Open access peer-reviewed chapter

Nanoparticles as New Therapeutic Agents against Candida albicans

Written By

Hilda Amelia Piñón Castillo, Laila Nayzzel Muñoz Castellanos, Rigoberto Martínez Chamorro, Reyna Reyes Martínez and Erasmo Orrantia Borunda

Submitted: 26 April 2018 Reviewed: 18 July 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.80379

From the Edited Volume

Candida Albicans

Edited by Doblin Sandai

Chapter metrics overview

1,972 Chapter Downloads

View Full Metrics

Abstract

Candida albicans is an opportunistic dimorphic yeast. This organism is pathogen associated to superficial and systemic infections. Actually, Candida albicans represents an emergent pathogen especially in a patient with some immunity compromises. Added to this, the use of antifungal in an indiscriminate form has increased the resistance of the existing drugs. In this aspect, the nanotechnology generates the possibility of creating new therapeutic agents. Nanoparticles are structures of 1–100 nm with special physicochemical characteristics that allow it to function as therapeutic agents or as carriers of these. Palladium, silver, and gold metallic nanoparticles and iron, titanium, zinc, and copper oxides have been used as growth inhibitors. These nanoparticles have been proved alone or in form of nanocomposites. The objective of this chapter is to describe the state of the art of the use of nanoparticles as inhibitors of the growth of Candida albicans, as well as the most relevant results regarding the mechanisms involved in this inhibition.

Keywords

  • nanotechnology
  • resistance
  • metallic and oxide nanoparticles

1. Introduction

Candidiasis is one of the most common infections worldwide. These diseases are generated by Candida albicans and Candida non-albicans group. Candida albicans is an opportunistic dimorphic yeast. It is a commensal organism in gastrointestinal, respiratory, and genitourinary tracts. Normally, it is a harmless organism that can turn into an opportunistic organism in immunocompromised or immunologically deficient individuals [1, 2, 3].

Candida albicans represents the predominant yeast in oral, genital, and cutaneous infections (≥90%) [2, 4, 5]. However, the invasion of blood torrent by Candida species is the infection more frequently diagnosed. The frequency of Candida albicans isolated in candidemia is of 46.3%, followed by C. glabrata (24.4%) [3, 6]. The mortality frequency in patients with invasive candidemia ranges from 40 to >70%, generating a loss of about 1.7 billion dollars every year [1, 7].

For the treatment of invasive infection by Candida, few options are available. The polyenes (amphotericin B and nystatin) are the chosen treatment by excellence. However other antifungals are used as azoles [8]. The azoles (fluconazole and itraconazole) and new azoles (ravuconazole, posaconazole, and voriconazole) are selected for prophylaxis, empirical therapy, and diagnostic-driven therapy [9]. The use of echinocandins (caspofungin, anidulafungin, and micafungin) is driven to therapy on invasive candidiasis [1].

However, the adverse effects of antifungal therapy as nephrotoxicity, red blood cell (RBC) toxicity, and arrhythmias for amphotericin B, hepatotoxicity associated with fluconazole, cardiotoxicity and gastrointestinal (GIT) disturbances attributed to itraconazole, voriconazole exhibiting neurological and hepatic toxicities, and posaconazole shown to elevate serum aminotransferase levels and cause mild-to-moderate hepatic toxicity limit its use. These adverse effects are attributed to the action mechanism. Azoles and polyenes have ergosterol as a target from the cell membrane, and the ergosterol has a high similarity between ergosterol and human cholesterol [8]. Azoles inhibit the ergosterol biosynthesis, specifically the lanosterol 14α-demethylase. This generates sterol precursor accumulation, resulting in a cell membrane instability and the loss of viability [3].

The echinocandins are more specific for fungi. This therapeutic antifungal inhibits the enzyme 1,3-b-D-glucan synthase and generates the depletion of polymer synthesis and osmotic stability loss [1, 10].

Due to the discriminated antifungal agent use, Candida albicans has developed different resistance mechanisms. Three principal mechanisms have been reported: the efflux pump, which is a membrane transport protein that ejects the azoles out of the cell; the structural changes of the enzyme target, mutations on lanosterol 14 α-demethylase, changing the three-dimensional structure, and, as consequence, the decreased affinity for this enzyme; and the last mechanism is the increase in the production of the target enzyme; this generates the inhibition of only a portion of the enzyme, and the rest are still active [1, 2, 3, 8, 11].

On the other hand, the survival characteristics of Candida albicans in the host make this organism difficult to kill. The virulence factors include the biofilm formation capacity, the yeast-hyphae morphogenesis, the adhesion to surfaces, and the capacity to secrete enzymes as hydrolases.

Because of this, the research for more specific antifungal therapies has been the focus of attention. In this area, the nanotechnology offers an opportunity to design and create a new pharmaceutical agent that can be specific. As pharmaceutical agent nanoparticles such as copper, silver, and palladium, among others, have been employed in order to inhibit the growth of any microorganism, in particular, Candida albicans.

The objective of this chapter is to describe the state of the art of the use of nanoparticles as inhibitors of the growth of Candida albicans, as well as the most relevant results regarding the mechanisms involved in this inhibition.

Advertisement

2. Candida albicans taxonomy

Fungi form one of the largest eukaryotic kingdoms; they have a very broad function of decomposing biopolymers and different compounds in dead or alive hosts, as well as synthesizing bioactive compounds. The yeast Candida albicans originally was described as Blastomycetes class belonging to Monilial order and Cryptococcaceae family. Actually, with gene sequencing, the new taxonomy classifies Candida albicans into Ascomycota, (dikaryon-producing fungi), Hemiascomycetes subclass, and Saccharomycotina order [12, 13].

Candida albicans is a common commensal on mucosal surfaces under normal conditions, mainly of the gastrointestinal and vaginal tracts. The commensal stage is frequently described as harmless to the host, and the fungus is controlled by the normal microbiota, epithelial barriers, and the innate immune system [14, 15]. This is the most important opportunistic medical pathogenic fungus, causing candidiasis, which is an acute, subacute, or chronic infection involving any part of the body [15]. The weakening of the defense mechanisms of the host and the adaptation capacity of the microorganism turn this fungus into an aggressive pathogen, starting with superficial infections into a systemic process, mainly in immunosuppressed or HIV patients [16].

Advertisement

3. Candida albicans cell wall

Candida albicans cell wall is composed mainly of carbohydrates; three basic main constituents are polysaccharides:

  1. Branched polymers of glucose with β-1,3 and β-1,6 bonds, they are called β-glucans.

  2. Unbranched polymers of N-acetyl-D-glucosamine (GlcNAc) with β-1,4 bonds; the chitin, which contributes to the fiber insolubility; and theβ-1,3 glucan-chitin complex, which is the major constituent of the inner wall. β-1,6 glucan links the components of the inner and outer walls; thus, chitin and β-1,3 are structural polysaccharides and convey strength and shape to the cell wall [15, 17, 18].

  3. Finally mannose polymers, mannans, covalently associated with proteins (glucomannan proteins).

The mannan of the outer cell wall is less structured but allows this fungus breakage of epithelial barriers or dysfunction of the immune system converting Candida albicans from a commensal to a pathogen. The cell walls also contain proteins, 6–25%, and minor amounts of lipids, 1–7% [15, 17, 18].

This arrangement is responsible for its cell morphology and provides protection to the cell against physical, chemical, and biological (immune host system) aggression [16].

Likewise, the cell wall participates in both growth and morphological transitions between yeast and hyphae; this conversion from the unicellular yeast to multicellular filamentous fungus is essential for Candida virulence [19]. It has cell surface proteins linked to glycosylphosphatidylinositol (GPI) [20], including an N-terminal signal sequence and a C-terminal GPI-anchor addition signal; some of these proteins are secreted to the extracellular environment, but do not remain cell associated. Some of them are hydrolytic enzymes, which hydrolyze complex substrates into small units, to carry out into the cell as a source of nutrition. If these polymers are host targets, all of these facilitate colonization or invasion of Candida albicans, and then such enzymes also function as virulence factors [17]. For therapeutic strategies, to establish the functional relationships between Candida albicans cell wall’s proteins and virulence is a challenge [20].

Advertisement

4. Candida albicans virulence factors

Candida spp. invade several anatomical hosts, mainly in patients who are immunocompromised or debilitated. This organism expresses several virulence factors that allow the host-pathogen interaction and that lead to the establishment of infection [21, 22].

Virulence factors include host recognition biomolecules (adhesins) due to cell surface hydrophobicity which enhanced adherence and resistance to phagocytosis; another important factor is its ability to grow either in yeast or in hyphal form (dimorphism, especially in the presence of N-acetyl-D-glucosamine) [23], secreting aspartyl proteases and phospholipases [21].

The phenomenon of the quorum sensing (QS) has been described as a contributor to morphogenesis, playing a role in regulating the yeast to filamentous fungus; one of the conditions is the dependence on cell density associated with the regulation by QS. The QS has an important role in the ability of Candida albicans to be virulent and survive within the host [23]. Candida albicans uses multiple redundant and/or synergistic signaling pathways to integrate host signals to promote hyphal development, tissue invasion, and virulence during infection, exacerbating the pathogenesis of invasive candidiasis [24].

Advertisement

5. Candida albicans biofilm formation

Candidavirulence is a result of its capacity to form biofilms on the surface of biomaterials and on catheters and prosthetic devices [25, 26, 27]. Candida albicans biofilms are composed of hyphae, pseudohyphae, yeast cells, and an extracellular matrix, where the hyphae play an integral role within this complex [26]. Biofilm formation presents an initial adherence of yeast cells (0–2 h), followed by germination and microcolony formation (2–4 h), filamentation (4–6 h), monolayer development (6–8 h), proliferation (8–24 h), and maturation (24–48 h). Serum and saliva preconditioning films and Candida albicans preconditioning films increased the initial attachment of the fungus but had minimal effect on subsequent biofilm formation [25, 28]. This kind of biofilm has clinical repercussions because it is the target of both chemical and immunological antifungals. Due of their intrinsic resistance to almost all antifungals in clinical use, increased resistance to this therapy, and the ability of cells within biofilms to withstand host immune defenses, the antifungal resistance of biofilms results most probably from the conjunction of several mechanisms that act in a time-dependent manner [26].

Advertisement

6. Candida albicans resistance mechanisms

In order for Candida albicans to survive, these microorganisms have developed diverse mechanisms. Biofilm formation, enzyme modification, and the pump efflux are the mechanisms reported to antibiotic resistance [29]. As described previously, biofilms are a complex of cells, DNA, and polymeric matrix, among others. The polymeric matrix has a crucial role in the antifungal resistance since it does not allow antibiotic penetration inside the biofilm. This phenomenon helps the internal cell to survive [25].

Resistance to azole antifungal has been described in three forms: [1] change in the affinity of cytochrome P450 sterol 14α-demethylase (gen Erg11p), [2] increase in the enzyme expression, and [3] changes in the sterol desaturase. Other alterations in an enzyme involved in the cell wall construction have been described [29, 30].

The pump efflux ejects the azoles outside the cell, without any effect inside of it. This efflux is mediated by an ATP-binding cassette (ABC) and the major facilitator superfamily. This is a complex family of genes that involve several CDR genes, including core membrane pore composed by ABC segments in the membrane; two segments are in the cytosolic side, which provide energy to pump the antibiotic outside of the cell [25, 29, 30].

It is an important highlight that this resistance could be a derivative in the resistance to other antifungals.

Candida albicans resistance could be due to one or a combination of all the resistance mechanisms described.

Advertisement

7. Nanotechnology as an antifungal

Nanotechnology is a new technology which manipulates the matter in a nanoscale order (1 × 10−9 m). In this order, the nanoparticles have been used as antimicrobial, as antifungal, or as carried molecules. Nanoparticles have a size from 1–100 nm on any of this dimension. In this scale, nanoparticles have particularity physicochemical characteristics [31]. Within these characteristics, their electron configuration confers an extraordinary quantum effect. These quantum effects enhance their pharmacokinetic effect and generate nanoparticles that can be used as a protein, antibodies, or inclusive specific molecule carriers [32].

As an antifungal, nanoparticles show cell wall damage, an oxidative stress increase, and DNA interaction [33]. Nevertheless, the NPs toxicity mechanisms are dependent on the nanoparticle nature, size, and shape and capping nature, among other characteristics.

This chapter reviewed some NPs characteristics such as nanoparticle nature, synthesis, size, and shape and their influence on the toxic mechanisms in Candida albicans and their influence on the virulence factors.

7.1. Silver nanoparticles

Silver has been used for several decades as a disinfectant. Inclusively, silver nitrate was used in burns in order to control infections. However, with the discovery of penicillin, the use of silver was slow. Nowadays, the increase of resistance to antifungal puts attention again on the use of silver as an antibiotic, though the use of silver ions implies complications in their solubility and availability [34].

In this sense, the silver nanoparticles (AgNPs) offer a possibility. AgNPs offer the possibility of using minimal doses and act in microorganisms which present antibiotic resistance or decrease the virulence factors.

The use of different AgNPs synthesis methods caused different effects. The general method is the salt reduction by a reducing agent. However, the addition stabilizing agent has been proved. The stabilizing or capping agent limits the NPs’ growth, avoids the agglomeration, and controls the shape, size, and superficial charge [35].

The NP synthesis by plant extract or secondary metabolites from microorganisms offers a new possibility. The phenolic compounds, proteins, anthocyanin, and carboxylic acid, among others, work as reducing agent and stabilizing or capping agent [35, 36].

The green AgNPs synthesis through the use of microorganisms as the use of Actinobacteria Pilimelia columellifera subsp. pallida [37], Bacillus species [38, 39], filamentous fungi Monascus purpureus [7], and Macromycetes as Pleurotus sajor-caju [40], as well as extract plant as Tulsi leaf extract [41] or extract from Phoenix dactylifera [34] has been proven. All these methods present inhibition of Candida albicans, with different efficiencies (Table 1).

Inhibition test NPs size (nm) Synthesis method Capped Reference
18 mm (inhibition zone) 15–20 Cassia roxburghii + Ketoconazole NR Moteriya et al. 2017 [42]
64 μg/mL (MIC) 12.7 Pilimelia columellifera subsp. pallida SL19 Proteins Wypij et al. 2017 [37]
16.7 ± 0.25 mm (inhibition zone) 2–8 Monascus purpureus NR El-Baz et al., 2016 [7]
10 μg/mL (BIC) 40–55 Dodonaea viscosa and Hyptis suaveolens Detection of functional groups as phenol, alcohol, proteins, heterocyclic amines Muthamil et al. 2018 [43]
40 μg/mL (MIC) 5–95 Bacillus safensis Proteins Lateef et al. 2015 [38]
12 mm (inhibition zone with 30 μL of solution) 18–46 Gracilaria corticata Detection of functional groups as alcohol, carbonyl, amino groups Kumar et al. 2013 [44]
250 μg/mL (MIC) and 500 μg/mL (MFC) 3.77–33.22 Pleurotus sajor-caju NR Musa et al. 2018 [40]
60 μg/mL (MIC90) and 120 (MFC) 2–7 Tulsi leaf extract NR Khatoon et al., 2015 [45]
20 mm (inhibition zone with 80 μg/20 mL) 21.65–41.05 Phoenix dactylifera Proteins Oves et al. 2018 [34]
75 ppm reduce 83%/150 ppm (MFC) 70 Red cabbage Anthocyanin compounds Ocsoy et al. 2017 [46]
08 μg/mL (MIC) 9–130 Enterococcus faecalis NR Ashajyothi 2016 [47]
40–60 μg/mL(MIC) 4–39 Citrus lemon aqueous Juice + CTAB Surrounded by a layer Rahisuddin et al. 2015 [48]
10.8 ± 0.8 (inhibition zone) 12–85 Clove extract Detection of functional groups such as methoxy, alcohol, carboxilic acid, aliphatic group Parlinska-Wojtan et al. 2017 [49]
21–12 cm in combination with antifungal (inhibition zone) 10–90 Flower broth of Tagetes erecta NR Padalia et al. 2015 [50]
12.14 μg/mL (MIC) 47.0 ± 2.0 J. curcas Protein layer Kumar et al. 2017 [51]
10.78 μg/mL (MIC) 7.6 ± 0.5 L. grandis
70% fold increase (combination with antifungal, inhibition zone) 1.836–135.78 B. subtilis 10,833 NR Ghiut et al. 2018 [39]
50% fold increase (combination with antifungal, inhibition zone) 1.012–73.83 B. amyloliquefaciens NR
14 mm(inhibition zone) 5–10 Ocimum sanctum + CTAB Not visible Aazam and Zaheer 2016 [35]
30 μg/mL (MIC) 4–36 Actinobacterial strain SF23 NR Anasane et al. 2016 [52]
40 μg/mL (MIC) 8–60 Actinobacteria C9 NR

Table 1.

Biogenic synthesized AgNPs against Candida albicans.

MIC, minimal inhibition concentration.

MFC, minimal fungicide concentration.

NR, not reported.

BIC, biofilm inhibitory concentration.

The microorganisms Bacillus amyloliquefaciens 1853 and Bacillus subtilis 10,833 have been used for the AgNPs synthesis. Culture media were employed as a reducing agent. The scanning electronic microscopy has shown AgNPs produced with capping. Bacillus strains have been used in the AgNPS production. Bacillus amyloliquefaciens strain 1853 produced AgNPs with a size of 1.071 and other of 73.83 nm. The Bacillus subtilis strain 10,833 produced tree population sizes of AgNPs at 1.863, 10.866, and 135.78 nm. AgNPs produced by both microorganisms were employed to evaluate the Candida albicans inhibition. No differences were found between AgNPs; inclusive addition of antibiotic with the AgNPs does not improve the inhibition [38].

Clove extract was used in order to produce AgNPs. The high content of eugenol, β-caryophyllene, humulene, chavicol, methyl salicylate, α-ylangene, and eugenone; flavonoids such as eugenin, rhamnetin, kaempferol, and eugenitin; triterpenoids like oleanolic acid, stigmasterol, and campesterol; and several sesquiterpenes allowed the clove to act as a reducing and capping agent in AgNPs synthesis. The DLS analysis showed AgNPs size of 12 and 85 nm. The AgNPs produced to generate a complete inhibition of Candida albicans in 40-fold dilution were applied. The results obtained suggest that the functional groups’ incorporation into the AgNPs generates a synergism between clove and NPs [49].

In order to evaluate the AgNPs effect on biofilms, some AgNPs have been evaluated. As described, biofilms are a Candida albicans virulence factor. The change from yeast to hyphae is crucial for the biofilm process. In this process, Candida albicans produces a polysaccharide matrix, carbohydrates, proteins, and signaling molecules [43]. Changes in the structural biofilm conformation of Candida albicans exposed to AgNPs were reported [28, 51, 53, 54].

The use of latex plants in the AgNPs green synthesis was proved. J. curcas and L. grandis extracts produced AgNPs with a size of 47 ± 2.0 nm and 7.6 ± 0.5 nm, respectively. The Fourier transform infrared (FTIR) spectra suggest the incorporation of phenolic OH, amides, amines, and aldehyde in the AgNPs as a stabilizing agent. The smallest AgNPs showed a better effect in plankton and biofilms of Candida albicans than the big ones. The AgNPs synthesized proved to be efficient in the inhibition of biofilm formation, changes in morphology and cell lysis when candida albicans was exposed [51].

Also, silver nitrate chemical reduction by sodium citrate was evaluated in intermediate and mature Candida albicans biofilms. The preformatted biofilms were exposed to 54 μg/mL of AgNPs. After 5 h of exposition, biofilms showed a decrease in the number of viable cells, and no differences were found in intermediate or mature biofilm [28].

The results obtained show the importance of the NPs size, being the smallest nanoparticles which show the highest inhibition [51]. Lara et al. (2015) showed that biofilm exposed to AgNPs exhibits a few Candida albicans filamentous formations. Electron microscopy analysis showed AgNPs accumulation and internalization [55].

Besides the AgNPs effect in biofilm, the AgNPs have been used alone showing the inhibition, cell wall damage, and incorporation of Ag into the cell. It was used in combination with different antifungals amphotericin b [51, 56] or nystatin and chlorhexidine digluconate, showing a synergic effect [54]. Other molecules as cationic carboxilane have been used, proving a high Candida albicans inhibition with small AgNPs concentration [1.8 mg/mL; [57]]. Also, the AgNPs have been capped with L-3,4-dihydroxyphenyl-alanine, showing a minimal fungicide concentration of 31.2–62.5 μg/mL [58].

The direction in the investigation with silver nanoparticles has been driven in the green synthesis due to the low toxicity generated in the human cell and the increase of the toxicity in bacteria and yeast [59]. Their chemical characteristics as a metallic oxidative state and superficial area, among others, allow the NPs to interact with the microorganism and inhibit their growth.

7.2. Zinc oxide nanoparticles

Due to ZnO low toxicity, ZnO has been used in a medical device such as drug carriers, antibacterial agents, and bioimaging probes, among others. For this reason, the ZnO nanoparticles (ZnONPs) have been used in Candida albicans control. The ZnONPs application is reflected by the annual production between 550 and 5500. This is more than 10–100 times higher than any other nanoparticle [60]. As the AgNPs, the ZnONPs characteristics depend on the size, synthesis method, and superficial charge, among others [61]. In order to control these characteristics, the green synthesis method has been used to fabricate ZnONPs with a size, shape, and superficial charge with the desirable characteristic [62, 63].

A promissory ZnONPs application area is their use as antimicrobial. Their effect has been evaluated in soil, plants, bacteria, and fungi [60, 64, 65, 66] ZnONPs different synthesis methods in ZnONPs have a big influence in the Candida albicans inhibition [68].

The use of Atalantia monophylla extracts produced ZnONPs 30 nm that have effect inhibition in Candida albicans growth. The author proposed the ZnONPs interaction with the cell wall, penetration to the cell, ion liberation, and the oxidative stress production, triggering the protein, lipid, and DNA oxidation [67].

The microwave aqueous extract from Suaeda aegyptiaca produced ZnONPs with a size of 80 nm with a capping produced by metabolites from the plant. The effect of the capping was studied. ZnONPs produced by precipitation method were studied and compared with the plant extract produced. ZnONPs Suaeda aegyptiaca produced a major inhibitory effect than the ZnONPs without capping agent, suggesting a synergic effect [68].

Extracts for different plants have been proved to synthesize ZnONPs in order to evaluate the toxicity of the resulting nanoparticles in Candida albicans inhibition (Table 2).

Inhibition test NPs size (nm) Synthesis method Capped Reference
24 mm (inhibition zone) 30 Atalantia monophylla NR Vijayakumar et al. 2018 [67]
80 μM (MIC)/>2590 μM (MFC) 8 Chelidonium majus NR Dobrucka et al. 2018 [69]
34 ± 1.28 mm (inhibition zone with 100 g/ml) 32–40 Glycosmis pentaphylla Alkane, C=O stretching, other groups Vijayakumar et al., 2018 [70]
25 μg/mL (MIC)/50 μg/mL (MFC) 18 Azadirachta indica Proteins Elumalai and Velmurugan 2015 [71]
Strong activity 36 S. myriocystum Protein, carbohydrates, flavonoids, tannins, mannitol Nagarajan and Kuppusamy 2013 [72]
12.5 mm (inhibition zone with 150 mg/ml) 68.64 Cassia auriculata Alkene, alcohol, ether, and alkane Padalia et al. 2018 [73]
20.3 ± 0.57 mm (inhibition zone with 200 μg/mL)/25 μg/mL (MIC) 50 μg/mL (MFC) 10–30 Vitex trifolia Proteins, alkanes, aromatic group Elumalai et al. 2015 [71]
18.01 mm (inhibition zone 100 μg/mL) 80 Suaeda aegyptiaca Alcohol, aldehyde, amine Rajabi et al. 2018 [68]
Inhibition test NPs size (nm) Synthesis method Capped Reference
24 mm (zone inhibition) 30 nm Atalantia monophylla NR Vijayakumar et al. 2018 [67]
80 μM (MIC)/>2590 μM (MFC) 8 nm Chelidonium majus NR Dobrucka et al. 2018 [74]

Table 2.

Biogenic synthesized ZnONPs against Candida albicans.

MIC, minimal inhibition concentration.

MFC, minimal fungicide concentration.

NR, not reported.

ZnONPs hybrid has been synthesized using chitosan, gelatin, and polystyrene, among others [75, 76, 77].

Dhillon et al. 2014 [75] used chitosan as a capping and bridge to carry on citric acid, glycerol, starch, and whey powder using the nanospray drying method. The ZnONPs produced have a size range of 93.2–402.5 nm. The smallest ZnONPs were the chitosan starch NPs. Aggregation effect was observed in the largest ZnONPs; this phenomenon could be explained by the superficial charge. However, these ZnONPs have a small effect in Candida albicans inhibition.

Collagen was used as a reducing and capping agent. The ZnONPs produced with this method showed a size between 20 and 50 nm negatively charged. ZnONPs coated by collagen showed a MIC of 99.7 ± 0.99 μg/mL, whereas the control with zinc acetate showed an MIC of 297.9 ± 2.0 μg/mL. Biofilms exposed to increase the concentration of ZnONPs showed inhibition as a result of the ZnONPs concentration [78].

Gelatin was used for the ZnONPs as well. ZnONPs produced with gelatin showed a diameter of 20 nm with a negative superficial charge. The ZnONPs produced were employed to biofilm exposition. ZnONPs in a concentration of 50 μg/mL showed a thickness diminution, resulting in a weak adherence biofilm compared with the biofilm treated with fluconazole [72].

Also, egg albumin was used as a ZnONPs template. The synthesis method produced spherical ZnONPs with a range between 10 and 20 nm. The FITR analysis showed the interaction of egg albumin and ZnONPs. In this work, ZnONPs characterization in the culture media showed importance. The use of 45 μg/ml in Sabouraud’s dextrose (SD) nutrient media does not significantly affect ZnONPs stability, size, and integrity. The synthesized ZnONPs showed Candida albicans inhibition in a dose-dependent manner. Changes in the Candida albicans morphology show cavity formation in cells exposed to 15 μg/ml [79].

ZnONPs were evaluated both alone and as a part of nanocomposites with polystyrene. The use of ZnONPs alone showed Candida albicans inhibition. However, when ZnONPs are used in nanocomposites, a major concentration of ZnONPs to inhibit Candida albicans is necessary [77].

Nanocomposites with ZnONPs have been produced with different materials in order to increase the growth inhibition of microorganisms [45, 46]. However, the results of the antifungal activity were contradictory; ZnONPs capped by surfactant in cotton fiber have proved the inhibition of Candida albicans in 94% [68], while the incorporation of ZnONPs in polylactic acid films does not inhibit the Candida albicans growth [80].

The principal ZnONPs toxicity mechanisms reported are (1) ion liberation, (2) interaction of ZnONPs with the cell wall, and (3) stress oxidative generation [81]. The ion liberation and the interaction of NPs with the cell trigger the generation of oxidative stress; with that the cell lost their viability [79].

7.3. Copper oxide nanoparticles

Copper exhibits good characteristics such as its antimicrobial activity, chemical stability, and thermal resistance. Due to their toxicity characteristic, copper has been used by Egyptians for water disinfection. The Aztecs used copper to treat sore throats, and the Persian and Indians used it to treat eye infection and venereal ulcers [82].

Copper has an advantage in comparison with other materials, is cheaper than any other, and is easy to oxidize to copper oxide nanoparticles (CuONPs). The CuONPs can easily make nanocomposites with polymers, macromolecules inclusive of other metals [83, 84].

The CuONPs have recovered importance in the investigation due to their low cost and the variety of applications. CuONPs have been used in the pharmaceutical, medicine, and electronic industries, among others.

Diverse synthesis methods have been proved to create CuONPs including chemical [85], electrochemical [86], and green synthesis [87].

The synthesis method influences the characteristics as the size, shape, and agglomeration of the CuONPs [88].

CuONP electrochemical production with different solvents and reaction times produces different NPs sizes from 3 to 200 nm. The CuONPs were obtained in water-acetonitrile, with sodium hydroxide as the electrolyte. The CuONPs obtained in these conditions showed a size around 20 nm. The Candida albicans exposition to 25 and 50 μg/ml generates a delay in the growth curves. Added to that, 50 μg/ml inhibited completely Candida albicans growth [86].

CuONPs prepared by precipitation method coated with acetate were evaluated together with fluconazole. CuONPs coated with acetate were more toxic than CuONPs without coating. CuONPs in combination with fluconazole showed the complete Candida albicans inhibition. The fractional inhibitory concentration index (FICI) suggested the additive to moderate synergism between CuONPs and fluconazole [89].

The use of dispersant as ethylene glycol and Tween 80 proved to be efficient in the production of CuONPs capable to the Candida albicans growth inhibition. Inhibition with CuONPs was more efficient than the exposition to itraconazole [90]. Also, the effect of CuONPs was evaluated in Candida albicans from oral cavities [91], showing that strain isolated from the patients needs until 1000 ppm to generate an inhibitory effect in Candida albicans.

CuONPs capped with pyrimidine derivatives have shown a size of 10 nm. The CuONPs showed interaction with DNA and antioxidant activity. However, the CuONPs showed Candida albicans inhibition [92].

Nanocomposites with the incorporation of Fe [93] Cd and Ba CuONPs [94], as well as the incorporation of CuONPs into cotton fiber [95, 96], polyurethane [97], polyester [98], copolymer microgel [99], and open polyurethane foams with starch powder [100], have been proved be efficient in the Candida albicans growth inhibition.

As in the other nanoparticles, the green synthesis has captured the investigators’ attention. Its methodology does not produce toxic subproduct in the nanoparticle production, and it covers the nanoparticles avoiding the agglomeration [31, 84, 87].

Extracts of garlic and ginger have been used for the oxidation of copper. Garlic produces the smallest CuONPs, with a size around 14.62–22.80 nm. However, ginger shows a big quantity of very small nanoparticles. Candida albicans has shown major inhibition when exposed to ginger ethanol extract NPs [101]. Few reports about the use of biological extract for the synthesis of CuONPs that inhibited Candida albicans were founded (Table 3).

Inhibition test NPs size (nm) Synthesis method Capped Reference
15.0 c ± 0.57735 (inhibition zone with 62 μg/mL)/1.93 ± 0.76376 (MIC) 31.7 Alginate NR El-Batal et al. 2018 [102]
10.5 (zone inhibition) 26 ± 4 Acalypha indica NR Sivaraj et al. 2014 [103]
22.5 mm (inhibition zone) Nanostructures
~8000
Viburnum opulus Cu3(PO4)2
3H2O
Ildiz et al. 2017 [104]
20 mm (inhibition zone with 50 μg/mL) NR Sargassum polycystum NR Vishnu et al. 2016 [105]
0.15 Log10 growth reduction 22 ± 1 nm Cellulose, chitosan, and keratin nanocomposites Cellulose, chitosan, and keratin Tram et al., 2017 [106]

Table 3.

Biogenic synthesized CuONPs against Candida albicans.

Cu2+ ions realized from CuONPs are the principal toxicity mechanisms reported. However, it is not the only one. Some authors reported the CuONP accumulation in the cell wall as well as the oxidative stress increase. However, more studies are necessary in order to understand the NPs toxicity mechanisms.

7.4. Other nanoparticles

Other metal nanoparticles as gold [AuNPs, [36]], palladium [PdNPs, [91]], and selenium [SeNPs, [92]], among others, have been proven. The microorganisms’ elimination by all these NPs has been proved to be effective.

Chemical reduction of palladium salt has produced PdNPs in an average size of 9 nm (±3.9 nm). The PdNPs showed a significative Candida albicans growth reduction at 150 ppm [107].

SeNPs were prepared by the reduction of selenium chloride inside K. pneumoniae. This synthesis produced SeNPs with a size range from 90 to 320 nm. The concentration necessary for growth inhibition of Candida albicans was 2000 μg/mL [109]. Another strategy for the Candida albicans inhibition induced the SeNPs formation inside of Lactobacillus and then used them as an antifungal. This strategy worked with L. plantarum and L. johnsonii. Candida albicans exposed to both microorganisms becomes more efficient than bacterium without NPs [108].

Schiff base ligand 2-((4,6-dihethoxypyr-imidine-2-yl)methyleneenamino)-6-methoxyphenol has been used in order to capped AuNPs and platinum nanoparticles (PtNPs). Nanoparticles synthetized presented a layer. The NPs showed a granular and spherical shape with size of 38.14 ± 4.5 and 58.64 ± 3.0 nm, respectively. Both nanoparticles showed greater inhibition than amphotericin [110].

Biogenic PdNP production with watery extract of Moringa oleifera flower was evaluated for the Candida albicans inhibition. However, these PdNPs had no effect in Candida albicans [111].

Candida albicans exposition to tellurium NPs produces complete inhibition with of ≥2000 μg/mL. Besides, the results showed a squalene accumulation and increase in the expression levels of the ERG1 gene of squalene monooxygenase enzyme [112].

7.5. Nanoparticle characteristics and toxicity and Candida albicans inhibition growth

The most important NPs characteristics are the size, shape, composition, superficial charge, and hydrodynamic. These characteristics directly influence their capacity to interact with the molecules or cells. Ocsoy et al. (2017) [46] indicate that the size is strictly related to charge density. The smallest of the NPs has the biggest charge density. This explains the big attraction between small NPs and their agglomeration. The NPs agglomeration is a phenomenon that changes their capacity to work as a caring or their toxicokinetic characteristic [33, 107]. In order to improve the distribution and decrease the agglomeration, a surfactant as Pluronic F® has been added to the synthesis process (Figure 1). Other surfactants as CTAB, SDS, cation surfactants [113], or PVP have been proved to be efficient [114].

Figure 1.

Scanning electronic microscopy of PdNPs produced by chemical reduction. (A) Without surfactant. (B) With Pluronic F® (5% W/V).

Another synthesis methodology that improves these characteristics is the use of subproducts from plants, algae, bacteria, fungi, yeast, etc. [115]. The high content of proteins, reducing sugars or anthocyanins, works as reducing agent; additionally, some of these molecules are adhered in the NPs surface [31]. This addition gives different characteristics to the NPs produced by chemical reduction. The NPs obtained with biogenic methods offer homogeneous NPs with antioxidant characteristics that allow the addition of other molecules and control their toxicity characteristics.

The biogenic NPs production has proven to be efficient inhibiting the Candida albicans growth (Tables 13). The differences in the concentration necessary to inhibit their growth are directly related to NPs size, the capping nature, and the microorganisms evaluated [116]. These are evident in the work of Anasane et al. 2016 [52] et al. (2013), where two microorganisms were evaluated in the production of AgNPs, obtaining two NPs sizes. The AgNPs production by the strain SF23 was smaller than that produced by the strain C9. Candida albicans presented an MIC of 30 μg/mL with the small AgNPs (size 4–36 nm). Wypij et al. (2017) [37] reported the AgNPs synthesis with Pilimelia columellifera subsp. pallida SL19, with a capping of proteins. The AgNPs size was on average 12.7 nm with a growth inhibition of 64 μg/mL. Tulsi leaf extract has been used to produce AgNPs with a size of 2–7 nm. The fungicide concentration in the AgNPs produced was of 120 μg/mL [25]. The production of ZnONPs by Azadirachta indica [61], Vitex trifolia [64], and Suaeda aegyptiaca [65], among others (Table 2), has been evaluated. The extract produced different ZnONPs sizes and capping (Table 2) that influence the growth inhibition.

Due to the NPs size (nanometer scale), diverse toxicity mechanisms have been proposed (Figure 2). One is the direct interaction of the nanoparticle with the cell membrane. The interaction of NPs microorganism is allowed by the NPs superficial charge but also the microorganism superficial charge. This interaction generates disruption of the cell wall and the leakage of ions and the intracellular material with the microorganism’s death [117, 118, 119].

Figure 2.

NPs toxicity mechanisms induced in Candida albicans. (1) NPs adhesion in the cell membrane causing disruption in the membrane and death. (2) NPs introduction to the cell with their posterior degradation and ion release. (3) NPs ROS generation outside the cell. (4) NPs and ions ROS generation in interaction with mitochondria. (6) Membrane lipid peroxidation with nucleus damage. (7) DNA damage by NPS and ions ROS generation.

Consequent to NPs dilution, the ion interaction with the cell is another toxicity mechanism. The ions are incorporated into the cell, and they could interact with thiol groups of proteins and enzymes leading to the inhibition of crucial biological activities [82]. Also, the NPs internalization has been reported; this generates the interaction of NPs and ions with molecules of biological importance as DNA or enzymes [33, 120]. Inside the cell, the NPs or the ions provoke Fenton’s type reaction. Due to this, the oxidative stress increases, and the proteins, lipids, and DNA release (Figure 2).

Khan et al. (2014) [121] write “when CoFe2O4P NCs gained a higher energy than Eg, the electrons (e) of CoFe2O4P NCs were promoted across the band gap to the conduction band (Ec), which creates a hole (h+) in the valence band (Ec). These e in the EC and holes in the Ev possibly have high reducing and oxidizing powers, respectively.” These electron movements generate the NPs reactivity and the oxidative stress generation. The same phenomenon is reported by Setyawati et al. (2014) [122] for TiONPs. These NPs adsorbed UV light, and the adsorbed UV light generates electron excitation, creating a hole in the valence band. The electrons and holes migrate to NPs superficies where they can react with oxygen or water generating oxidative stress. The superficial e generates the liberation of hydroxyl increasing the mitochondria membrane depolarization and liberation of cytochrome c. Added to this, the results show DNA damage [120]. However, the oxidative stress is not the only mechanisms involved in the AgNPs toxicity; also the membrane fluidity, ergosterol content, and cellular and ultrastructure morphology are altered [123]. The squalene monooxygenase expression in Candida albicans exposed to TeNPs showed an increase [112]. The squalene monooxygenase catalyzes the conversion of squalene to 2,3-oxidosqualene; inhibition in this step generates the squalene increase and the disruption of the Candida albicans cell wall [107].

Advertisement

8. Conclusion

The antibiotic resistance increase in many microorganisms has encouraged researchers to focus their efforts on the synthesis and design of multiple effective compounds in order to combat the resistance mechanism of microorganisms. NPs offer a new and effective “antibiotic” that could work against the resistance mechanism in Candida albicans.

The biogenic NPs offer a new generation of NPs that work as a carrier. These NPs offer the possibility to drive the toxicity effect to a specific target, in such way the collateral damage could be diminished.

However, the NPs toxicity mechanisms are not completely understood. The oxidative stress is one of the described mechanisms; it is possible to investigate other mechanisms involving similarity to oxidative stress. More studies are necessary to understand the influence that nanoparticles have in the Candida albicans resistance mechanisms and if this exposition could not generate new ones.

Advertisement

Acknowledgments

This work was partially financed by F-PROMEP-38/REV-04. SEP-23-005 and ECOS NORD 263456.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Simon J, Sun HY, Leong HN, Barez MYC, Huang PY, Talwar D, et al. Echinocandins in invasive candidiasis. Mycoses. 2013;56:601-609. DOI: 10.1111/myc.12085
  2. 2. Dadar M, Tiwari R, Karthik K, Chakraborty S, Shahali Y, Dhama K. Candida albicans—Biology, molecular characterization, pathogenicity, and advances in diagnosis and control—An update. Microbial Pathogenesis. 2018;117:128-138. DOI: 10.1016/j.micpath.2018.02.028
  3. 3. Gonçalves SS, Souza ACR, Chowdhary A, Meis JF, Colombo AL. Epidemiology and molecular mechanisms of antifungal resistance in Candida and Aspergillus. Mycoses. 2016;59:198-219. DOI: 10.1111/myc.12469
  4. 4. de Bedout C, Gómez BL. Candida y candidiasis invasora: Un reto continuo para su diagnóstico temprano. Infection. 2010;14:159-171. DOI: 10.1016/S0123-9392(10)70133-8
  5. 5. de Sa NP, de Paula LFJ, Lopes LFF, Cruz LIB, Matos TTS, Lino CI, et al. In vivo and in vitro activity of a bis-arylidenecyclo-alkanone against fluconazole-susceptible and resistant isolates of Candida albicans. The Journal of Global Antimicrobial Resistance; 2018. DOI: 10.1016/j.jgar.2018.04.012
  6. 6. De la Torre-Saldaña VA, Mayté MV, et al. Factores de riesgo y epidemiología de la candidemia en el Hospital Juárez de México. Medicina Interna de México. 2014;30:121-132
  7. 7. El-Baz AF, El-Batal AI, Abomosalam FM, Tayel AA, Shetaia YM, Yang ST. Extracellular biosynthesis of anti-Candida silver nanoparticles using Monascus purpureus. Journal of Basic Microbiology. 2016;56:531-540. DOI: 10.1002/jobm.201500503
  8. 8. Salci TP, Negri M, Abadio AKR, Svidzinski TIE, Kioshima ÉS. Targeting Candida spp. to develop antifungal agents. Drug Discovery Today. 2018;23:802-814. DOI: 10.1016/j.drudis.2018.01.003
  9. 9. Zhou S, Sun Z, Ye Z, Wang Y, Wang L, Xing L, et al. In vitro photodynamic inactivation effects of benzylidene cyclopentanone photosensitizers on clinical fluconazole-resistant Candida albicans. Photodiagnosis and Photodynamic Therapy. 2018;22:178-186. DOI: 10.1016/j.pdpdt.2018.04.001
  10. 10. Caplan T, Polvi EJ, Xie JL, Buckhalter S, Leach MD, Robbins N, et al. Functional genomic screening reveals core modulators of echinocandin stress responses in Candida albicans. Cell Reports. 2018;23:2292-2298. DOI: 10.1016/j.celrep.2018.04.084
  11. 11. Mane A, Vidhate P, Kusro C, Waman V, Saxena V, Kulkarni-Kale U, et al. Molecular mechanisms associated with Fluconazole resistance in clinical Candida albicans isolates from India. Mycoses. 2016;59:93-100. DOI: 10.1111/myc.12439
  12. 12. Arenas R. Micología Médica Ilustrada3era edici. McGraw-Hill; 2014
  13. 13. Choi J, Kim S-H. A genome tree of life for the fungi kingdom. Proceedings of the National Academy of Sciences of the United States of America. 2017;114:201711939. DOI: 10.1073/pnas.1711939114
  14. 14. Chaffin WL. Candida albicans cell wall proteins. Microbiology and Molecular Biology Reviews. 2008;72:495-544. DOI: 10.1128/MMBR.00032-07
  15. 15. Gow NAR, Hube B. Importance of the Candida albicans cell wall during commensalism and infection. Current Opinion in Microbiology. 2012;15:406-412. DOI: 10.1016/j.mib.2012.04.005
  16. 16. Ruiz-Herrera J, Victoria Elorza M, Valentín E, Sentandreu R. Molecular organization of the cell wall of Candida albicans and its relation to pathogenicity. FEMS Yeast Research. 2006;6:14-29. DOI: 10.1111/j.1567-1364.2005.00017.x
  17. 17. Chaffin WL, López-ribot JL, Casanova M, Gozalbo D, Martínez JP, Lo L. Cell wall and secreted proteins of Candida albicans: Identification, function, and expression. Microbiology and Molecular Biology Reviews. 1998;62:130-180
  18. 18. Lipke PN, Ovalle R. Cell wall architecture in yeast: New structure and new challenges. Journal of Bacteriology. 1998;180:3735-3740
  19. 19. Ramos AP, Desgarennes CP. Artículo de revisión Factores de virulencia en; 2005. pp. 12-27
  20. 20. Spreghini E, Davis DA, Subaran R, Kim M, Mitchell AP. Roles of Candida albicans Dfg5p and Dcw1p cell surface proteins in growth and hypha formation. Eukaryotic Cell. 2003;2:746-755. DOI: 10.1128/EC.2.4.746-755.2003
  21. 21. Calderone RA, Fonzi WA. Virulence factors of Candida Albicans. Trends in Microbiology. 2001;9:327-335
  22. 22. Younes S, Bahnan W, Dimassi HI, Khalaf RA. The Candida albicans Hwp2 is necessary for proper adhesion, biofilm formation and oxidative stress tolerance. Microbiological Research. 2011;166:430-436. DOI: 10.1016/j.micres.2010.08.004
  23. 23. Kruppa M. Quorum sensing and Candida albicans. Mycoses. 2009;52:1-10. DOI: 10.1111/j.1439-0507.2008.01626.x
  24. 24. Kadosh D, Lopez-Ribot JL. Candida albicans: Adapting to succeed. Cell Host & Microbe. 2013;14:483-485. DOI: 10.1016/j.chom.2013.10.016
  25. 25. Ramage G, Vandewalle K, Wickes BL, López-Ribot JL. Characteristics of biofilm formation by Candida albicans (xtt). Revista Iberoamericana de Micología. 2001;18:163-170 DOI: 200118163 [pii]
  26. 26. Coleman JJ, Okoli I, Tegos GP, Holso EB, Wagner FF, Hamblin MR, et al. Characterization of plant-derived Saponin natural products against Candida albicans. ACS Chemical Biology. 2010;5:321-332
  27. 27. Monteiro DR, Silva S, Negri M, Gorup LF, de Camargo ER, Oliveira R, et al. Silver colloidal nanoparticles: Effect on matrix composition and structure of Candida albicans and Candida glabrata biofilms. Journal of Applied Microbiology. 2013;114:1175-1183. DOI: 10.1111/jam.12102
  28. 28. Monteiro DR, Takamiya AS, Feresin LP, Gorup LF, de Camargo ER, Delbem ACB, et al. Susceptibility of Candida albicans and Candida glabrata biofilms to silver nanoparticles in intermediate and mature development phases. Journal of Prosthodontic Research. 2015;59:42-48. DOI: 10.1016/j.jpor.2014.07.004
  29. 29. Salari S, Khosravi AR, Mousavi SAA, Nikbakht-Brojeni GH. Mechanisms of resistance to fluconazole in Candida albicans clinical isolates from Iranian HIV-infected patients with oropharyngeal candidiasis. The Journal de Mycologie Médicale. 2016;26:35-41. DOI: 10.1016/j.mycmed.2015.10.007
  30. 30. Maebashi K, Niimi M, Kudoh M, Fischer FJ, Makimura K, Niimi K, et al. Mechanisms of fluconazole resistance in Candida albicans isolates from Japanese AIDS patients. The Journal of Antimicrobial Chemotherapy. 2001;47:527-536. DOI: 10.1093/jac/47.5.527
  31. 31. Saratale RG, Saratale GD, Shin HS, Jacob JM, Pugazhendhi A, Bhaisare M, et al. New insights on the green synthesis of metallic nanoparticles using plant and waste biomaterials: Current knowledge, their agricultural and environmental applications. Environmental Science and Pollution Research. 2018;25:10164-10183. DOI: 10.1007/s11356-017-9912-6
  32. 32. Carpenter AW, Slomberg DL, Rao KS, Schoenfisch MH. Influence of scaffold size on bactericidal activity of nitric oxide-releasing silica nanoparticles. ASC Nano. 2011;5:7235-7244. DOI: 10.1111/j.1743-6109.2008.01122.x.Endothelial
  33. 33. Carlos A-Q, Blanca S-R, Roco I-R, Amelia P-CH, Lorena M-FS, Alberto D-M, et al. Toxicity effects in pathogen microorganisms exposed to silver nanoparticles. Nanoscience and Nanotechnology Letters. 2017;9:165-173. DOI: 10.1166/nnl.2017.2275
  34. 34. Oves M, Aslam M, Rauf MA, Qayyum S, Qari HA, Khan MS, et al. Antimicrobial and anticancer activities of silver nanoparticles synthesized from the root hair extract of Phoenix dactylifera. Materials Science and Engineering: C. 2018;89:429-443. DOI: 10.1016/j.msec.2018.03.035
  35. 35. Aazam ES, Zaheer Z. Growth of Ag-nanoparticles in an aqueous solution and their antimicrobial activities against Gram positive, Gram negative bacterial strains and Candida fungus. Bioprocess and Biosystems Engineering. 2016;39:575-584. DOI: 10.1007/s00449-016-1539-3
  36. 36. Shankar PD, Shobana S, Karuppusamy I, Pugazhendhi A, Ramkumar VS, Arvindnarayan S, et al. A review on the biosynthesis of metallic nanoparticles (gold and silver) using bio-components of microalgae: Formation mechanism and applications. Enzyme and Microbial Technology. 2016;95:28-44. DOI: 10.1016/j.enzmictec.2016.10.015
  37. 37. Wypij M, Czarnecka J, Dahm H, Rai M, Golinska P. Silver nanoparticles from Pilimelia columellifera subsp. pallida SL19 strain demonstrated antifungal activity against fungi causing superficial mycoses. Journal of Basic Microbiology. 2017;57:793-800. DOI: 10.1002/jobm.201700121
  38. 38. Lateef A, Ojo SA, Oladejo SM. Anti-candida, anti-coagulant and thrombolytic activities of biosynthesized silver nanoparticles using cell-free extract of Bacillus safensis LAU 13. Process Biochemistry. 2016;51:1406-1412. DOI: 10.1016/j.procbio.2016.06.027
  39. 39. Ghiuță I, Cristea D, Croitoru C, Kost J, Wenkert R, Vyrides I, et al. Characterization and antimicrobial activity of silver nanoparticles, biosynthesized using Bacillus species. Applied Surface Science. 2018;438:66-73. DOI: 10.1016/j.apsusc.2017.09.163
  40. 40. Musa SF, Yeat TS, Kamal LZM, Tabana YM, Ahmed MA, El Ouweini A, et al. Pleurotus sajor-caju can be used to synthesize silver nanoparticles with antifungal activity against Candida albicans. Journal of the Science of Food and Agriculture. 2018;98:1197-1207. DOI: 10.1002/jsfa.8573
  41. 41. Khatoon N, Mishra A, Alam H, Manzoor N, Sardar M. Biosynthesis, characterization, and antifungal activity of the silver nanoparticles against pathogenic candida species. BioNanoScience. 2015;5:65-74. DOI: 10.1007/s12668-015-0163-z
  42. 42. Moteriya P, Padalia H, Chanda S. Characterization, synergistic antibacterial and free radical scavenging efficacy of silver nanoparticles synthesized using Cassia roxburghii leaf extract. Journal, Genetic Engineering & Biotechnology. 2017;15:505-513. DOI: 10.1016/j.jgeb.2017.06.010
  43. 43. Muthamil S, Devi VA, Balasubramaniam B, Balamurugan K, Pandian SK. Green synthesized silver nanoparticles demonstrating enhanced in vitro and in vivo antibiofilm activity against Candida spp. Journal of Basic Microbiology. 2018;58:343-357. DOI: 10.1002/jobm.201700529
  44. 44. Kumar P, Senthamil Selvi S, Govindaraju M. Seaweed-mediated biosynthesis of silver nanoparticles using Gracilaria corticata for its antifungal activity against Candida spp. Applied Nanoscience. 2013;3:495-500. DOI: 10.1007/s13204-012-0151-3
  45. 45. Khatoon UT, Nageswara Rao GVS, Mohan KM, Ramanaviciene A, Ramanavicius A. Antibacterial and antifungal activity of silver nanospheres synthesized by tri-sodium citrate assisted chemical approach. Vacuum. 2017;146:259-265. DOI: 10.1016/j.vacuum.2017.10.003
  46. 46. Ocsoy I, Temiz M, Celik C, Altinsoy B, Yilmaz V, Duman F. A green approach for formation of silver nanoparticles on magnetic graphene oxide and highly effective antimicrobial activity and reusability. Journal of Molecular Liquids. 2017;227:147-152. DOI: 10.1016/j.molliq.2016.12.015
  47. 47. Ashajyothi C, Prabhurajeshwar C, Handral HK, Kelmani CR. Investigation of antifungal and anti-mycelium activities using biogenic nanoparticles: An eco-friendly approach. Environmental Nanotechnology, Monitoring and Management. 2016;5:81-87. DOI: 10.1016/j.enmm.2016.04.002
  48. 48. Rahisuddin Al-Thabaiti SA, Khan Z, Manzoor N. Biosynthesis of silver nanoparticles and its antibacterial and antifungal activities towards Gram-positive, Gram-negative bacterial strains and different species of Candida fungus. Bioprocess and Biosystems Engineering. 2015;38:1773-1781. DOI: 10.1007/s00449-015-1418-3
  49. 49. Parlinska-wojtan M, Depciuch J, Fryc B, Kus-liskiewicz M. Green synthesis and antibacterial effects of aqueous colloidal solutions of silver nanoparticles using clove eugenol. Applied Organometallic Chemistry. 2018;32:1-9. DOI: 10.1002/aoc.4276
  50. 50. Padalia H, Moteriya P, Chanda S. Green synthesis of silver nanoparticles from marigold flower and its synergistic antimicrobial potential. Arabian Journal of Chemistry. 2015;8:732-741. DOI: 10.1016/j.arabjc.2014.11.015
  51. 51. Kumar S, Bhattacharya W, Singh M, Halder D, Mitra A. Plant latex capped colloidal silver nanoparticles: A potent anti-biofilm and fungicidal formulation. Journal of Molecular Liquids. 2017;230:705-713. DOI: 10.1016/j.molliq.2017.01.004
  52. 52. Anasane N, Golińska P, Wypij M, Rathod D, Dahm H, Rai M. Acidophilic actinobacteria synthesised silver nanoparticles showed remarkable activity against fungi-causing superficial mycoses in humans. Mycoses. 2016;59:157-166. DOI: 10.1111/myc.12445
  53. 53. Hamid S, Zainab S, Faryal R, Ali N. Deterrence in metabolic and biofilms forming activity of Candida species by mycogenic silver nanoparticles. Journal of Applied Biomedicine. 2017;15:249-255. DOI: 10.1016/j.jab.2017.02.003
  54. 54. Monteiro DR, Silva S, Negri M, Gorup LF, de Camargo ER, Oliveira R, et al. Antifungal activity of silver nanoparticles in combination with nystatin and chlorhexidine digluconate against Candida albicans and Candida glabrata biofilms. Mycoses. 2013;56:672-680. DOI: 10.1111/myc.12093
  55. 55. Lara HH, Romero-Urbina DG, Pierce C, Lopez-Ribot JL, Arellano-Jiménez MJ, Jose-Yacaman M. Effect of silver nanoparticles on Candida albicans biofilms: An ultrastructural study. Journal of Nanobiotechnology. 2015;13:1-12. DOI: 10.1186/s12951-015-0147-8
  56. 56. Kumar CG, Poornachandra Y. Biodirected synthesis of Miconazole-conjugated bacterial silver nanoparticles and their application as antifungal agents and drug delivery vehicles. Colloids and Surfaces B. 2015;125:110-119. DOI: 10.1016/j.colsurfb.2014.11.025
  57. 57. Peña-González CE, Pedziwiatr-Werbicka E, Martín-Pérez T, Szewczyk EM, Copa-Patiño JL, Soliveri J, et al. Antibacterial and antifungal properties of dendronized silver and gold nanoparticles with cationic carbosilane dendrons. International Journal of Pharmaceutics. 2017;528:55-61. DOI: 10.1016/j.ijpharm.2017.05.067
  58. 58. Selvaraj M, Pandurangan P, Ramasami N, Rajendran SB, Sangilimuthu SN, Perumal P. Highly potential antifungal activity of quantum-sized silver nanoparticles against Candida albicans. Applied Biochemistry and Biotechnology. 2014;173:55-66. DOI: 10.1007/s12010-014-0782-9
  59. 59. Li D, Liu Z, Yuan Y, Liu Y, Niu F. Green synthesis of gallic acid-coated silver nanoparticles with high antimicrobial activity and low cytotoxicity to normal cells. Process Biochemistry. 2015;50:357-366. DOI: 10.1016/j.procbio.2015.01.002
  60. 60. Rajput VD, Minkina TM, Behal A, Sushkova SN, Mandzhieva S, Singh R, et al. Effects of zinc-oxide nanoparticles on soil, plants, animals and soil organisms: A review. Environmental Nanotechnology, Monitoring and Management. 2018;9:76-84
  61. 61. Mirzaeia Hamed DM. Zinc oxide nanoparticles: Biological synthesis and biomedical applications. Ceramics International. 2013;43:1-11. DOI: 10.1016/j.ijnurstu.2008.11.006
  62. 62. Chinnasamy C, Tamilselvam P, Karthick B, Sidharth B, Senthilnathan M. Green synthesis, characterization and optimization studies of zinc oxide nano particles using costusigneus leaf extract. Materials Today: Proceedings. 2018;5:6728-6735. DOI: 10.1016/j.matpr.2017.11.331
  63. 63. Ahmed S, Chaudhry SA, Ikram S. A review on biogenic synthesis of ZnO nanoparticles using plant extracts and microbes: A prospect towards green chemistry. Journal of Photochemistry and Photobiology B: Biology. 2017;166:272-284. DOI: 10.1016/j.jphotobiol.2016.12.011
  64. 64. Hou J, Wu Y, Li X, Wei B, Li S, Wang X. Toxic effects of different types of zinc oxide nanoparticles on algae, plants, invertebrates, vertebrates and microorganisms. Chemosphere. 2018;193:852-860. DOI: 10.1016/j.chemosphere.2017.11.077
  65. 65. Madhumitha G, Elango G, Roopan SM. Biotechnological aspects of ZnO nanoparticles: Overview on synthesis and its applications. Applied Microbiology and Biotechnology. 2016;100:571-581. DOI: 10.1007/s00253-015-7108-x
  66. 66. Ahmed S, Annu SAC, Ikram S. A review on biogenic synthesis of ZnO nanoparticles using plant extracts and microbes: A prospect towards green chemistry. Journal of Photochemistry and Photobiology B: Biology. 2017;166:272-284. DOI: 10.1016/j.jphotobiol.2016.12.011
  67. 67. Vijayakumar S, Mahadevan S, Arulmozhi P, Sriram S, Praseetha PK. Green synthesis of zinc oxide nanoparticles using Atalantia monophylla leaf extracts: Characterization and antimicrobial analysis. Materials Science in Semiconductor Processing. 2018;82:39-45. DOI: 10.1016/j.mssp.2018.03.017
  68. 68. Rajabi HR, Naghiha R, Kheirizadeh M, Sadatfaraji H, Mirzaei A, Alvand ZM. Microwave assisted extraction as an efficient approach for biosynthesis of zinc oxide nanoparticles: Synthesis, characterization, and biological properties. Materials Science and Engineering: C. 2017;78:1109-1118. DOI: 10.1016/j.msec.2017.03.090
  69. 69. Dobrucka R, Dlugaszewska J, Kaczmarek M. Cytotoxic and antimicrobial effects of biosynthesized ZnO nanoparticles using of Chelidonium majus extract. Biomedical Microdevices. 2018;20. DOI: 10.1007/s10544-017-0233-9
  70. 70. Vijayakumar S, Krishnakumar C, Arulmozhi P, Mahadevan S, Parameswari N. Biosynthesis, characterization and antimicrobial activities of zinc oxide nanoparticles from leaf extract of Glycosmis pentaphylla (Retz.) DC. Microbial Pathogenesis. 2018;116:44-48. DOI: 10.1016/j.micpath.2018.01.003
  71. 71. Elumalai K, Velmurugan S. Green synthesis, characterization and antimicrobial activities of zinc oxide nanoparticles from the leaf extract of Azadirachta indica (L.). Applied Surface Science. 2015;345:329-336. DOI: 10.1016/j.apsusc.2015.03.176
  72. 72. Nagarajan S, Kuppusamy KA. Extracellular synthesis of zinc oxide nanoparticle using seaweeds of gulf of Mannar, India. Journal of Nanobiotechnology. 2013;11:39. DOI: 10.1186/1477-3155-11-39
  73. 73. Padalia H, Moteriya P, Chanda S. Synergistic antimicrobial and cytotoxic potential of zinc oxide nanoparticles synthesized using Cassia auriculata leaf extract. BioNanoScience. 2018:196-206
  74. 74. Dobrucka R, Długaszewska J. Biosynthesis and antibacterial activity of ZnO nanoparticles using Trifolium pratense flower extract. Saudi Journal of Biological Sciences. 2015:517-523. DOI: 10.1016/j.sjbs.2015.05.016
  75. 75. Dhillon GS, Kaur S, Brar SK. Facile fabrication and characterization of chitosan-based zinc oxide nanoparticles and evaluation of their antimicrobial and antibiofilm activity. International Nano Letters. 2014;4:107. DOI: 10.1007/s40089-014-0107-6
  76. 76. Divya M, Vaseeharan B, Abinaya M, Vijayakumar S, Govindarajan M, Alharbi NS, et al. Biology biopolymer gelatin-coated zinc oxide nanoparticles showed high antibacterial, antibiofilm and anti-angiogenic activity. The Journal of Photochemistry and Photobiology B: Biology. 2018;178:211-218. DOI: 10.1016/j.jphotobiol.2017.11.008
  77. 77. Serkhacheva NS, Yashina NV, Prokopov NI, Gaynanova AA, Kuz’micheva GM, Domoroshchina EN, et al. Bactericidal properties of nanoscale zinc(II) and titanium (IV) oxides of different nature and their nanocomposites with polystyrene. Nanotechnologies in Russia. 2016;11:99-109. DOI: 10.1134/S1995078016010146
  78. 78. Vijayakumar S, Vaseeharan B. Antibiofilm, anti cancer and ecotoxicity properties of collagen based ZnO nanoparticles. Advanced Powder Technology. 2018;29:2331-2345
  79. 79. Shoeb M, Singh BR, Khan JA, Khan W, Singh BN, Singh HB, et al. ROS-dependent anticandidal activity of zinc oxide nanoparticles synthesized by using egg albumen as a biotemplate. Advances in Natural Sciences: Nanoscience and Nanotechnology. 2013;4:1-12. DOI: 10.1088/2043-6262/4/3/035015
  80. 80. Mizielińska M, Łopusiewicz Ł, Mȩżyńska M, Bartkowiak A. The influence of accelerated UV-A and Q-sun irradiation on the antimicrobial properties of coatings containing ZnO nanoparticles. Molecules. 2017;22:12-15. DOI: 10.3390/molecules22091556
  81. 81. Agarwal H, Menon S, Venkat Kumar S, Rajeshkumar S. Mechanistic study on antibacterial action of zinc oxide nanoparticles synthesized using green route. Chemico-Biological Interactions. 2018;286:60-70. DOI: 10.1016/J.CBI.2018.03.008
  82. 82. Longano D, Ditaranto N, Sabbatini L, Torsi L. In: Nicola Coffi MR, editor. Nano Antimicrobials. 2012. pp. 85-117
  83. 83. Ingle AP, Duran N, Rai M. Bioactivity, mechanism of action, and cytotoxicity of copper-based nanoparticles: A review. Applied Microbiology and Biotechnology. 2014;98:1001-1009. DOI: 10.1007/s00253-013-5422-8
  84. 84. Natan M, Banin E. From Nano to Micro: Using nanotechnology to combat microorganisms and their multidrug resistance. FEMS Microbiology Reviews. 2017;41:302-322. DOI: 10.1093/femsre/fux003
  85. 85. Bogdanovic U, Lazic V, Vodnik V, Budimir M, Markovic Z, Dimitrijevic S. Copper nanoparticles with high antimicrobial activity. Materials Letters. 2014;128:75-78. DOI: 10.1016/j.matlet.2014.04.106
  86. 86. Katwal R, Kaur H, Sharma G, Naushad M, Pathania D. Electrochemical synthesized copper oxide nanoparticles for enhanced photocatalytic and antimicrobial activity. Journal of Industrial and Engineering Chemistry. 2015;31:173-184
  87. 87. Din MI, Arshad F, Hussain Z, Mukhtar M. Green adeptness in the synthesis and stabilization of copper nanoparticles: Catalytic, antibacterial, cytotoxicity, and antioxidant activities. Nanoscale Research Letters. 2017;12:638. DOI: 10.1186/s11671-017-2399-8
  88. 88. Tamayo L, Azócar M, Kogan M, Riveros A, Páez M. Copper-polymer nanocomposites: An excellent and cost-effective biocide for use on antibacterial surfaces. Materials Science and Engineering: C. 2016;69:1391-1409. DOI: 10.1016/j.msec.2016.08.041
  89. 89. Weitz IS, Maoz M, Panitz D, Eichler S, Segal E. Combination of CuO nanoparticles and fluconazole: Preparation, characterization, and antifungal activity against Candida albicans. Journal of Nanoparticle Research. 2015;17. DOI: 10.1007/s11051-015-3149-4
  90. 90. Ramyadevi J, Jeyasubramanian K, Marikani A, Rajakumar G, Rahuman AA. Synthesis and antimicrobial activity of copper nanoparticles. Materials Letters. 2012;71:114-116. DOI: 10.1016/j.matlet.2011.12.055
  91. 91. Amiri M, Etemadifar Z, Daneshkazemi A, Nateghi M. Antimicrobial effect of copper oxide nanoparticles on some oral bacteria and candida species. Journal of Dental Biomaterials. 2017;4:347-352. DOI: 28959764s
  92. 92. Adwin Jose P, Dhaveethu Raja J, Sankarganesh M, Rajesh J. Evaluation of antioxidant, DNA targeting, antimicrobial and cytotoxic studies of imine capped copper and nickel nanoparticles. Journal of Photochemistry and Photobiology B: Biology. 2018;178:143-151. DOI: 10.1016/j.jphotobiol.2017.11.005
  93. 93. Pugazhendhi A, Kumar SS, Manikandan M, Saravanan M. Photocatalytic properties and antimicrobial efficacy of Fe doped CuO nanoparticles against the pathogenic bacteria and fungi. Microbial Pathogenesis. 2018;122:84-89
  94. 94. Arunadevi R, Kavitha B, Rajarajan M, Suganthi A, Jeyamurugan A. Investigation of the drastic improvement of photocatalytic degradation of Congo red by monoclinic Cd, Ba-CuO nanoparticles and its antimicrobial activities. Surfaces and Interfaces. 2018;10:32-44. DOI: 10.1016/j.surfin.2017.11.004
  95. 95. Eremenko AM, Petrik IS, Smirnova NP, Rudenko AV, Marikvas YS. Antibacterial and antimycotic activity of cotton fabrics, impregnated with silver and binary silver/copper nanoparticles. Nanoscale Research Letters. 2016;11:1-9. DOI: 10.1186/s11671-016-1240-0
  96. 96. El-nahhal IM, Elmanama AA, Amara N, Qodih FS, Selmane M, Chehimi MM. The efficacy of surfactants in stabilizing coating of nano-structured CuO particles onto the surface of cotton fibers and their antimicrobial activity. Materials Chemistry and Physics. 2018;215:221-228
  97. 97. Savelyev Y, Gonchar A, Movchan B, Gornostay A, Vozianov S, Rudenko A, et al. Antibacterial polyurethane materials with silver and copper nanoparticles. Materials Today: Proceedings. 2017;4:87-94. DOI: 10.1016/j.matpr.2017.01.196
  98. 98. Ballo MKS, Rtimi S, Kiwi J, Pulgarin C, Entenza JM, Bizzini A. Fungicidal activity of copper-sputtered flexible surfaces under dark and actinic light against azole-resistant Candida albicans and Candida glabrata. Journal of Photochemistry and Photobiology B: Biology. 2017;174:229-234. DOI: 10.1016/j.jphotobiol.2017.07.030
  99. 99. Khan J, Siddiq M, Akram B, Ashraf MA. In-situ synthesis of CuO nanoparticles in P(NIPAM-co-AAA) microgel, structural characterization, catalytic and biological applications. Arabian Journal of Chemistry. 2018;11:897-909. DOI: 10.1016/j.arabjc.2017.12.018
  100. 100. Reza H, Saeed M, Dorraji S, Fakhrzadeh V, Eslami H. Starch-based polyurethane/CuO nanocomposite foam: Antibacterial effects for infection control. International Journal of Biological Macromolecules. 2018;111:1076-1082
  101. 101. El AA, Gehan R, Ayman AG, El Khateeb Y, Hassaan MM. Eco-friendly synthesis of metal nanoparticles using ginger and garlic extracts as biocompatible novel antioxidant and antimicrobial agents. Journal of Nanostructure in Chemistry. 2018;8:71-81. DOI: 10.1007/s40097-018-0255-8
  102. 102. El-Batal AI, Al-Hazmi NE, Mosallam FM, El-Sayyad GS. Biogenic synthesis of copper nanoparticles by natural polysaccharides and Pleurotus ostreatus fermented fenugreek using gamma rays with antioxidant and antimicrobial potential towards some wound pathogens. Microbial Pathogenesis. 2018;118:159-169. DOI: 10.1016/j.micpath.2018.03.013
  103. 103. Sivaraj R, Rahman PKSM, Rajiv P, Narendhran S, Venckatesh R. Biosynthesis and characterization of Acalypha indica mediated copper oxide nanoparticles and evaluation of its antimicrobial and anticancer activity. Spectrochimica Acta Part A. 2014;129:25-258. DOI: 10.1016/j.saa.2014.03.027
  104. 104. Nilay Ildiz IO, Baldemir A, Altinkaynak C, Özdemir N, Yilmazc V. Self assembled snowball-like hybrid nanostructures comprising Viburnum opulus L. extract and metal ions for antimicrobial and catalytic applications. Enzyme and Microbial Technology. 2017;102:60-66
  105. 105. Vishnu SRI, Ramaswamy P, Narendhran S, Sivaraj R. Potentiating effect of ecofriendly synthesis of copper oxide nanoparticles using brown alga: Antimicrobial and anticancer activities. Bulletin of Materials Science. 2016;39:361-364. DOI: 10.1007/s12034-016-1173-3
  106. 106. Tran CD, Makuvaza J, Munson E, Bennett B. Biocompatible copper oxide nanoparticle composites from cellulose and chitosan: Facile synthesis, unique structure, and antimicrobial activity. ACS Applied Materials & Interfaces. 2017;9:42503-42515. DOI: 10.1021/acsami.7b11969
  107. 107. Athie-García MS, Piñón-Castillo HA, Muñoz-Castellanos LN, Ulloa-Ogaz AL, Martínez-Varela PI, Quintero-Ramos A, et al. Cell wall damage and oxidative stress in Candida albicans ATCC10231 and Aspergillus niger caused by palladium nanoparticles. Toxicology In Vitro. 2018;48:111-120. DOI: 10.1016/j.tiv.2018.01.006
  108. 108. Kazempour ZB, Yazdi MH, Rafii F, Shahverdi AR. Sub-inhibitory concentration of biogenic selenium nanoparticles lacks post antifungal effect for Aspergillus niger and Candida albicans and stimulates the growth of Aspergillus niger. Iranian Journal of Microbiology. 2013;5:81-85
  109. 109. Kheradmand E, Rafii F, Yazdi MH, Sepahi AA. The antimicrobial effects of selenium nanoparticle-enriched probiotics and their fermented broth against Candida albicans. Daru. 2014;22:1-6. DOI: 10.1186/2008-2231-22-48
  110. 110. Sankarganesh M, Adwin Jose P, Dhaveethu Raja J, Kesavan MP, Vadivel M, Rajesh J, et al. New pyrimidine based ligand capped gold and platinum nano particles: Synthesis, characterization, antimicrobial, antioxidant, DNA interaction and in vitro anticancer activities. Journal of Photochemistry and Photobiology B: Biology. 2017;176:44-53. DOI: 10.1016/j.jphotobiol.2017.09.013
  111. 111. Anand K, Tiloke C, Phulukdaree A, Ranjan B, Chuturgoon A, Singh S, et al. Biology biosynthesis of palladium nanoparticles by using Moringa oleifera flower extract and their catalytic and biological properties. Journal of Photochemistry and Photobiology, B: Biology. 2016;165:87-95. DOI: 10.1016/j.jphotobiol.2016.09.039
  112. 112. Zare B, Sepehrizadeh Z, Faramarzi MA, Soltany-Rezaee-Rad M, Rezaie S, Shahverdi AR. Antifungal activity of biogenic tellurium nanoparticles against Candida albicans and its effects on squalene monooxygenase gene expression. Biotechnology and Applied Biochemistry. 2014;6(1):395-400. DOI: 10.1002/bab.1180
  113. 113. Aiad I, Marzouk MI, Shaker SA, Ebrahim NE, Abd-Elaal AA, Tawfik SM. Antipyrine cationic surfactants capping silver nanoparticles as potent antimicrobial agents against pathogenic bacteria and fungi. Journal of Molecular Liquids. 2017;243:572-583. DOI: 10.1016/j.molliq.2017.08.072
  114. 114. Javed R, Ahmed M, ul Haq I, Nisa S, Zia M. PVP and PEG doped CuO nanoparticles are more biologically active: Antibacterial, antioxidant, antidiabetic and cytotoxic perspective. Materials Science and Engineering: C. 2017;79:108-115. DOI: 10.1016/j.msec.2017.05.006
  115. 115. Durán N, Nakazato G, Seabra AB. Antimicrobial activity of biogenic silver nanoparticles, and silver chloride nanoparticles: An overview and comments. Applied Microbiology and Biotechnology. 2016;100:6555-6570. DOI: 10.1007/s00253-016-7657-7
  116. 116. Monteiro DR, Silva S, Negri M, Gorup LF, De Camargo ER, Oliveira R, et al. Silver nanoparticles: Influence of stabilizing agent and diameter on antifungal activity against Candida albicans and Candida glabrata biofilms. Letters in Applied Microbiology. 2012;54:383-391. DOI: 10.1111/j.1472-765X.2012.03219.x
  117. 117. Kim KJ, Sung WS, Suh BK, Moon SK, Choi JS, Kim JG, et al. Antifungal activity and mode of action of silver nano-particles on Candida albicans. Biometals. 2009;22:235-242. DOI: 10.1007/s10534-008-9159-2
  118. 118. Dos Santos CA, Seckler MM, Ingle AP, Gupta I, Galdiero S, Galdiero M, et al. Silver nanoparticles: Therapeutical uses, toxicity, and safety issues. Journal of Pharmaceutical Sciences. 2014;103:1931-1944. DOI: 10.1002/jps.24001
  119. 119. Qasim M, Singh BR, Naqvi AH, Paik P, Das D. Silver nanoparticles embedded mesoporous SiO2 nanosphere: An effective anticandidal agent against Candida albicans 077. Nanotechnology. 2015;26. DOI: 10.1088/0957-4484/26/28/285102
  120. 120. Hwang IS, Lee J, Hwang JH, Kim KJ, Lee DG. Silver nanoparticles induce apoptotic cell death in Candida albicans through the increase of hydroxyl radicals. The FEBS Journal. 2012;279:1327-1338. DOI: 10.1111/j.1742-4658.2012.08527.x
  121. 121. Alam J, Qasim M, Raj B, Khan W, Das D, Naqvi AH. Polyaniline/CoFe2O4 nanocomposite inhibits the growth of Candida albicans 077 by ROS production. Comptes Rendus Chimie. 2014;17:91-102. DOI: 10.1016/j.crci.2013.08.006
  122. 122. Setyawati MI, Fang W, Chia SL, Leong DT. Nanotoxicology of commonmetal oxide based nanomaterials: their ROS-y and non-ROS-y consequences. Asia-Pacific Journal of Chemical Engineering. 2013;8:205-217. DOI: 10.1002/apj.1680
  123. 123. Radhakrishnan VS, Dwivedi SP, Siddiqui MH, Prasad T. In vitro studies on oxidative stress-independent, Ag nanoparticles-induced cell toxicity of Candida albicans, an opportunistic pathogen. International Journal of Nanomedicine. 2018;13:91-96. DOI: 10.2147/IJN.S125010

Written By

Hilda Amelia Piñón Castillo, Laila Nayzzel Muñoz Castellanos, Rigoberto Martínez Chamorro, Reyna Reyes Martínez and Erasmo Orrantia Borunda

Submitted: 26 April 2018 Reviewed: 18 July 2018 Published: 05 November 2018