of versatile biomedical nanomaterials. International Journal of Advanced Research. 2016;4:411-427
  74. 74. Sudarshan NR, Hoover DG, Knorr D. Antibacterial action of chitosan. Food Biotechnology. 1992;6:257-272
  75. 75. Madhumathi K, Kumar PTS, Abhilash S, Sreeja V, Tamura H, Manzoor K, et al. Development of novel chitin/nanosilver composite scaffolds for wound dressing applications. Journal of Materials Science: Materials in Medicine. 2009;21(2):807-813
  76. 76. Quan H, Zhu F, Han X, Xu Z, Zhao Y, Miao Z. Mechanism of anti-angiogenic activities of chitooligosaccharides may be through inhibiting heparanase activity. Medical Hypotheses. 2009;73(2):205-206
  77. 77. Pangestuti R, Kim S-K. Neuroprotective properties of chitosan and its derivatives. Marine Drugs. 2010;8(7):2117-2128
  78. 78. Li Q , Dunn EJ, Grandmaison EW, Goosen MFA. Applications and properties of chitosan. Journal of Bioactive and Compatible Polymers. 1992;7:370-397
  79. 79. Fernandes JC, Tavaria FK, Soares JC, Ramos ÓS, Monteiro MJ, Pintado ME, et al. Antimicrobial effects of chitosans and chitooligosaccharides, upon Staphylococcus aureus and Escherichia coli, in food model systems. Food Microbiology. 2008;25(7):922-928
  80. 80. Wang Y, Zhou P, Yu J, Pan X, Wang P, Lan W, et al. Antimicrobial effect of chitooligosaccharides produced by chitosanase from Pseudomonas CUY8. Asia Pacific Journal of Clinical Nutrition. 2007;16:174-177
  81. 81. Khanafari A, Marandi R, Sanatei S. Recovery of chitin and chitosan from shrimp waste by chemical and microbial methods. Iranian Journal of Environmental Health Science and Engineering. 2008;5:19-24
  82. 82. Li XF, Feng XQ , Yang S. A mechanism of antibacterial activity of chitosan against Gram-negative bacteria. Chinese Journal of Polymer Science. 2010;31:148-153
  83. 83. Li XF, Feng XQ , Yang S, Fu GQ , Wang TP, Su ZX. Chitosan kills Escherichia coli through damage to be of cell membrane mechanism. Carbohydrate Polymers. 2010b;79(3):493-499
  84. 84. Limam Z, Selmi S, Sadok S, El-Abed A. Extraction and characterization of chitin and chitosan from crustacean by-products: Biological and physicochemical properties. African Journal of Biotechnology. 2011;10:640-647
  85. 85. Benhabiles M, Salah R, Lounici H, Drouiche N, Goosen M, Mameri N. Antibacterial activity of chitin, chitosan and its oligomers prepared from shrimp shell waste. Food Hydrocolloids. 2012;29(1):48-56
  86. 86. Yang E-J, Kim J-G, Kim J-Y, Kim S, Lee N, Hyun C-G. Anti-inflammatory effect of chitosan oligosaccharides in RAW 2647 cells. Open Life Sciences. 2010;5(1)
  87. 87. Yu T, Yu C, Chen F, Sheng K, Zhou T, Zunun M, et al. Integrated control of blue mold in pear fruit by combined application of chitosan, a biocontrol yeast and calcium chloride. Postharvest Biology and Technology. 2012;69:49-53
  88. 88. Rahman M, Mahmud T, Kadir J, Rahman RA, Begum M. Enhancing the efficacy of Burkholderia cepacia B23 with calcium chloride and chitosan to control anthracnose of papaya during storage. The Plant Pathology Journal. 2009;25(4):361-368
  89. 89. El-Ghaouth A, Smilanick JL, Wilson CL. Enhancement of the performance of Candida saitoana by the addition of glycolchitosan for the control of postharvest decay of apple and citrus fruit. Postharvest Biology and Technology. 2000;19(1):103-110
  90. 90. El-Ghaouth A, Smilanick JL, Brown GE, Ippolito A, Wisniewski M, Wilson CL. Application of Candida saitoana and glycolchitosan for the control of postharvest diseases of apple and citrus fruit under semi-commercial conditions. Plant Disease. 2000;84(3):243-248
  91. 91. Capdeville GD, Wilson CL, Beer SV, Aist JR. Alternative disease control agents induce resistance to blue mold in harvested ‘red delicious’ apple fruit. Phytopathology. 2002;92(8):900-908
  92. 92. Kaur P, Thakur R, Barnela M, Chopra M, Manuja A, Chaudhury A. Synthesis, characterization and in vitro evaluation of cytotoxicity and antimicrobial activity of chitosan-metal nanocomposites. Journal of Chemical Technology & Biotechnology. 2014;90(5):867-873
  93. 93. Hashim AF, Alghuthaymi MA, Vasil’kov AY, Abd-Elsalam KA. Polymer inorganic nanocomposites: A sustainable antimicrobial agents. In: Prasad R, editor. Advances and Applications through Fungal Nanobiotechnology. Switzerland: Springer International Publishing; 2016. pp. 265-290
  94. 94. Brunel F, Gueddari NEE, Moerschbacher BM. Complexation of copper(II) with chitosan nanogels: Toward control of microbial growth. Carbohydrate Polymers. 2013;92(2):1348-1356
  95. 95. Vokhidova NR, Sattarov ME, Kareva ND, Rashidova SS. Fungicide features of the nanosystems of silkworm (Bombyx mori) chitosan with copper ions. Microbiology. 2014;83(6):751-753
  96. 96. Sagoo S. The antimicrobial action of chitosan [PhD thesis]. Newington: South Bank University; 2003
  97. 97. Sagoo S, Board R, Roller S. Chitosan potentiates the antimicrobial action of sodium benzoate on spoilage yeasts. Letters in Applied Microbiology. 2002;34(3):168-172
  98. 98. Pospieszny H, Chirkov S, Atabekov J. Induction of antiviral resistance in plants by chitosan. Plant Science. 1991;79(1):63-68
  99. 99. Kulikov SN, Chirkov SN, Il’Ina AV, Lopatin SA, Varlamov VP. Effect of the molecular weight of chitosan on its antiviral activity in plants. Applied Biochemistry and Microbiology. 2006;42(2):200-203
  100. 100. Yang F-L, Li X-G, Zhu F, Lei C-L. Structural characterization of nanoparticles loaded with garlic essential oil and their insecticidal activity against Tribolium castaneum (Herbst) (Coleoptera: Tenebrionidae). Journal of Agricultural and Food Chemistry. 2009;57(21):10156-10162
  101. 101. Mgbeahuruike E, Fyhrquist P, Vuorela H, Julkunen-Tiitto R, Holm Y. Alkaloid-rich crude extracts, fractions and piperamide alkaloids of piper guineense possess promising antibacterial effects. Antibiotics. 2018;7(4):98
  102. 102. Nychas GJE, Skandamis PN. Antimicrobials from herbs and spices. In: Roller S, editor. Natural Antimicrobials for the Minimal Processing of Foods. Cambridge: Woodhead Publishing Limited; 2000. pp. 176-192
  103. 103. Ippolito A, Nigro F. Natural antimicrobials in postharvest storage of fresh fruits and vegetables. In: Roller S, editor. Natural Antimicrobials for the Minimal Processing of Foods. Cambridge, UK: Woodhead Publishing Limited; 2000. pp. 201-224
  104. 104. Fang SW, Li CF, Shih DYC. Antifungal activity of chitosan and its preservative effect on low-sugar candied kumquat. Journal of Food Protection. 1994;57(2):136-140. DOI: 10.4315/0362-028x-57.2.136
  105. 105. Roller S, Covill N. The antifungal properties of chitosan in laboratory media and apple juice. International Journal of Food Microbiology. 1999;47(1-2):67-77
  106. 106. Badawy MEI, Ahmed M, Rabea EI. Bactericidal and fungicidal activities of different molecular weight chitosan samples. Pest Control. 2006;14:19-34
  107. 107. Martinez LR, Mihu MR, Han G, Frases S, Cordero RJ, Casadevall A, et al. The use of chitosan to damage Cryptococcus neoformans biofilms. Biomaterials. 2010;31(4):669-679. DOI: 10.1016/j.biomaterials.2009.09.087
  108. 108. Jiang T, Feng L, Li J. Changes in microbial and postharvest quality of shiitake mushroom (Lentinus edodes) treated with chitosan-glucose complex coating under cold storage. Food Chemistry. 2012;131(3):780-786. DOI: 10.1016/j.foodchem.2011.08.087
  109. 109. Tikhonov VE, Stepnova EA, Babak VG, Yamskov IA, Palma-Guerrero J, Jansson HB, et al. Bactericidal and antifungal activities of a low molecular weight chitosan and its N-/2(3)-(dodec-2-enyl)succinoyl/-derivatives. Carbohydrate Polymers. 2006;64(1):66-72
  110. 110. Ramos-García M, Bosquez-Molina E, Hernández-Romano J, Zavala-Padilla G, Terrés-Rojas E, Alia-Tejacal I, et al. Use of chitosan-based edible coatings in combination with other natural compounds, to control Rhizopus stolonifer and Escherichia coli DH5α in fresh tomatoes. Crop Protection. 2012;38:1-6. DOI: 10.1016/j.cropro.2012.02.016
  111. 111. Ziani K, Fernández-Pan I, Royo M, Maté JI. Antifungal activity of films and solutions based on chitosan against typical seed fungi. Food Hydrocolloids. 2009;23(8):2309-2314. DOI: 10.1016/j.foodhyd.2009.06.005
  112. 112. Vu KD, Hollingsworth RG, Leroux E, Salmieri S, Lacroix M. Development of edible bioactive coating based on modified chitosan for increasing the shelf life of strawberries. Food Research International. 2011;44(1):198-203. DOI: 10.1016/j.foodres.2010.10.037
  113. 113. Zhang M, Tan T, Yuan H, Rui C. Insecticidal and fungicidal activities of chitosan and oligo-chitosan. Journal of Bioactive and Compatible Polymers. 2003;18(5):391-400. DOI: 10.1177/0883911503039019
  114. 114. Sebti I, Martial-Gros A, Carnet-Pantiez A, Grelier S, Coma V. Chitosan polymer as bioactive coating and film against Aspergillus niger contamination. Journal of Food Science. 2005;70(2):M100–M104. DOI: 10.1111/j.1365-2621.2005.tb07098.x
  115. 115. Bhaskara Reddy MV, Arul J, Ait-Barka E, Angers P, Richard C, Castaigne F. Effect of chitosan on growth and toxin production by Alternaria alternata f. sp. lycopersici. Biocontrol Science and Technology. 1998;8(1):33-43. DOI: 10.21273/hortsci.32.3.467f
  116. 116. Dash M, Chiellini F, Ottenbrite R, Chiellini E. Chitosan—A versatile semi-synthetic polymer in biomedical applications. Progress in Polymer Science. 2011;36(8):981-1014
  117. 117. Ifuku S. Chitin and chitosan nanofibers: Preparation and chemical modifications. Molecules. 2014;19(11):18367-18380
  118. 118. Xing K, Chen XG, Kong M, Liu CS, Cha DS, Park HJ. Effect of oleoyl-chitosan nanoparticles as a novel antibacterial dispersion system on viability, membrane permeability and cell morphology of Escherichia coli and Staphylococcus aureus. Carbohydrate Polymers. 2009;76(1):17-22
  119. 119. Xing K, Chen XG, Liu CS, Cha DS, Park HJ. Oleoyl-chitosan nanoparticles inhibits Escherichia coli and Staphylococcus aureus by damaging the cell membrane and putative binding to extracellular or intracellular targets. International Journal of Food Microbiology. 2009;132(2-3):127-133
  120. 120. Kong M, Chen XG, Xing K, Park HJ. Antimicrobial properties of chitosan and mode of action: A state of the art review. International Journal of Food Microbiology. 2010;144(1):51-63
  121. 121. Lin G, Zhang Q , Lin X, Zhao D, Jia R, Gao N, et al. Enhanced photoluminescence of gallium phosphide by surface plasmon resonances of metallic nanoparticles. RSC Advances. 2015;5(60):48275-48280
  122. 122. Chien RC, Yen MT, Mau JL. Antimicrobial and antitumor activities of chitosan from shiitake stipes, compared to commercial chitosan from crab shells. Carbohydrate Polymers. 2016;138:259-264
  123. 123. Severino R, Ferrari G, Vu KD, Donsì F, Salmieri S, Lacroix M. Antimicrobial effects of modified chitosan based coating containing nanoemulsion of essential oils, modified atmosphere packaging and gamma irradiation against Escherichia coli O157:H7 and Salmonella Typhimurium on green beans. Food Control. 2015;50:215-222
  124. 124. Yang TC, Chou CC, Li CF. Antibacterial activity of N-alkylated disaccharide chitosan derivatives. International Journal of Food Microbiology. 2005;97(3):237-245
  125. 125. Ignatova M, Starbova K, Markova N, Manolova N, Rashkov I. Electrospun nano-fibre mats with antibacterial properties from quaternised chitosan and polyvinyl alcohol. Carbohydrate Research. 2006;341(12):2098-2107
  126. 126. Jia Z, Shen D, Xu W. Synthesis and antibacterial activities of quaternary ammonium salt of chitosan. Carbohydrate Research. 2001;333(1):1-6
  127. 127. Xie Y, Liu X, Chen Q . Synthesis and characterization of water-soluble chitosan derivate and its antibacterial activity. Carbohydrate Polymers. 2007;69(1):142-147
  128. 128. Qin C, Li H, Xiao Q , Liu Y, Zhu J, Du Y. Water-solubility of chitosan and its antimicrobial activity. Carbohydrate Polymers. 2006;63(3):367-374
  129. 129. Li Z, Yang F, Yang R. Synthesis and characterization of chitosan derivatives with dual-antibacterial functional groups. International Journal of Biological Macromolecules. 2015;75:378-387
  130. 130. Márquez IG, Akuaku J, Cruz I, Cheetham J, Golshani A, Smith ML. Disruption of protein synthesis as antifungal mode of action by chitosan. International Journal of Food Microbiology. 2013;164(1):108-112
  131. 131. Ing LY, Zin NM, Sarwar A, Katas H. Antifungal activity of chitosan nanoparticles and correlation with their physical properties. International Journal of Biomaterials. 2012:1-9
  132. 132. Raafat D, Sahl H-G. Chitosan and its antimicrobial potential—A critical literature survey. Microbial Biotechnology. 2009;2(2):186-201
  133. 133. Helander IM, Nurmiaho-Lassila EL, Ahvenainen R, Rhoades J, Roller S. Chitosan disrupts the barrier properties of the outer membrane of Gram-negative bacteria. International Journal of Food Microbiology. 2001;71(2-3):235-244
  134. 134. Zakrzewska A, Boorsma A, Brul S, Hellingwerf KJ, Klis FM. Transcriptional response of Saccharomyces cerevisiae to the plasma membrane-perturbing compound chitosan. Eukaryotic Cell. 2005;4(4):703-715
  135. 135. Je JY, Kim SK. Chitosan derivatives killed bacteria by disrupting the outer and inner membrane. Journal of Agricultural and Food Chemistry. 2006;54(18):6629-6633
  136. 136. Torr KM, Chittenden C, Franich RA, Kreber B. Advances in understanding bioactivity of chitosan and chitosan oligomers against selected wood-inhabiting fungi. Holzforschung. 2005;59(5):559-567
  137. 137. Helander IM, Wright AV, Mattila-Sandholm TM. Potential of lactic acid bacteria and novel antimicrobials against Gram-negative bacteria. Trends in Food Science & Technology. 1997;8(5):146-150
  138. 138. Kong M, Chen XG, Liu CS, Liu CG, Meng XH, Yu LJ. Antibacterial mechanism of chitosan microspheres in a solid dispersing system against E. coli. Colloids and Surfaces B: Biointerfaces. 2008;65(2):197-202
  139. 139. Raafat D, Bargen KV, Haas A, Sahl H-G. Insights into the mode of action of chitosan as an antibacterial compound. Applied and Environmental Microbiology. 2008;74(23):3764-3773
  140. 140. Rabea EI, Badawy ME-T, Stevens CV, Smagghe G, Steurbaut W. Chitosan as antimicrobial agent: Applications and mode of action. Biomacromolecules. 2003;4(6):1457-1465
  141. 141. Goy RC, Britto D, Assis OBG. A review of the antimicrobial activity of chitosan. Polímeros. 2009;19:241-247
  142. 142. Sahariah P, Másson M. Antimicrobial chitosan and chitosan derivatives: A review of the structure–activity relationship. Biomacromolecules. 2017;18(11):3846-3868
  143. 143. Palma-Guerrero J, Lopez-Jimenez J, Pérez-Berná A, Huang I-C, Jansson HB, Salinas J, et al. Membrane fluidity determines sensitivity of filamentous fungi to chitosan. Molecular Microbiology. 2010;75:1021-1032
  144. 144. Hadwiger LA, Beckman JM, Adams MJ. Localization of fungal components in the pea-Fusarium interaction detected immunochemically with anti-chitosan and anti-fungal cell wall antisera. Plant Physiology. 1981;67(1):170-175
  145. 145. Tarsi R, Muzzarelli R, Guzmàn C, Pruzzo C. Inhibition of Streptococcus mutans adsorption to hydroxyapatite by low-molecular-weight chitosans. Journal of Dental Research. 1997;76(2):665-672
  146. 146. Liu XF, Guan YL, Yang DZ, Li Z, Yao KD. Antibacterial action of chitosan and carboxymethylated chitosan. Journal of Applied Polymer Science. 2000;79(7):1324-1335
  147. 147. Hosseinnejad M, Jafari SM. Evaluation of different factors affecting antimicrobial properties of chitosan. International Journal of Biological Macromolecules. 2016;85:467-475
  148. 148. Barakova NV, Sharova NY, Juskauskajte AR, Mityukov AS, Romanov VA, Nsengumuremyi D. Fungicidal activity of ultradisperse humic sapropel suspensions. Agronomy Research. 2017;15(3):639-648
  149. 149. Avdeyeva LN, Kovalyenko TA, Krivonos OI, Plaksin GV, Strunina NN. Determination of sapropel chemical composition. Chemistry and Chemical Engineering. 2009;52:121-123 (in Russian)
  150. 150. Nsengumuremyi D, Barakova NV, Romanov VA, Guzeva AV. The effect of sapropel extracts on microflora and physicochemical parameters of dried distillers’ grain. Agronomy Research. 2018;16(S2):1457-1465
  151. 151. Nsengumuremyi D, Adadi P, Ukolova MV, Barakova NV. Effects of ultradisperse humic sapropel suspension on microbial growth and fermentation parameters of barley distillate. Fermentation. 2019;5:24
  152. 152. Kireycheva LV, Khokhlova OB. Elemental composition of humic substances insapropel deposits. RAAS Newsletter. 2000;4:59-62 (in Russian)
  153. 153. Gorbunovskaya OM, Kurzo VB. New methods of sapropel composition analysis. Chemistry and Chemical Engineering. 2001;2:73-81 (in Russian)
  154. 154. Dolgopolov VN. Trials of Gumival™ yield better bovine, swine and fowl stock productivity. In: Proceedings of Conference on Summaries and Perspectives of Humin Preparations in Livestock Industry; Moscow. 2006. pp. 40-43 (in Russian)
  155. 155. Kulikova NA, Filippova OI, Abros’kin DP, Klyain OI. Modern approach to humic substance biological activity assessment. In: Proceedings of International Conference on Biodiagnostics in Soil Ecology Analysis. Moscow: BINOM; 2013. pp. 116-111 (in Russian)
  156. 156. Platonov VV, Khadartsev AA, Fridzon KY, Chunosov SN. Chemical composition and biological activity of sapropel from the Lake Glubokoe (Tatarstan). Journal of New Medical Technologies. 2014;21(3):199-204 (in Russian)
  157. 157. Rumyantsev VA, Mityukov AS, Kryukov LN, Yaroshevich GS. Unique properties of humic substances from sapropel. Doklady Earth Sciences. 2017;473(2):482-484
  158. 158. Shtin SM. Lake Sapropels Complex Utilization. Moscow: MSU; 2005 (in Russian)
  159. 159. Kosov VI. Sapropel Resources, Engineering, Geoecology. Saint Petersburg: Nauka; 2007. p. 244 (in Russian)
  160. 160. Schepetkin I, Khlebnikov A, Kwon BS. Medical drugs from humus matter: Focus on mumie. Drug Development Research. 2002;57(3):140-159
  161. 161. Buzlama AV, Chernov YN. Pharmacological properties, action pathways and prospectives of humic substances in medicine: An analysis. Russian Journal of Experimental and Clinical Pharmacology. 2010;73:43-48 (in Russian)
  162. 162. Ouyang K, Walker SL, Yu XY, Gao CH, Huanga Q , Cai P. Metabolism, survival, and gene expression of Pseudomonas putida to hematite nanoparticles mediated by surface-bound humic acid. Environmental Science: Nano. 2018;5:682-695
  163. 163. Wu M, Song M, Liu M, Jiang C, Li Z. Fungicidal activities of soil humic/fulvic acids as related to their chemical structures in greenhouse vegetable fields with cultivation chronosequence. Scientific Reports. 2016;6(1):32858
  164. 164. Ong KJ, Felix LC, Boyle D, Ede JD, Ma G, Veinot JG, et al. Humic acid ameliorates nanoparticle-induced developmental toxicity in zebrafish. Environmental Science: Nano. 2017;4(1):127-137
  165. 165. Ezhkov VO, Yapparov AK, Ezhkova AM, Yapparov IA, Ezhkova GO, Faizrakhmanov RN, et al. Study of the action of nanostructured sapropel at different doses on the morphological and functional condition of the gastrointestinal tract in albino mice. Nanotechnologies in Russia. 2016;11(7-8):497-505
  166. 166. Perdue EM. Chemical composition, structure, and metal binding properties. In: Hessen DO, Tranvik L, editors. Aquatic Humic Substances: Ecology and Biogeochemistry. Berlin: Springer-Verlag; 1998. pp. 41-61
  167. 167. Cosgrove DJ. Creeping, walls, softening fruit, and penetrating pollen tubes: The growing role of expansins. Proceedings of the National Academy of Sciences of the United States of America. 1997;94:5504-5505
  168. 168. Lipke PN, Ovalle R. Cell wall architecture in yeast: New structure and new challenges. Journal of Bacteriology. 1998;15:3735-3740
  169. 169. Gubbins PO, Anaissie EJ. Antifungal therapy. In: Clinical Mycology. 2nd ed. 2009. pp. 161-195
  170. 170. Bittner M. Direct effects of humic substances on organisms [Master’s thesis]. Brno: Masaryk University; 2006
  171. 171. Lofts S, Tipping E, Sanchez A, Dodd B. Modelling the role of humic acid in radiocaesium distribution in a British upland peat soil. Journal of Environmental Radioactivity. 2002;61(2):133-147. DOI: 10.1016/S0265-931X(01)00118-7
  172. 172. García-Mina J, Antolín M, Sanchez-Diaz M. Metal-humic complexes and plant micronutrient uptake: A study based on different plant species cultivated in diverse soil types. Plant and Soil. 2004;258(1):57-68
  173. 173. Heil CA. Influence of humic, fulvic and hydrophilic acids on the growth, photosynthesis and respiration of the dinoflagellate Prorocentrum minimum (Pavillard) Schiller. Harmful Algae. 2005;4(3):603-618
  174. 174. Schulten H-R, Leinweber P. New insights into organic-mineral particles: Composition, properties and models of molecular structure. Biology and Fertility of Soils. 2000;30(5-6):399-432
  175. 175. Wang W-H, Ray CM, Jones MN. The fate of 14C-labelled hurnic substances in rice cells in culture. Journal of Plant Physiology. 1999;154(2):203-211
  176. 176. Steinberg CEW, Höss S, Brüggemann R. Further evidence that humic substances have the potential to modulate the reproduction of the nematode Caenorhabditis elegans. International Review of Hydrobiology. 2002;87(1):121
  177. 177. Vigneault B, Percot A, Lafleur M, Campbell PGC. Permeability changes in model and phytoplankton membranes in the presence of aquatic humic substances. Environmental Science & Technology. 2000;34(18):3907-3913
  178. 178. Almatov KT, Akhmerov AI. Mumie effect on oxidative phosphorylation and enzymes of mitochondrial respiratory chain of rat liver, small and large intestine. Tashkent State Medical Institute. 1977;7:11-14
  179. 179. Ghosal S. Shilajit. Part 15. Shilajit: Its origin and vital significance. In: Mukerjee B, editor. Traditional Medicine. New Delhi: Oxford & IBH; 1993. pp. 308-319
  180. 180. Bhattacharya SK, Sen AP, Ghosal S. Effects of shilajit on biogenic free radicals. Phytotherapy Research. 1995;9(1):56-59
  181. 181. Ghosal S, Tripathi VK, Chauhan S. Active constituents of Emblica officinalis: Part 1—The chemistry and antioxidative effects of two new hydrolysable tannins, Emblicanin A and B. Indian Journal of Chemistry: Section B Organic and Medicinal Chemistry. 1996;35:941-948
  182. 182. Quaggiotti S. Effect of low molecular size humic substances on nitrate uptake and expression of genes involved in nitrate transport in maize (Zea mays L.). Journal of Experimental Botany. 2004;55(398):803-813
  183. 183. Visser S. Effect of humic substances on mitochondrial respiration and oxidative phosphorylation. Science of The Total Environment. 1987;62:347-354
  184. 184. Lu F-J, Tseng S-N, Li M-L, Shih S-R. In vitro anti-influenza virus activity of synthetic humate analogues derived from protocatechuic acid. Archives of Virology. 2002;147(2):273-284
  185. 185. Bittner M, Janošek J, Hilscherová K, Giesy J, Holoubek I, Bláha L. Activation of Ah receptor by pure humic acids. Environmental Toxicology. 2006;21(4):338-342
  186. 186. Vallini G, Pera A, Avio L, Valdrighi M, Giovannetti M. Influence of humic acids on laurel growth, associated rhizospheric microorganisms, and mycorrhizal fungi. Biology and Fertility of Soils. 1993;16(1):1-4
  187. 187. Höss S, Bergtold M, Haitzer M, Traunspurger W, Steinberg CE. Refractory dissolved organic matter can influence the reproduction of Caenorhabditis elegans (Nematoda). Freshwater Biology. 2001;46(1):1-10
  188. 188. Meinelt T, Schreckenbach K, Knopf K, Wienke A, StuBer A, Steinberg CEW. Humic substances affect physiological condition and sex ratio of swordtail (Xiphophorus helleri Heckel). Aquatic Sciences. 2004;66(2):239-245. DOI: 10.1007/s00027-004-0706-9
  189. 189. Timofeyev MA, Wiegand C, Burnison BK, Shatilina ZM, Pflugmacher S, Steinberg CE. Impact of natural organic matter (NOM) on freshwater amphipods. Science of The Total Environment. 2004;319(1-3):115-121
  190. 190. Kang S, Herzberg M, Rodrigues DF, Elimelech M. Antibacterial effects of carbon nanotubes: Size does matter! Langmuir. 2008;24(13):6409-6413
  191. 191. Buzea C, Pacheco II, Robbie K. Nanomaterials and nanoparticles: Sources and toxicity. Biointerphases. 2007;2(4):MR17–MR71
  192. 192. Cataldo F, Da Ros T. Medicinal Chemistry and Pharmacological Potential of Fullerenes and Carbon Nanotubes. Trieste: Springer; 2008
  193. 193. Wang JT, Chen C, Wang E, Kawazoe Y. A new carbon allotrope with six-fold helical chains in all-sp2 bonding networks. Scientific Report. 2014;4:4339
  194. 194. Sokolov VI, Stankevich IV. The fullerenes-new allotropic forms of carbon: Molecular and electronic structure, and chemical properties. Russian Chemical Reviews. 1993;62(5):419-435
  195. 195. Gibson N, Shenderova O, Luo TJM, Moseenkov S, Bondar V, Puzyr A, et al. Colloidal stability of modified nanodiamond particles. Diamond and Related Materials. 2009;18:620-626
  196. 196. Maleki Dizaj S, Mennati A, Jafari S, Khezri K, Adibkia K. Antimicrobial activity of carbon-based nanoparticles. Advanced Pharmaceutical Bulletin. 2015;5(1):19-23. DOI: 10.5681/apb.2015.003
  197. 197. Kovač T, Borišev I, Crevar B, Kenjeri FC, Kovač M, Strelec I, et al. Fullerol C60(OH)24 nanoparticles modulate aflatoxin B1 biosynthesis in Aspergillus flavus. Scientific Reports. 2018;8:12855. DOI: 10.1038/s41598-018-31305-9
  198. 198. Hao Y, Cao X, Ma C, Zhang Z, Zhao N, Ali A, et al. Potential applications and antifungal activities of engineered nanomaterials against Gray mold disease agent Botrytis cinerea on rose petals. Frontiers in Plant Science. 2017;8:1332. DOI: 10.3389/fpls.2017.01332
  199. 199. Sawangphruk M, Srimuk P, Chiochan P, Sangsri T, Siwayaprahm P. Synthesis and antifungal activity of reduced graphene oxide nanosheets. Carbon. 2012;50(14):5156-5516. DOI: 10.1016/j.carbon.2012.06.056
  200. 200. Hudler GW. Magical Mushrooms Mischievous Molds. Vol. 7. Princeton, NJ: Princeton University; 1998. p. 7
  201. 201. Wang X, Liu X, Chen J, Han H, Yuan Z. Evaluation and mechanism of antifungal effects of carbon nanomaterials in controlling plant fungal pathogen. Carbon. 2014;68:798-806. DOI: 10.1016/j.carbon.2013.11.072
  202. 202. Wang X, Zhou Z, Chen F. Surface modification of carbon nanotubes with an enhanced antifungal activity for the control of plant fungal pathogen. Materials. 2017;10:1375. DOI: 10.3390/ma10121375
  203. 203. Canuto de Menezes BR, Rodrigues KF, da Silva Fonseca BC, Ribas RG, do Amaral Montanheiro TL, Gilmar Patrocínio Thim GP. Recent advances in the use of carbon nanotubes as smart biomaterials. Journal of Materials Chemistry B. 2019;7:1343-1360. DOI: 10.1039/C8TB02419G
  204. 204. Klumpp C, Kostarelos K, Prato M, Bianco A. Functionalized carbon nanotubes as emerging nanovectors for the delivery of therapeutics. Biochimica et Biophysica Acta (BBA) – Biomembranes. 2006;1758(3):404-412. DOI: 10.1016/j.bbamem.2005.10.008
  205. 205. Panchakarla LS, Govindaraj A. Covalent and non-covalent functionalization and solubilization of double-walled carbon nanotubes in nonpolar and aqueous media. Journal of Chemical Sciences. 2008;120(6):607-611
  206. 206. Zare-Zardini H, Amiri A, Shanbedi M, Memarpoor-Yazdi M, Asoodeh A. Studying of antifungal activity of functionalized multiwalled carbon nanotubes by microwave-assisted technique. Surface and Interface Analysis. 2013;45:751-755. DOI: 10.1002/sia.5152
  207. 207. Katerine I, Romina AA, Jorge ES, Natalia B, José RVB, Carlos RG, et al. Antifungal activity of cotton fabrics finished modified silica-silver carbon-based hybrid nanoparticles. Textile Research Journal. 2019;89(5):825-833. DOI: 10.1177/0040517518755792
  208. 208. Yuan X, Zhang X, Sun L, Wei Y, Wei X. Cellular toxicity and immunological effects of carbon-based nanomaterials. Particle and Fibre Toxicology. 2019;16:18. DOI: 10.1186/s12989-019-0299-z
  209. 209. Saha D, Heldt CL, Gencoglu MF, Vijayaragavan KS, Chen J, Saksule A. A study on the cytotoxicity of carbon-based materials. Materials Science and Engineering C. 2016;68:101-108. DOI: 10.1016/j.msec.2016.05.094
  210. 210. Al-Jumaili A, Alancherry S, Bazaka K, Jacob MV. Review on the antimicrobial properties of carbon nanostructures. Materials. 2017;10:1066. DOI: 10.3390/ma10091066
  211. 211. Gurunathan S, Han JW, Dayem AA, Eppakayala V, Kim JH. Oxidative stress-mediated antibacterial activity of graphene oxide and reduced graphene oxide in Pseudomonas aeruginosa. International Journal of Nanomedicine. 2012;7:5901-5914. DOI: 10.2147/IJN.S37397
  212. 212. Krishnamoorthy K, Veerapandian M, Zhang LH, Yun K, Kim SJ. Antibacterial efficiency of graphene nanosheets against pathogenic bacteria via lipid peroxidation. Journal of Physical Chemistry C. 2012;116:17280-17287. DOI: 10.1021/jp3047054
  213. 213. Higashisaka K, Yoshioka Y, Tsutsumi Y. Applications and safety of nanomaterials used in the food industry. Food Safety. 2015;3(2):39-47
  214. 214. Zande MVD, Vandebriel RJ, Doren EV, Kramer E, Rivera ZH, Serrano-Rojero CS, et al. Distribution, elimination, and toxicity of silver nanoparticles and silver ions in rats after 28-day oral exposure. ACS Nano. 2012;6(8):7427-7442
  215. 215. Guan M, Zhu QL, Liu Y, Bei YY, Gu ZL, Zhang XN, et al. Uptake and transport of a novel anticancer drug-delivery system: Lactosyl-norcantharidin-associated N-trimethyl chitosan nanoparticles across intestinal Caco-2 cell monolayers. International Journal of Nanomedicine. 2012;7:1921-1930
  216. 216. Saremi S, Dinarvand R, Kebriaeezadeh A, Ostad SN, Atyabi F. Enhanced oral delivery of docetaxel using thiolated chitosan nanoparticles: preparation, in vitro and in vivo studies. BioMed Research International. 2013:1-8
  217. 217. He B, Lin P, Jia Z, Du W, Qu W, Yuan L, et al. The transport mechanisms of polymer nanoparticles in Caco-2 epithelial cells. Biomaterials. 2013;34(25):6082-6098
  218. 218. Geiser M, Rothen-Rutishauser B, Kapp N, Schürch S, Kreyling W, Schulz H, et al. Ultrafine particles cross cellular membranes by nonphagocytic mechanisms in lungs and in cultured cells. Environmental Health Perspectives. 2005;113(11):1555-1560
  219. 219. He C, Hu Y, Yin L, Tang C, Yin C. Effects of particle size and surface charge on cellular uptake and biodistribution of polymeric nanoparticles. Biomaterials. 2010;31(13):3657-3666
  220. 220. Kim YS, Song MY, Park JD, Song KS, Ryu HR, Chung YH, et al. Subchronic oral toxicity of silver nanoparticles. Particle Fibre Toxicology. 2010;7:20
  221. 221. Wang J, Zhou G, Chen C, Yu H, Wang T, Ma Y, et al. Acute toxicity and biodistribution of different sized titanium dioxide particles in mice after oral administration. Toxicology Letters. 2007;168(2):176-185
  222. 222. Warheit D, Brown S, Donner E. Acute and subchronic oral toxicity studies in rats with nanoscale and pigment grade titanium dioxide particles. Food and Chemical Toxicology. 2015;84:208-224
  223. 223. Sharma V, Singh P, Pandey AK, Dhawan A. Induction of oxidative stress, DNA damage and apoptosis in mouse liver after sub-acute oral exposure to zinc oxide nanoparticles. Mutation Research/Genetic Toxicology and Environmental Mutagenesis. 2012;745(1-2):84-91
  224. 224. Cho WS, Kang BC, Lee JK, Jeong J, Che JH, Seok SH. Comparative absorption, distribution, and excretion of titanium dioxide and zinc oxide nanoparticles after repeated oral administration. Particle and Fibre Toxicology. 2013;10(1):9

Written By

Daniel Nsengumuremyi, Parise Adadi, Gavers K. Oppong, Nadezhda V. Barakova and Elena F. Krivoshapkina

Reviewed: 17 June 2020 Published: 15 September 2020