Next Article in Journal
Prognostic and Immunological Role of STK38 across Cancers: Friend or Foe?
Previous Article in Journal
Evaluation of Free Light Chains (FLCs) Synthesis in Response to Exposure to SARS-CoV-2
Previous Article in Special Issue
Recent Advances in Prion Inactivation by Plasma Sterilizer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nonthermal Plasma Effects on Fungi: Applications, Fungal Responses, and Future Perspectives

by
Lucia Hoppanová
1,* and
Svetlana Kryštofová
2,*
1
Department of Biophysics and Electrophysiology, Institute of Molecular Physiology and Genetics, Centre of Biosciences, Slovak Academy of Sciences, Dúbravská Cesta 9, 841 04 Bratislava, Slovakia
2
Institute of Biochemistry and Microbiology, Faculty of Chemical and Food Technology, Slovak University of Technology, Radlinského 9, 812 37 Bratislava, Slovakia
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(19), 11592; https://doi.org/10.3390/ijms231911592
Submission received: 31 August 2022 / Revised: 20 September 2022 / Accepted: 26 September 2022 / Published: 30 September 2022
(This article belongs to the Special Issue Plasma Microbiology)

Abstract

:
The kingdom of Fungi is rich in species that live in various environments and exhibit different lifestyles. Many are beneficial and indispensable for the environment and industries, but some can threaten plants, animals, and humans as pathogens. Various strategies have been applied to eliminate fungal pathogens by relying on chemical and nonchemical antifungal agents and tools. Nonthermal plasma (NTP) is a potential tool to inactivate pathogenic and food-contaminating fungi and genetically improve fungal strains used in industry as enzyme and metabolite producers. The NTP mode of action is due to many highly reactive species and their interactions with biological molecules. The interaction of the NTP with living cells is believed to be synergistic yet not well understood. This review aims to summarize the current NTP designs, applications, and challenges that involve fungi, as well as provide brief descriptions of underlying mechanisms employed by fungi in interactions with the NTP components

1. Introduction

The motivation for understanding the effect of nonthermal plasma (NTP) treatment on fungi or other microorganisms stems from the unique and complex nature of plasma and the complexity of processes triggered in the fungal cells upon interaction with the plasma components. NTP in medicine, agriculture, and food processing is used to devitalize and decontaminate various surfaces and liquids. The application of NTP could expand to biotechnology for fungal breeding and antifungal resistance management. Recently, several excellent reviews summarized achievements in the utilization of various types of NTP devices in antifungal treatment [1,2], but very few elaborated in depth on molecular mechanisms triggered by NTP [3]. At the moment, we face a lack of a better understanding of molecular mechanisms and experience some difficulties regarding the methodology influenced by many variables in the experimental setup of the plasma devices, as well as biological differences in fungal species and cell types and biological sample handling. Since plasma has gained significant attention in antifungal treatment in recent years, this review aims to equip readers with the most recent information on NTP compositions and designs, direct and indirect applications, and molecular mechanisms employed by fungi in response to NTP.
The review is organized into several chapters. Section 2 introduces the NTP systems used in the fungal treatments, plasma generation, composition, and biological mechanisms that can be triggered by plasma in fungal cells. It also summarizes biologically active reactive species present in plasma and their effects on fungi. Section 3 provides an overview of plasma applications in medicine, agriculture, food preservation, biotechnology, and the protection of cultural objects.

2. NTP Devices, Plasma Composition, Biologically Active Agents, and Fungal Responses

2.1. Plasma Classification and Configurations of NTP Systems

Plasma, considered to be the fourth state of matter, is a fully or partially ionized gas formed of charged particles (free electrons, positively and negatively charged ions), free radicals, neutral gas particles (excited atoms and molecules), photons (in the visible and UV regions of the spectrum), and the electromagnetic field [4,5,6].
Plasma is a quasi-neutral system of free electrons and ions that exhibit collective behavior. This behavior is reflected in the plasma response to deviations from neutrality and applied external electromagnetic fields and in its ability to sustain many different waveforms and oscillations. Plasma is produced by an ionization process that generates a certain number of free electrons and positive ions. If quasi-neutrality applies, the number density of electrons ne is approximately equal to the ion density of ni, ne ≈ ni, and ne is called the plasma density. Typical plasma density units are cm−3. The plasma density at atmospheric pressure can vary from 109 to 1019 cm−3, corresponding to a degree of ionization from 10−10 to 1 [7,8]. Depending on the method of formation, plasma can be high-temperature (thermal) and low-temperature (nonthermal) plasma (NTP). Thermal plasma is created by heating a gas to a temperature at which electrons are torn from atoms and ions are formed. It can reach a temperature of up to 106 K. NTP is generated by an electric discharge when the generated ions reach a temperature close to the environment (maximum 340 K), which predestines NTP for use in many applications [9,10,11,12]. NTP is often referred to as nonequilibrium plasma because it is not in thermodynamic equilibrium. Nonequilibrium plasma is characterized by the temperature of electrons ranging from a few eV to 10 eV, while the temperature of heavy particles varies from room temperature to a level comparable to the electron temperature but usually lower [13]. NTP is easily formed in the air at atmospheric pressure using various discharges. In addition to air, plasma can also be created in other gases such as nitrogen, oxygen, argon, or carbon dioxide. The most commonly used electric discharges are corona discharge, dielectric barrier discharge, and plasma jet (Figure 1).

2.1.1. Corona Discharge

A corona discharge (Figure 1A) can be observed as a luminous glow. Near to sharp electrodes such as thin wires, spikes, or edges in a highly non-uniform electric field with high intensities, the active region of corona and plasma generation occurs [15]. Point-to-plate geometry, which is a sharply curved electrode arranged as a counterpart to a flat one, is a typical electrode geometry. Corona discharges can be operated in direct current or pulsed mode, where the pointed electrode has a negative or positive potential [9]. Corona discharges are used in various industrial applications [16,17,18,19,20,21].

2.1.2. Dielectric Barrier Discharge

Due to its configuration and flexibility of electrode shapes, dielectric barrier discharge (DBD) (Figure 1b) is one of the most commonly used plasma systems. DBD plasma is generated by a high voltage applied between two metal electrodes, which are covered with a dielectric material (glass, ceramic, or polymer), and micro-discharges do not occur [15,22,23]. These sources operate at frequencies of 50 Hz to 500 kHz, while the voltage amplitude can be up to tens of kV. The gap between the electrodes can be several µm to several cm. Volume and surface DBDs are the most well-known configurations of this arrangement used to treat biological objects. Volumetric DBD is also known as industrial corona [24]. It consists of two parallel plates in a plane, or the electrodes can be curved in the shape of a cylinder. The surface DBD is composed of parallel electrodes separated by a dielectric barrier layer, while the plasma is formed in an uneven electric field. In the surface DBD configuration, the gap between the discharges is flexible, allowing the treatment of objects of different sizes. The disadvantage of this arrangement is the device’s lifetime, which is limited by contact of plasma with the electrodes [25]. The advantages of volume and surface DBD are combined in a coplanar configuration where a dielectric barrier layer covers pairs of linear parallel electrodes with opposite polarity. Electrodes can have an area of up to a few cm2, which makes this type of plasma particularly suitable for processing large surfaces.

2.1.3. Plasma Jet

A plasma jet is not considered a plasma discharge. It is a specific configuration of other discharges, e.g., corona discharge, DBD, and microwave discharge [15]. An auxiliary gas (usually noble gases) flows through the two electrodes generating the plasma, which pushes the plasma out of the electrodes. A stream of active particles burning as a small jet is created. A plasma jet makes a stable, homogeneous, and uniform discharge at atmospheric pressure. It is used in plasma sources called jets, torches, or pens [15]. The disadvantage is that the plasma jet is only suitable for treating small surfaces. When treating large areas, it is necessary to use several jets in a row [26].

2.2. Biologically Active Agents Generated by Plasma

In NTP, depending on the parameters (gas composition, humidity, and temperature), biologically active agents (BAAs) are formed as a result of many physical and chemical processes. Among BAAs generated by plasma, we include, for example, ROS (reactive oxygen species), RNS (reactive nitrogen species), UV radiation, radiation in the visible and infrared spectrum, charged particles, alternating electric field, and heat [4,9,27]. In recent years, many experiments have been conducted dealing with the importance of individual BAAs generated by plasma in the inactivation process of microorganisms [27,28,29,30]. It is difficult to objectively evaluate which plasma component is the most effective because different types of plasma sources do not have to generate BAA in the same amount, and it is always necessary to identify them. Each of these factors inactivates microorganisms independently, but they are much more effective if their synergistic effect is used [31,32,33], making NTP unique. Of all BAAs generated by NTP, ROS and RNS (RONS) are the most critical inactivating agents of plasma, and NTP has been shown to induce oxidative stress, which can result in cell damage or death [27,29,31,32,34,35,36]. RONS are responsible for several biological reactions, from intracellular DNA breaks to protein damage to outer membrane oxidation [28].
Depending on the type of plasma source used and the conditions of plasma generation, the electric field can contribute to the inactivation of microorganisms. Processes similar to electroporation and disruption of cell morphology may occur during NTP biomass treatment. Plasma treatment can break the cell membrane, which then loses integrity, resulting in the leakage of cytoplasmic components out of the cell [27,37,38].
UV radiation has mutagenic to lethal effects and is widely used in sterilizing rooms and spaces. Nevertheless, UV photons originating from the plasma play only a minor role in the inactivation process [27,29,30]. Plasma-generated UV radiation does not have such a striking impact on cells as the use of UV lamps. In addition, many microorganisms contain protective pigments, such as melanin, in the cell wall of fungi, which to some extent, can protect against UV damage [36].
The effect of NTPs and, thus, BAAs originating from plasma on biological material is dose-dependent, although “dose” is still not a precisely defined term [32]. So far, it has been found that the plasma effect on the treated biological material is more substantial with higher plasma power, more prolonged exposure of the biological material, and closer placement of the material to the plasma or the electrode surface. For example, low doses of NTP cause mammalian cells to proliferate, higher doses cause apoptosis, and even higher doses may cause necrosis [32].

2.3. Fungal Molecular Mechanisms in Response to NTP

NTP generated in ambient air can produce reactive species (RONS), such as free electrons, atomic oxygen, singlet oxygen (1O2), .O2, .H, .OH, .NO, .NO2, O3, and atomic oxygen. In aqueous liquids, primary RONS react with each other, and compounds such as H2O2, NO2, and NO3 are formed. Their formation leads to extracellular and intracellular liquid acidification [39,40,41,42]. Most fungal pathogens grow well in an acidic environment but struggle in alkaline conditions [43,44]. It is well known that fungi generally have a wide pH optimum (4–9 pH units). Nevertheless, the drop in intracellular pH could contribute to maintaining membrane potential in the plasma oxidized fungal cell membrane [42].
Reactive species are believed to be a major factor responsible for the effects of plasma on living cells. Although more studies regarding the molecular action of the NTP have been published on bacteria, mammalian, and plant cells than on fungi, many of the mechanisms may be shared by different species [45]. The function of ROS has been well studied in fungal cell signaling. ROS are intracellularly produced as metabolic byproducts under normal physiological conditions during development or stress responses [46]. ROS can react in excess with biomolecules, such as proteins, lipids, and DNA, which can harm cells. Therefore, cells possess several ROS-scavenging systems. The antioxidant systems are composed of nonenzymatic and enzymatic types [47]. The major nonenzymatic antioxidant is tripeptide glutathione, which forms a disulfide bond between cysteines of two glutathione molecules, resulting in the generation of an oxidized form of glutathione. In A. flavus, plasma treatment led to a significant decrease in the reduced form of glutathione, indicating a potent oxidative attack during plasma treatment which likely caused depletion of the reduced glutathione [48]. In addition to glutathione, some other organic compounds in fungi exhibit scavenging properties, such as ascorbic acid, carotenoids, flavonoids, alkaloids, mannitol, and trehalose [49,50]. In addition to non-protein ROS scavengers, thioredoxin proteins, their respective reductases, and antioxidant enzymes such as catalases, superoxide dismutases, and peroxidases are involved in cellular protection against ROS. The role of the antioxidant enzymes in fungal defense in response to plasma treatment was confirmed in A. flavus and S. cerevisiae [48,51].
ROS generated by plasma sources are characterized by a short lifetime and their ability to interact with reduced functional groups of organic compounds in cells [52]. ROS oxidation of cysteine residues in proteins leads to the generation of cysteine sulfenic acid (–SOH) and disulfide bonds between two cysteines. The formation of disulfide bonds is a reversible modification. In yeast S. cerevisiae, transcription factor Yap1 responds to plasma treatment by rapid translocation from the cytoplasm to nucleus. The translocation is initiated by forming disulfide bonds in the protein region governing the transport into the nucleus [53]. Yap1 activates the expression of antioxidant stress response genes.
Sulfenic acid can be oxidized to sulfinic (–SO2H) or sulfonic (–SO3H) acid. This cysteine modification is, however, irreversible [54] and damaging to cells. In addition to cysteine, methionine possesses a sulfur-containing side chain susceptible to oxidation. The oxidized methionine, methionine sulfoxide, is one of the important post-translational modifications [55] that ROS can affect. At the moment, there is very little information on cysteine and methionine oxidations in fungi following plasma treatment.
The major targets of ROS from plasma are fungal cell walls and cytoplasmic membranes. FTIR analysis and electron microscopy in Aspergillus sp. indicated chemical (polysaccharide oxidation) and physical changes (dehydration, ruptures) in cell surface structures [42,48,56,57,58]. Currently, we do not have many studies regarding the nature of ROS interaction with cell surfaces of fungal cells and the depth they can penetrate. ROS are divided into long- and short-lived species. It was reported that the interplay of those species and their concentration gradients and penetrability with the cell surface might initiate a sequence of cell responses [42,59]. Although fungi do not synthesize polyunsaturated fatty acids, malondialdehyde (MDA) formation was determined in fungi after plasma treatment [42,48], indicating lipid peroxidation. Protein and potassium leakage and membrane potential reduction suggested the loss of membrane integrity. Damage to cell membranes inflicted by reactive species also led to mitochondrial malfunction, endoplasmic reticulum stress, defects in protein folding, and intracellular calcium increase [42,53,60,61,62].
In addition to proteins, lipids, and polysaccharides, ROS target nucleic acids. In eukaryotic cells, single-strand and double-strand break formations were reported, along with forming oxidized bases such as 8-oxodeoxyguanosine [63,64]. These breaks are subjected to DNA repair mechanisms which could result in mutations or cell death if the damage overwhelms the DNA repair machinery. Apoptosis-like markers such as chromatin condensation, phosphatidylserine presence on the outer plasma membrane, decrease in mitochondrial transmembrane potential, and cell-cycle arrest [61,65] were determined in yeasts. However, yeast mutants lacking genes for the proapoptotic proteins Yca1p, Aif1p, and Nuc1p (metacaspase, apoptosis-inducing factor, endonuclease G) did not differ significantly in sensitivity from the wildtype when treated with NTP [51]. These results indicate that fungi might have a plasma-specific type of death that does not require the activation of the fungal programmed cell death pathway.

3. NTP Technology in the Management of Fungal Contamination, Disease Control, Protection of Heritage Objects, and Strain Improvement

Microbial inactivation using NTP is especially suitable when traditional decontamination methods are ineffective. Since the differences in the structure and size of cells, their metabolic activity, and the ability to cope with reactive molecules in different microorganisms are not sufficiently studied, a complete generalization of the effects of plasma is not possible. Many studies confirmed the applicability of NTP for the inactivation of fungal cells (Table 1), which show lower sensitivity to NTP than bacteria [66,67,68].
Fungal cells were effectively inactivated by plasma after only a few minutes of exposure to NTP. The action mechanism is based on damage to the structure of cell envelopes and oxidation of macromolecules, similar to bacteria [81]. The level of oxidative stress induced by NTP is a critical factor for cell fate determination. Plasma-generated ROS contribute most to fungal inactivation. The NTP can induce two modes of cell death (apoptosis or necrosis) in fungal cells dependent on treatment time [71]. The most studied fungal genera include Aspergillus sp., Penicillium sp., Fusarium sp., and others. Šimončicová et al. [48] investigated the effect of plasma on A. flavus hyphae, reporting massive structural changes, increased membrane permeability, and DNA degradation. The DNA damage by plasma-induced intracellular RONS was also confirmed in Cordyceps pruinosa spores [77]. Julák et al. [78] observed a delay in the growth of Aspergillus oryzae and Alternaria sp. after exposure of conidia to plasma. This phenomenon is probably related to the mechanism of plasma effects on fungal cells. After nonlethal damage, revitalization processes begin restoring damaged components and functions. Yeasts, especially Candida sp. and Saccharomyces sp., are frequently used as model organisms. Tyczkowska-Sieroń et al. [73] studied changes in the genome of Candida albicans after exposure to a sublethal dose of plasma. They identified six single-nucleotide variants, six insertions, and five deletions and also demonstrated that, of the 19 hydrolytic enzymes, nine were inactive, nine temporarily decreased the activity, and one constitutively increased the activity after plasma exposure. Carbon assimilation and drug sensitivity were not affected by plasma. Hence, they concluded that the changes in surviving C. albicans cells did not impose significant danger to the environment, especially regarding drug resistance and pathogenicity.
Some microorganisms can form mono- or polymicrobial aggregates referred to as biofilms. This structure protects pathogenic microorganisms from antimicrobial agents and the immune system. According to some estimates, a pathogen biofilm is present in the body in up to 80% of diseases. C. albicans is one of the most common human opportunistic yeasts. Infections caused by C. albicans are associated with their ability to form a biofilm. Several studies proved the positive effect of plasma on biofilm inactivation [82,83,84]. The complete killing of C. albicans cells in the biofilm was observed after 8 min of plasma treatment [84]. A study of A. flavus biofilm showed that plasma treatment has detrimental effects on the biofilm structure. At the same time, it pointed out that the fungicidal effect of plasma may depend on the initial concentration of the inoculum [82].

3.1. Plasma Medicine

NTP generated at atmospheric pressure shows promising biomedical applications leading to the emergence of plasma medicine that includes the inactivation of bacteria, fungi, viruses, and endospores, blood clotting, wound healing, and tooth whitening. Applications in antitumor therapy are also being studied, where plasma exhibits an antitumor effect on a wide range of cancer cell lines [33,85,86,87,88].
Fungal infections cause a complex set of disease states that cause tissue destruction or may result from inflammation caused by the presence of the fungus [89]. Among the relatively common fungal diseases are candidiasis, onychomycosis, and dermatophytosis. Older people, people with organ transplants, HIV-positive people, and diabetics are especially prone to developing candidal infections [90,91]. Borges et al. [92] tested plasma jet as a possible effective tool for preventing oral candidiasis in vivo. After only 5 min, they observed a significant decrease in the viability of the C. albicans biofilm. Histological analyses revealed a significantly lower incidence of inflammatory changes and a substantial reduction in candidal tissue invasion in the plasma-treated group. Park et al. [93] found that 1–5 min application of no-ozone cold plasma inhibited the growth of C. albicans by approximately 2 log.
Dermatophytosis is a term used to describe fungal infections caused by fungi that colonize the surface of the skin, hair, or nails. The most common are representatives of the genera Epidermophyton, Microsporum, and Trichophyton [94]. The effect of plasma in preventing dermatophytosis was monitored with silver nanoparticles. Such treatment decreased the minimum inhibitory concentration of nanoparticles, increased mycelial permeability to nanoparticles, and increased the effectiveness of healing and suppression of disease symptoms on the skin [95]. In guinea pigs infected with Trichophyton mentagrophytes, plasma treatment shortened and attenuated the infection and significantly reduced the viability of the pathogen without adverse effects on the animal model [96].
Onychomycosis is a nail fungal infection that afflicts almost 6% of the population worldwide [97,98]. About 70% of these infections are caused by dermatophytes (Trichophyton rubrum (more than 50%), Trichophyton mentagrophytes, Epidermophytonon floccosum, Microsporum spp., Trichophyton violaceum, Trichophyton verrucosum, Trichophyton krajdenii, and Arthroderma spp.), about 20% are caused by non-dermatophyte molds (Scopulariopsis brevicaulis, Aspergillus spp., Acremonium, Fusarium spp., Alternaria alternata, and Neoscytalidium), and 10–20% are caused by yeasts (Candida spp.) [99]. Bulson et al. [100] observed the complete killing of C. albicans and T. mentagrophytes by plasma in suspension and on nails. A similar plasma effect was also observed in the inactivation of T. rubrum, Trichophyton interdigital, and Trichophyton benhamiae [101].

3.2. Plasma Food Technology and Agriculture

To fulfill the needs of an ever-growing population, it is necessary to ensure a sufficient amount of high-quality raw materials. In this case, NTP is a suitable alternative to the already used technologies [102,103,104]. Plasma has been effectively used to decontaminate various food surfaces such as fruit, vegetables, and meat (Table 2). Park et al. [105] investigated the effect of plasma on the reduction of Cladosporium cladosporioides and Penicillium citrinum on the surface of dried filefish fillets. After 20 min of treatment, they determined a 0.9–1 log reduction of CFU/g, but at the same time observed a decrease in overall sensory acceptance. Plasma treatment caused a reduction of viable fungal spores on beef jerky but harmed off-color, flavor, and overall acceptability [106]. Royintarat et al. [107] used the synergistic effect of ultrasound and plasma-activated water (indirect plasma action) to reduce microbial contamination of chicken meat. Sudheesh and Sunooj [108] used plasma to treat fresh-cut fruits and vegetables. In addition to the inactivation of the microbial cell, they observed a decrease in enzymatic activity (pectin methylesterase and polyphenol oxidase), which is also related to the browning speed. There was also a decrease in antioxidant content and antioxidant activity. Thanks to the possibility of plasma generation in liquids [109], this method is also appropriate for decontaminating water, milk, and fruit juices [110,111]. Xiang et al. [112] used plasma to inactivate the yeast Zygosaccharomyces rouxii in apple juice. Treatment of juice with plasma for 140 s reduced Z. rouxii by approximately 5 log. At the same time, plasma caused significant changes in apple juice’s pH, acidity, and color parameters, but had no effect on the content of total soluble solids, reducing sugars, and total phenols. The changes in apple juice acidity may be related to the production of acidogenic molecules such as NOx or H+ dissociated from H2O and other components in apple juice during DBD plasma treatment. No significant changes in physicochemical properties were observed in tomato juice [113]. However, several studies have demonstrated the effects of plasma on components determining food quality, such as pH, proteins and enzymes, sugars, lipids, vitamins, and others [110,114].
The safety of crops and food is critical because of the health risk and the enormous economic losses. NTP can also be used for disinfection post-harvest fruits and vegetables. DSBD plasma was effectively used to inhibit the growth of natural microbiota and the natural decay of blueberries. After less than 15 min, only modest effects of plasma on blueberry quality were observed. However, 20 min treatment resulted in severe oxidative damage to the peels [115]. Plasma treatment did not significantly change the taste, aroma, color, and texture of kumquat [116] or the color and hardness of paprika during storage [117]. The treatment of mung bean sprouts with PAW did not cause significant changes in mung bean’s total phenolic and flavonoid content and sensory properties [118]. Using a microwave plasma jet significantly increased mandarin peel’s entire phenolic content and antioxidant activity [119]. Liu et al. [120] developed a high-field plasma system at atmospheric pressure to control and keep the storage area clean and to keep plants such as vegetables, fruits, and flowers fresh for longer. The study showed that, with the help of the plasma system, fresh fruits (bananas, grapefruits) are preserved much longer compared to the conventional methods. The amount of ethane emitted during storage was also reduced. Ambrico et al. [121] found that pretreatment of cherries with plasma leads to increased resistance to subsequent fungal infection. It is also worth mentioning a study showing that NTP can degrade pesticide residues in fruits and vegetables [122].
Contamination of food with mycotoxins is a global problem. Despite implementing various measures in agriculture, the contamination of raw materials during storage and processing cannot be completely prevented. Another problem is that, due to the high stability of mycotoxins against thermal, physical, and chemical influences, it is impossible to remove them altogether during food processing [148]. Mycotoxins spoil food and feed, threaten human and animal health, and hinder international trade [149]. Approximately 25% of the world’s crops are contaminated with mycotoxins each year, resulting in enormous agricultural and industrial losses estimated in the billions of dollars. The main mycotoxin-producing fungal genera include Aspergillus, Fusarium, and Penicillium. While species of the genera Aspergillus and Penicillium contaminate food and feed during storage, species of the genus Fusarium colonize crops directly in fields and plantations [148]. NTP was effectively applied for inactivating mycotoxin producers. Therefore, questions arose about whether plasma could be used for mycotoxin degradation. Aflatoxin B1 was completely degraded after plasma treatment of corn kernels [134], approximately 73% degradation was observed on hazelnuts [137], and a 45–56% reduction was achieved on rice and wheat [136]. Hojnik et al. [150] investigated the possible cytotoxic and genotoxic potential of aflatoxin B1 (AFB1) plasma degradation products on human hepatocellular carcinoma cells. Cytotoxic and genotoxic effects of NTP-treated AFB1 compared to NTP-untreated AFB1 were not confirmed. Hoppanová et al. [57,58] investigated changes in aflatoxin and ochratoxin production in response to plasma-induced oxidative stress. Their results clearly showed that NTP can significantly reduce viable cells. However, the cells that survived the plasma treatment were able to produce mycotoxins at an increased rate in the early stages of growth and their production slowed down in the later stages of growth. From a practical point of view, this means that, even after decontamination of food with plasma, it is still necessary to follow the principles of proper and safe food storage.
Many studies point to the positive results of using plasma in agriculture (Table 2). In addition to seed disinfection, plasma can improve the germination rate of many seeds, which can lead to enhanced production [67,142,147,151,152,153,154]. It has been shown that irrigation using plasma-activated water leads to better growth of radishes, tomatoes, and peppers [155]. Changes in the seed’s surface properties were also observed, thanks to which their wettability and water absorption increased [66,67,156,157]. It was observed that just 10 s of plasma treatment changes the surface of cereal seeds from hydrophobic to hydrophilic. Due to the better wettability of the seeds, it is necessary to apply a lower volume of chemical fungicides. By combining physical (NTP) and chemical (fungicide) treatment of cereal seeds, it is possible to effectively reduce the required amount of chemical fungicide and stimulate the germination and early growth parameters of the seed [158]. NTP could be an alternative for reducing the amount of chemical fungicides used in agriculture and for the degradation of toxic chemical compounds such as phenols and azo-dyes [159].

3.3. Plasma and Cultural Heritage Objects

Due to their high enzymatic activity and ability to grow even at low aw values, fungi can grow on paper, parchment, paintings, textiles, and other materials. Thus, they play a crucial role in damaging cultural heritage. Among the most widespread fungal genera damaging historical objects are Alternaria sp., Aspergillus sp., Aureobasidium pullulans, Fusarium sp., Mucor sp., Penicillium sp., Botrytis cinerea, Trichoderma harzianum a Trichoderma viride, Cladosporium cladosporioides, and Epicoccum nigrum [160]. NTP is a possible and effective method of inactivating fungal contamination to effectively save historical artifacts. DBD plasma is used to stabilize documents containing iron gall inks [161]. Low-temperature ADRE (atmospheric discharge with runaway electron) plasma can decontaminate the surfaces of various lignocellulosic materials from five types of filamentous fungi (A. alternata, Cladosporium herbarum, Penicillium chrysogenum, A. niger, and Trichoderma atroviride). The least sensitive to ADRE plasma treatment were the filamentous fungi P. chrysogenum and A. niger, which were most represented in archives and libraries [162]. These studies indicate that NTP is a promising alternative to other convective methods of inactivating fungal contamination of historical objects.

3.4. Plasma in Biotechnology

In the previous sections, we presented many studies focused mainly on the inhibition and inactivation of fungi in various industries such as medicine, agriculture, and food control. However, not all fungal genera are undesirable for humans. Many fungal species produce interesting substances (antibiotics, pigments, and enzymes). In recent years, studies have been emerging investigating the positive effect of NTP on beneficial fungi. Improving the beneficial aspects of fungi using plasma occurs in two ways, through mutagenic or non-mutagenic changes. Studies using plasma for mutagenesis of fungal cells are summarized in Table 3.
Most studies [175,176,177,178,179] used the ARTP plasma mutation system, formed by a radiofrequency atmospheric-pressure glow discharge plasma jet, to mutagenize fungal cells [180]. The Saccharomyces cerevisiae mutant prepared by the ARTP mutation system produced approximately 57% more glutathione, and an improvement in glutathione synthetase activity was also observed [181]. After chemical–physical mutagenesis, Rhodotorula mucilaginosa K4, with a 67% greater concentration of carotenoids than Rhodotorula mucilaginosa KC8, was obtained [178]. The mutated strain JNDY-13, which was obtained with T. reesei RUT-C30 as the parental strain, had an increased production of cellulases, which may be related to a mutation in the galactokinase gene. Upregulation of cellulase and hemicellulase genes was also noted in this mutant [174]. In the C. tropicalis mutant, in addition to an increase in xylitol production, an increase in xylose reductase gene expression and activity was observed [170]. Feng et al. [171] applied to A. oryzae KA-11 a combined mutagenesis program that included microwave mutagenesis, UV irradiation, heat-LiCl, and ARTP. Kojic acid production was increased by 47.0%, 87.1%, 126.2%, and 292.3% compared to the starting strain KA-11 after each stage of mutagenesis. From the obtained results, it is clear that the best results were obtained with ARTP mutagenesis.
Several studies focused on improving spore germination and protein secretion in a non-mutagenic way. A study by Farasat et al. [182] evaluated the effect of NTP on the production of recombinant phytase in the yeast Pichia pastoris, as well as the structure and function of the phytase enzyme. The yeast produced higher amounts of recombinant phytase after direct or indirect exposure to plasma. Plasma treatment of a commercial phytase solution with NTP caused up to a 125% increase in enzyme activity. It was also shown that this protein maintained its secondary structure after plasma treatment, while the tertiary structure was slightly changed. Veerana et al. treated A. oryzae cells with two plasma discharges, specifically a micro dielectric barrier discharge (MDBD) in nitrogen [183] and a plasma jet in the air [184]. Using MDBD plasma, they achieved a significant increase in the percentage of spore germination after 2 and 5 min of treatment. They also observed a 7.4–9.3% increase in α-amylase activity 24 and 48 h after plasma treatment [183]. After treatment with a plasma jet, they noted an approximately 10% increase in spore germination after 5 and 10 min of treatment and a significant increase in α-amylase activity 24–96 h after plasma treatment [184].

4. Summary and Prospects

Most studies explored NTP’s application in fungal decontamination, plasma medicine, seed protection, fungal breeding, food processing, preservation, and cultural heritage protection. Despite many advantages that could be exploited, we seem to have reached a point where we must carefully evaluate the positives and negatives when applying this technology to treat fungi. The research concerning fungi and plasma is even more complex because it involves various plasma source configurations, dose determination, working gas compositions, biological and nonbiological matrixes, or liquids (Figure 2).
When working with filamentous fungi, we face many challenges that stem from fungal diversity, the ability to form complex structures, and the formation of hundreds of types of cells that respond to plasma treatment differently. On the one hand, NTP could help combat the emergence of novel pathogens and antifungal-resistant strains by reducing antifungal agents. Nevertheless, on the other hand, the generation and potential spread of genetically modified strains should be of concern when the large-scale employment of NTP is planned. In the future, we have to address not only technical challenges. We must also fill those gaps in understanding the molecular mechanisms involved in fungal interactions with reactive species present in plasma.

Author Contributions

L.H. and S.K. wrote and reviewed the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the Slovak Research and Development Agency APVV-20-0257, APVV-16-0216, and Scientific grant agency VEGA 1/0663/22.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Veerana, M.; Yu, N.; Ketya, W.; Park, G. Application of non-thermal plasma to fungal resources. J. Fungi 2022, 8, 102. [Google Scholar] [CrossRef] [PubMed]
  2. Mravlje, J.; Regvar, M.; Vogel-Mikuš, K. Development of cold plasma technologies for surface decontamination of seed fungal pathogens: Present status and perspectives. J. Fungi 2021, 7, 650. [Google Scholar] [CrossRef] [PubMed]
  3. Polčic, P.; Machala, Z. Effects of non-thermal plasma on yeast Saccharomyces cerevisiae. Int. J. Mol. Sci. 2021, 22, 2247. [Google Scholar] [CrossRef]
  4. Misra, N.N.; Tiwari, B.K.; Raghavarao, K.S.M.S.; Cullen, P.J. Nonthermal plasma inactivation of food-borne pathogens. Food Eng. Rev. 2011, 3, 159–170. [Google Scholar] [CrossRef]
  5. Bellan, P.M. Fundamentals of Plasma Physics; Cambridge University Press: Cambridge, UK, 2008; ISBN 9781139449731. [Google Scholar]
  6. Fridman, A. Plasma Chemistry; Cambridge University Press: Cambridge, UK, 2008; ISBN 9781139471732. [Google Scholar]
  7. Gershman, S. Pulsed Electrical Discharge in Gas Bubbles in Water. Ph.D. Thesis, Rutgers the State University of New Jersey-New Brunswick, New Brunswick, NJ, USA, 2008. [Google Scholar]
  8. Reedijk, J. Reference Module in Chemistry, Molecular Sciences and Chemical Engineering; Elsevier: Oxford, UK, 2014. [Google Scholar]
  9. Ehlbeck, J.; Schnabel, U.; Polak, M.; Winter, J.; von Woedtke, T.; Brandenburg, R.; von dem Hagen, T.; Weltmann, K.D. Low temperature atmospheric pressure plasma sources for microbial decontamination. J. Phys. D Appl. Phys. 2010, 44, 13002. [Google Scholar] [CrossRef]
  10. Šimončicová, J.; Kryštofová, S.; Medvecká, V.; Ďurišová, K.; Kaliňáková, B. Technical applications of plasma treatments: Current state and perspectives. Appl. Microbiol. Biotechnol. 2019, 103, 5117–5129. [Google Scholar] [CrossRef]
  11. Gupta, T.T.; Ayan, H. Application of non-thermal plasma on biofilm: A review. Appl. Sci. 2019, 9, 3548. [Google Scholar] [CrossRef]
  12. Šrámková, P.; Zahoranová, A.; Kelar, J.; Kelar Tučeková, Z.; Stupavská, M.; Krumpolec, R.; Jurmanová, J.; Kováčik, D.; Černák, M. Cold atmospheric pressure plasma: Simple and efficient strategy for preparation of poly(2-oxazoline)-based coatings designed for biomedical applications. Sci. Rep. 2020, 10, 9478. [Google Scholar] [CrossRef]
  13. Guo, H.; Zhang, X.N.; Chen, J.; Li, H.-P.; Ostrikov, K. Non-equilibrium synergistic effects in atmospheric pressure plasmas. Sci. Rep. 2018, 8, 4783. [Google Scholar] [CrossRef]
  14. Laroque, D.A.; Seó, S.T.; Valencia, G.A.; Laurindo, J.B.; Carciofi, B.A.M. Cold plasma in food processing: Design, mechanisms, and application. J. Food Eng. 2022, 312, 110748. [Google Scholar] [CrossRef]
  15. Scholtz, V.; Pazlarova, J.; Souskova, H.; Khun, J.; Julak, J. Nonthermal plasm0061—A tool for decontamination and disinfection. Biotechnol. Adv. 2015, 33 Pt 2, 1108–1119. [Google Scholar] [CrossRef]
  16. Ruangwong, K.; Rongsangchaicharean, T.; Thammaniphit, C.; Onwimol, D.; Srisonphan, S. Atmospheric corona discharge plasma for rice (Oryza sativa L.) seed surface modification, fungi decontamination, and shelf life extension. Plasma Med. 2020, 10, 191–201. [Google Scholar] [CrossRef]
  17. Porsev, E.G.; Druzhinina, N.S. Electric corona discharge as a basis for managing the quality of solid cereals. IOP Conf. Ser. Mater. Sci. Eng. 2019, 560, 12172. [Google Scholar] [CrossRef]
  18. Jose, J.; Ramanujam, S.; Philip, L. Applicability of pulsed corona discharge treatment for the degradation of chloroform. Chem. Eng. J. 2019, 360, 1341–1354. [Google Scholar] [CrossRef]
  19. Hassani, O.F.; Merbahi, N.; Oushabi, A.; Elfadili, M.H.; Kammouni, A.; Oueldna, N. Effects of corona discharge treatment on surface and mechanical properties of Aloe Vera fibers. Mater. Today Proc. 2020, 24, 46–51. [Google Scholar] [CrossRef]
  20. Singh, R.K.; Philip, L.; Ramanujam, S. Continuous flow pulse corona discharge reactor for the tertiary treatment of drinking water: Insights on disinfection and emerging contaminants removal. Chem. Eng. J. 2019, 355, 269–278. [Google Scholar] [CrossRef]
  21. KS Narayanan, S.S.; Wang, X.; Paul, J.; Paley, V.; Weng, Z.; Ye, L.; Zhong, Y. Disinfection and electrostatic recovery of N95 respirators by corona discharge for safe reuse. Environ. Sci. Technol. 2021, 55, 15351–15360. [Google Scholar] [CrossRef]
  22. Misra, N.N.; Yepez, X.; Xu, L.; Keener, K. In-package cold plasma technologies. J. Food Eng. 2019, 244, 21–31. [Google Scholar] [CrossRef]
  23. Wagner, H.E.; Brandenburg, R.; Kozlov, K.V.; Sonnenfeld, A.; Michel, P.; Behnke, J.F. The barrier discharge: Basic properties and applications to surface treatment. Vacuum 2003, 71, 417–436. [Google Scholar] [CrossRef]
  24. Pykönen, M.; Silvaani, H.; Preston, J.; Fardim, P.; Toivakka, M. Plasma activation induced changes in surface chemistry of pigment coating components. Colloids Surf. A Physicochem. Eng. Asp. 2009, 352, 103–112. [Google Scholar] [CrossRef]
  25. Kogelschatz, U. Dielectric-barrier discharges: Their history, discharge physics, and industrial applications. Plasma Chem. Plasma Process. 2003, 23, 1–46. [Google Scholar] [CrossRef]
  26. Bermudez-Aguirre, D. Advances in Cold Plasma Applications for Food Safety and Preservation; Elsevier Science: Amsterdam, The Netherlands, 2019; ISBN 9780128149225. [Google Scholar]
  27. Deng, X.; Shi, J.; Kong, M.G. Physical mechanisms of inactivation of Bacillus subtilis spores using cold atmospheric plasmas. IEEE Trans. Plasma Sci. 2006, 34, 1310–1316. [Google Scholar] [CrossRef]
  28. Moisan, M.; Barbeau, J.; Moreau, S.; Pelletier, J.; Tabrizian, M.; Yahia, L.H. Low-temperature sterilization using gas plasmas: A review of the experiments and an analysis of the inactivation mechanisms. Int. J. Pharm. 2001, 226, 1–21. [Google Scholar] [CrossRef]
  29. Laroussi, M.; Leipold, F. Evaluation of the roles of reactive species, heat, and UV radiation in the inactivation of bacterial cells by air plasmas at atmospheric pressure. Int. J. Mass Spectrom. 2004, 233, 81–86. [Google Scholar] [CrossRef]
  30. Dobrynin, D.; Fridman, G.; Friedman, G.; Fridman, A. Physical and biological mechanisms of direct plasma interaction with living tissue. New J. Phys. 2009, 11, 115020. [Google Scholar] [CrossRef]
  31. Fridman, G.; Friedman, G.; Gutsol, A.; Shekhter, A.B.; Vasilets, V.N.; Fridman, A. Applied plasma medicine. Plasma Process. Polym. 2008, 5, 503–533. [Google Scholar] [CrossRef]
  32. Graves, D.B. Low temperature plasma biomedicine: A tutorial review. Phys. Plasmas 2014, 21, 80901. [Google Scholar] [CrossRef]
  33. Fridman, G.; Shereshevsky, A.; Jost, M.M.; Brooks, A.D.; Fridman, A.; Gutsol, A.; Vasilets, V.; Friedman, G. Floating electrode dielectric barrier discharge plasma in air promoting apoptotic behavior in melanoma skin cancer cell lines. Plasma Chem. Plasma Process. 2007, 27, 163–176. [Google Scholar] [CrossRef]
  34. Jiang, F.; Zhang, Y.; Dusting, G.J. NADPH oxidase-mediated redox signaling: Roles in cellular stress response, stress tolerance, and tissue repair. Pharmacol. Rev. 2011, 63, 218. [Google Scholar] [CrossRef]
  35. Arjunan, K.P.; Sharma, V.K.; Ptasinska, S. Effects of atmospheric pressure plasmas on isolated and cellular DNA—A review. Int. J. Mol. Sci. 2015, 16, 2971–3016. [Google Scholar] [CrossRef] [Green Version]
  36. Misra, N.N.; Yadav, B.; Roopesh, M.S.; Jo, C. Cold plasma for effective fungal and mycotoxin control in foods: Mechanisms, inactivation effects, and applications. Compr. Rev. Food Sci. Food Saf. 2019, 18, 106–120. [Google Scholar] [CrossRef] [PubMed]
  37. Hong, Y.F.; Kang, J.G.; Lee, H.Y.; Uhm, H.S.; Moon, E.; Park, Y.H. Sterilization effect of atmospheric plasma on Escherichia coli and Bacillus subtilis endospores. Lett. Appl. Microbiol. 2009, 48, 33–37. [Google Scholar] [CrossRef] [PubMed]
  38. Laroussi, M.; Mendis, D.A.; Rosenberg, M. Plasma interaction with microbes. New J. Phys. 2003, 5, 41. [Google Scholar] [CrossRef]
  39. Bauer, G. Intercellular singlet oxygen-mediated bystander signaling triggered by long-lived species of cold atmospheric plasma and plasma-activated medium. Redox Biol. 2019, 26, 101301. [Google Scholar] [CrossRef] [PubMed]
  40. Machala, Z.; Tarabová, B.; Sersenová, D.; Janda, M.; Hensel, K. Chemical and antibacterial effects of plasma activated water: Correlation with gaseous and aqueous reactive oxygen and nitrogen species, plasma sources and air flow conditions. J. Phys. D Appl. Phys. 2018, 52, 34002. [Google Scholar] [CrossRef]
  41. Kondeti, V.S.S.K.; Phan, C.Q.; Wende, K.; Jablonowski, H.; Gangal, U.; Granick, J.L.; Hunter, R.C.; Bruggeman, P.J. Long-lived and short-lived reactive species produced by a cold atmospheric pressure plasma jet for the inactivation of Pseudomonas aeruginosa and Staphylococcus aureus. Free. Radic. Biol. Med. 2018, 124, 275–287. [Google Scholar] [CrossRef]
  42. Xu, H.; Zhu, Y.; Du, M.; Wang, Y.; Ju, S.; Ma, R.; Jiao, Z. Subcellular mechanism of microbial inactivation during water disinfection by cold atmospheric-pressure plasma. Water Res. 2021, 188, 116513. [Google Scholar] [CrossRef]
  43. Wheeler, K.A.; Hurdman, B.F.; Pitt, J.I. Influence of pH on the growth of some toxigenic species of Aspergillus, Penicillium and Fusarium. Int. J. Food Microbiol. 1991, 12, 141–149. [Google Scholar] [CrossRef]
  44. Rousk, J.; Bååth, E.; Brookes, P.C.; Lauber, C.L.; Lozupone, C.; Caporaso, J.G.; Knight, R.; Fierer, N. Soil bacterial and fungal communities across a pH gradient in an arable soil. ISME J. 2010, 4, 1340–1351. [Google Scholar] [CrossRef]
  45. Liao, X.; Liu, D.; Xiang, Q.; Ahn, J.; Chen, S.; Ye, X.; Ding, T. Inactivation mechanisms of non-thermal plasma on microbes: A review. Food Control 2017, 75, 83–91. [Google Scholar] [CrossRef]
  46. Zhang, Z.; Chen, Y.; Li, B.; Chen, T.; Tian, S. Reactive oxygen species: A generalist in regulating development and pathogenicity of phytopathogenic fungi. Comput. Struct. Biotechnol. J. 2020, 18, 3344–3349. [Google Scholar] [CrossRef]
  47. Heller, J.; Tudzynski, P. Reactive oxygen species in phytopathogenic fungi: Signaling, development, and disease. Annu. Rev. Phytopathol. 2011, 49, 369–390. [Google Scholar] [CrossRef] [PubMed]
  48. Šimončicová, J.; Kaliňáková, B.; Kováčik, D.; Medvecká, V.; Lakatoš, B.; Kryštofová, S.; Hoppanová, L.; Palušková, V.; Hudecová, D.; Ďurina, P.; et al. Cold plasma treatment triggers antioxidative defense system and induces changes in hyphal surface and subcellular structures of Aspergillus flavus. Appl. Microbiol. Biotechnol. 2018, 102, 6647–6658. [Google Scholar] [CrossRef] [PubMed]
  49. Blokhina, O.; Virolainen, E.; Fagerstedt, K.V. Antioxidants, oxidative damage and oxygen deprivation stress: A review. Ann. Bot. 2003, 91, 179–194. [Google Scholar] [CrossRef] [PubMed]
  50. Hasanuzzaman, M.; Bhuyan, M.; Zulfiqar, F.; Raza, A.; Mohsin, S.M.; Mahmud, J.A.; Fujita, M.; Fotopoulos, V. Reactive oxygen species and antioxidant defense in plants under abiotic stress: Revisiting the crucial role of a universal defense regulator. Antioxidants 2020, 9, 681. [Google Scholar] [CrossRef]
  51. Polčic, P.; Pakosová, L.; Chovančíková, P.; Machala, Z. Reactive cold plasma particles generate oxidative stress in yeast but do not trigger apoptosis. Can. J. Microbiol. 2018, 64, 367–375. [Google Scholar] [CrossRef]
  52. Hara, H.; Adachi, T. Molecular mechanisms of non-thermal atmospheric pressure plasma-induced cellular responses. Jpn. J. Appl. Phys. 2021, 60, 20501. [Google Scholar] [CrossRef]
  53. Itooka, K.; Takahashi, K.; Izawa, S. Fluorescence microscopic analysis of antifungal effects of cold atmospheric pressure plasma in Saccharomyces cerevisiae. Appl. Microbiol. Biotechnol. 2016, 100, 9295–9304. [Google Scholar] [CrossRef]
  54. Spickett, C.M.; Pitt, A.R.; Morrice, N.; Kolch, W. Proteomic analysis of phosphorylation, oxidation and nitrosylation in signal transduction. Biochim. Biophys. Acta (BBA)—Proteins Proteom. 2006, 1764, 1823–1841. [Google Scholar] [CrossRef]
  55. Hoshi, T.; Heinemann, S.H. Regulation of cell function by methionine oxidation and reduction. J. Physiol. 2001, 531, 1–11. [Google Scholar] [CrossRef]
  56. Hojnik, N.; Modic, M.; Ni, Y.; Filipič, G.; Cvelbar, U.; Walsh, J.L. Effective fungal spore inactivation with an environmentally friendly approach based on atmospheric pressure air plasma. Environ. Sci. Technol. 2019, 53, 1893–1904. [Google Scholar] [CrossRef] [PubMed]
  57. Hoppanová, L.; Dylíková, J.; Kováčik, D.; Medvecká, V.; Ďurina, P.; Kryštofová, S.; Zahoranová, A.; Kaliňáková, B. The effect of cold atmospheric pressure plasma on Aspergillus ochraceus and ochratoxin A production. Antonie Leeuwenhoek 2020, 113, 1479–1488. [Google Scholar] [CrossRef] [PubMed]
  58. Hoppanová, L.; Dylíková, J.; Kováčik, D.; Medvecká, V.; Ďurina, P.; Kryštofová, S.; Hudecová, D.; Kaliňáková, B. Non-thermal plasma induces changes in aflatoxin production, devitalization, and surface chemistry of Aspergillus parasiticus. Appl. Microbiol. Biotechnol. 2022, 106, 2107–2119. [Google Scholar] [CrossRef]
  59. Bogaerts, A.; Yusupov, M.; Razzokov, J.; Van der Paal, J. Plasma for cancer treatment: How can RONS penetrate through the cell membrane? Answers from computer modeling. Front. Chem. Sci. Eng. 2019, 13, 253–263. [Google Scholar] [CrossRef]
  60. Itooka, K.; Takahashi, K.; Kimata, Y.; Izawa, S. Cold atmospheric pressure plasma causes protein denaturation and endoplasmic reticulum stress in Saccharomyces cerevisiae. Appl. Microbiol. Biotechnol. 2018, 102, 2279–2288. [Google Scholar] [CrossRef]
  61. Ma, R.N.; Feng, H.Q.; Liang, Y.D.; Zhang, Q.; Tian, Y.; Su, B.; Zhang, J.; Fang, J. An atmospheric-pressure cold plasma leads to apoptosis in Saccharomyces cerevisiae by accumulating intracellular reactive oxygen species and calcium. J. Phys. D Appl. Phys. 2013, 46, 285401. [Google Scholar] [CrossRef]
  62. Veerana, M.; Mitra, S.; Ki, S.-H.; Kim, S.-M.; Choi, E.-H.; Lee, T.; Park, G. Plasma-mediated enhancement of enzyme secretion in Aspergillus oryzae. Microb. Biotechnol. 2021, 14, 262–276. [Google Scholar] [CrossRef]
  63. Kurita, H.; Haruta, N.; Uchihashi, Y.; Seto, T.; Takashima, K. Strand breaks and chemical modification of intracellular DNA induced by cold atmospheric pressure plasma irradiation. PLoS ONE 2020, 15, e0232724. [Google Scholar] [CrossRef]
  64. Gaur, N.; Kurita, H.; Oh, J.-S.; Miyachika, S.; Ito, M.; Mizuno, A.; Cowin, A.J.; Allinson, S.; Short, R.D.; Szili, E.J. On cold atmospheric-pressure plasma jet induced DNA damage in cells. J. Phys. D Appl. Phys. 2020, 54, 35203. [Google Scholar] [CrossRef]
  65. Čtvrtečková, L.; Pichová, A.; Scholtz, V.; Khun, J.; Julák, J. Non-thermal plasma-induced apoptosis in yeast Saccharomyces cerevisiae. Contrib. Plasma Phys. 2019, 59, e201800064. [Google Scholar] [CrossRef]
  66. Zahoranová, A.; Hoppanová, L.; Šimončicová, J.; Tučeková, Z.; Medvecká, V.; Hudecová, D.; Kaliňáková, B.; Kováčik, D.; Černák, M. Effect of cold atmospheric pressure plasma on maize seeds: Enhancement of seedlings growth and surface microorganisms inactivation. Plasma Chem. Plasma Process. 2018, 38, 969–988. [Google Scholar] [CrossRef]
  67. Los, A.; Ziuzina, D.; Akkermans, S.; Boehm, D.; Cullen, P.J.; Van Impe, J.; Bourke, P. Improving microbiological safety and quality characteristics of wheat and barley by high voltage atmospheric cold plasma closed processing. Food Res. Int. 2018, 106, 509–521. [Google Scholar] [CrossRef] [PubMed]
  68. Soušková, H.; Scholtz, V.; Julák, J.; Kommová, L.; Savická, D.; Pazlarová, J. The survival of micromycetes and yeasts under the low-temperature plasma generated in electrical discharge. Folia Microbiol. 2011, 56, 77–79. [Google Scholar] [CrossRef] [PubMed]
  69. Intanon, W.; Vichiansan, N.; Leksakul, K.; Boonyawan, D.; Kumla, J.; Suwannarach, N.; Lumyong, S. Inhibition of the aflatoxin-producing fungus Aspergillus flavus by a plasma jet system. J. Food Process. Preserv. 2021, 45, e15045. [Google Scholar] [CrossRef]
  70. Siadati, S.; Pet’ková, M.; Kenari, A.J.; Kyzek, S.; Gálová, E.; Zahoranová, A. Effect of a non-thermal atmospheric pressure plasma jet on four different yeasts. J. Phys. D Appl. Phys. 2020, 54, 25204. [Google Scholar] [CrossRef]
  71. Xu, H.; Ma, R.; Zhu, Y.; Du, M.; Zhang, H.; Jiao, Z. A systematic study of the antimicrobial mechanisms of cold atmospheric-pressure plasma for water disinfection. Sci. Total Environ. 2020, 703, 134965. [Google Scholar] [CrossRef]
  72. Ki, S.H.; Noh, H.; Ahn, G.R.; Kim, S.H.; Kaushik, N.K.; Choi, E.H.; Lee, G.J. Influence of nonthermal atmospheric plasma-activated water on the structural, optical, and biological properties of Aspergillus brasiliensis spores. Appl. Sci. 2020, 10, 6378. [Google Scholar] [CrossRef]
  73. Tyczkowska-Sieroń, E.; Kałużewski, T.; Grabiec, M.; Kałużewski, B.; Tyczkowski, J. Genotypic and phenotypic changes in Candida albicans as a result of cold plasma treatment. Int. J. Mol. Sci. 2020, 21, 8100. [Google Scholar] [CrossRef]
  74. Xu, H.; Zhu, Y.; Cui, D.; Du, M.; Wang, J.; Ma, R.; Jiao, Z. Evaluating the roles of OH radicals, H2O2, ORP and pH in the inactivation of yeast cells on a tissue model by surface micro-discharge plasma. J. Phys. D Appl. Phys. 2019, 52, 395201. [Google Scholar] [CrossRef]
  75. Fukuda, S.; Kawasaki, Y.; Izawa, S. Ferrous chloride and ferrous sulfate improve the fungicidal efficacy of cold atmospheric argon plasma on melanized Aureobasidium pullulans. J. Biosci. Bioeng. 2019, 128, 28–32. [Google Scholar] [CrossRef]
  76. Wu, M.; Liu, C.; Chiang, C.; Lin, Y.; Lin, Y.; Chang, Y.; Wu, J. Inactivation effect of Colletotrichum gloeosporioides by Long-lived chemical species using atmospheric-pressure corona plasma-activated water. IEEE Trans. Plasma Sci. 2019, 47, 1100–1104. [Google Scholar] [CrossRef]
  77. Noh, H.; Kim, J.E.; Kim, J.Y.; Kim, S.H.; Han, I.; Lim, J.S.; Ki, S.H.; Choi, E.H.; Lee, G.J. Spore viability and cell wall integrity of Cordyceps pruinosa treated with an electric shock-free, atmospheric-pressure air plasma jet. Appl. Sci. 2019, 9, 3921. [Google Scholar] [CrossRef]
  78. Julák, J.; Soušková, H.; Scholtz, V.; Kvasničková, E.; Savická, D.; Kříha, V. Comparison of fungicidal properties of non-thermal plasma produced by corona discharge and dielectric barrier discharge. Folia Microbiol. 2018, 63, 63–68. [Google Scholar] [CrossRef] [PubMed]
  79. Nishime, T.M.C.; Borges, A.C.; Koga-Ito, C.Y.; Machida, M.; Hein, L.R.O.; Kostov, K.G. Non-thermal atmospheric pressure plasma jet applied to inactivation of different microorganisms. Surf. Coat. Technol. 2017, 312, 19–24. [Google Scholar] [CrossRef]
  80. Lee, G.J.; Park, G.; Choi, E.H. Optical and biological properties of plasma-treated Neurospora crassa spores as studied by absorption, circular dichroism, and Raman spectroscopy. J. Korean Phys. Soc. 2017, 71, 670–678. [Google Scholar] [CrossRef]
  81. Recek, N.; Zhou, R.; Zhou, R.; Te’o, V.S.J.; Speight, R.E.; Mozetič, M.; Vesel, A.; Cvelbar, U.; Bazaka, K.; Ostrikov, K. Improved fermentation efficiency of S. cerevisiae by changing glycolytic metabolic pathways with plasma agitation. Sci. Rep. 2018, 8, 8252. [Google Scholar] [CrossRef]
  82. Los, A.; Ziuzina, D.; Boehm, D.; Cullen Patrick, J.; Bourke, P. Inactivation efficacies and mechanisms of gas plasma and plasma-activated water against Aspergillus flavus spores and biofilms: A comparative study. Appl. Environ. Microbiol. 2020, 86, e02619-19. [Google Scholar] [CrossRef]
  83. Borges, A.C.; Castaldelli Nishime, T.M.; Kostov, K.G.; de Morais Gouvêa Lima, G.; Lacerda Gontijo, A.V.; de Carvalho, J.N.M.M.; Yzumi Honda, R.; Yumi Koga-Ito, C. Cold atmospheric pressure plasma jet modulates Candida albicans virulence traits. Clin. Plasma Med. 2017, 7–8, 9–15. [Google Scholar] [CrossRef]
  84. He, M.; Duan, J.; Xu, J.; Ma, M.; Chai, B.; He, G.; Gan, L.; Zhang, S.; Duan, X.; Lu, X.; et al. Candida albicans biofilm inactivated by cold plasma treatment in vitro and in vivo. Plasma Process. Polym. 2020, 17, 1900068. [Google Scholar] [CrossRef]
  85. Bekeschus, S.; Schmidt, A.; Kramer, A.; Metelmann, H.R.; Adler, F.; von Woedtke, T.; Niessner, F.; Weltmann, K.D.; Wende, K. High throughput image cytometry micronucleus assay to investigate the presence or absence of mutagenic effects of cold physical plasma. Environ. Mol. Mutagen. 2018, 59, 268–277. [Google Scholar] [CrossRef]
  86. Nam, S.H.; Ok, S.M.; Kim, G.C. Tooth bleaching with low-temperature plasma lowers surface roughness and Streptococcus mutans adhesion. Int. Endod. J. 2018, 51, 479–488. [Google Scholar] [CrossRef] [PubMed]
  87. Saadati, F.; Mahdikia, H.; Abbaszadeh, H.-A.; Abdollahifar, M.-A.; Khoramgah, M.S.; Shokri, B. Comparison of direct and indirect cold atmospheric-pressure plasma methods in the B16F10 melanoma cancer cells treatment. Sci. Rep. 2018, 8, 7689. [Google Scholar] [CrossRef] [PubMed]
  88. Xu, D.; Xu, Y.; Cui, Q.; Liu, D.; Liu, Z.; Wang, X.; Yang, Y.; Feng, M.; Liang, R.; Chen, H.; et al. Cold atmospheric plasma as a potential tool for multiple myeloma treatment. Oncotarget 2018, 9, 18002–18017. [Google Scholar] [CrossRef] [PubMed]
  89. Brown, G.D.; Denning, D.W.; Gow, N.A.; Levitz, S.M.; Netea, M.G.; White, T.C. Hidden killers: Human fungal infections. Sci. Transl. Med. 2012, 4, 165rv13. [Google Scholar] [CrossRef]
  90. Da Silva, L.C.F.; de Almeida Freitas, R.; de Andrade, M.P., Jr.; Piva, M.R.; Martins-Filho, P.R.S.; de Santana Santos, T. Oral lesions in renal transplant. J. Craniofac. Surg. 2012, 23, e214–e218. [Google Scholar] [CrossRef]
  91. De la Rosa García, E.; de la Rosa-García, E.; Miramontes Zapata, M.; Miramontes-Zapata, M.; Sanchez-Vargas, L.-O.; Sánchez-Vargas, L.O.; Mondragón Padilla, A.; Mondragón-Padilla, A. Oral colonisation and infection by Candida sp. in diabetic and non-diabetic patients with chronic kidney disease on dialysis. Nefrología 2013, 33, 764–770. [Google Scholar] [CrossRef]
  92. Borges, A.C.; Lima, G.d.M.G.; Nishime, T.M.C.; Gontijo, A.V.L.; Kostov, K.G.; Koga-Ito, C.Y. Amplitude-modulated cold atmospheric pressure plasma jet for treatment of oral candidiasis: In Vivo study. PLoS ONE 2018, 13, e0199832. [Google Scholar] [CrossRef]
  93. Park, N.-S.; Yun, S.-E.; Lee, H.-Y.; Lee, H.J.; Choi, J.-H.; Kim, G.-C. No-ozone cold plasma can kill oral pathogenic microbes in H2O2-dependent and independent manner. Sci. Rep. 2022, 12, 7597. [Google Scholar] [CrossRef]
  94. Smith, M.B.; McGinnis, M.R. CHAPTER 82—Dermatophytosi. In Tropical Infectious Diseases: Principles, Pathogens and Practice, 3rd ed.; Guerrant, R.L., Walker, D.H., Weller, P.F., Eds.; W.B. Saunders: Edinburgh, UK, 2011; pp. 559–564. [Google Scholar]
  95. Ouf, S.A.; El-Adly, A.A.; Mohamed, A.-A.H. Inhibitory effect of silver nanoparticles mediated by atmospheric pressure air cold plasma jet against dermatophyte fungi. J. Med. Microbiol. 2015, 64, 1151–1161. [Google Scholar] [CrossRef]
  96. Scholtz, V.; Soušková, H.; Švarcová, M.; Kříha, V.; Živná, H.; Julák, J. Inactivation of dermatophyte infection by nonthermal plasma on animal model. Med. Mycol. 2016, 55, 422–428. [Google Scholar] [CrossRef]
  97. Dubljanin, E.; Džamić, A.; Vujčić, I.; Grujičić, S.Š.; Arsenijević, V.A.; Mitrović, S.; Čalovski, I.Č. Epidemiology of onychomycosis in Serbia: A laboratory-based survey and risk factor identification. Mycoses 2017, 60, 25–32. [Google Scholar] [CrossRef] [PubMed]
  98. Maraki, S.; Mavromanolaki, V.E. Epidemiology of onychomycosis in Crete, Greece: A 12-year study. Mycoses 2016, 59, 798–802. [Google Scholar] [CrossRef] [PubMed]
  99. Lipner, S.R.; Scher, R.K. Onychomycosis: Clinical overview and diagnosis. J. Am. Acad. Dermatol. 2019, 80, 835–851. [Google Scholar] [CrossRef] [PubMed]
  100. Bulson, J.M.; Liveris, D.; Derkatch, I.; Friedman, G.; Geliebter, J.; Park, S.; Singh, S.; Zemel, M.; Tiwari, R.K. Non-thermal atmospheric plasma treatment of onychomycosis in an in vitro human nail model. Mycoses 2020, 63, 225–232. [Google Scholar] [CrossRef] [PubMed]
  101. Lux, J.; Dobiáš, R.; Kuklová, I.; Litvik, R.; Scholtz, V.; Soušková, H.; Khun, J.; Mrázek, J.; Kantorová, M.; Jaworská, P.; et al. Inactivation of dermatophytes causing onychomycosis and its therapy using non-thermal plasma. J. Fungi 2020, 6, 214. [Google Scholar] [CrossRef]
  102. Nwabor, O.F.; Onyeaka, H.; Miri, T.; Obileke, K.; Anumudu, C.; Hart, A. A cold plasma technology for ensuring the microbiological safety and quality of foods. Food Eng. Rev. 2022, 14, 1–20. [Google Scholar] [CrossRef]
  103. Jadhav, H.B.; Annapure, U. Consequences of non-thermal cold plasma treatment on meat and dairy lipids—A review. Future Foods 2021, 4, 100095. [Google Scholar] [CrossRef]
  104. Shelar, A.; Singh, A.V.; Dietrich, P.; Maharjan, R.S.; Thissen, A.; Didwal, P.N.; Shinde, M.; Laux, P.; Luch, A.; Mathe, V.; et al. Emerging cold plasma treatment and machine learning prospects for seed priming: A step towards sustainable food production. RSC Adv. 2022, 12, 10467–10488. [Google Scholar] [CrossRef]
  105. Park, S.Y.; Ha, S.-D. Application of cold oxygen plasma for the reduction of Cladosporium cladosporioides and Penicillium citrinum on the surface of dried filefish (Stephanolepis cirrhifer) fillets. Int. J. Food Sci. Technol. 2015, 50, 966–973. [Google Scholar] [CrossRef]
  106. Yong, H.I.; Lee, H.; Park, S.; Park, J.; Choe, W.; Jung, S.; Jo, C. Flexible thin-layer plasma inactivation of bacteria and mold survival in beef jerky packaging and its effects on the meat’s physicochemical properties. Meat Sci. 2017, 123, 151–156. [Google Scholar] [CrossRef]
  107. Royintarat, T.; Choi, E.H.; Boonyawan, D.; Seesuriyachan, P.; Wattanutchariya, W. Chemical-free and synergistic interaction of ultrasound combined with plasma-activated water (PAW) to enhance microbial inactivation in chicken meat and skin. Sci. Rep. 2020, 10, 1559. [Google Scholar] [CrossRef] [PubMed]
  108. Sudheesh, C.; Sunooj, K.V. 14—Cold plasma processing of fresh-cut fruits and vegetables. In Fresh-Cut Fruits and Vegetables; Siddiqui, M.W., Ed.; Academic Press: Cambridge, MA, USA, 2020; pp. 339–356. [Google Scholar]
  109. Misra, N.N.; Pankaj, S.K.; Segat, A.; Ishikawa, K. Cold plasma interactions with enzymes in foods and model systems. Trends Food Sci. Technol. 2016, 55, 39–47. [Google Scholar] [CrossRef]
  110. Coutinho, N.M.; Silveira, M.R.; Rocha, R.S.; Moraes, J.; Ferreira, M.V.S.; Pimentel, T.C.; Freitas, M.Q.; Silva, M.C.; Raices, R.S.L.; Ranadheera, C.S.; et al. Cold plasma processing of milk and dairy products. Trends Food Sci. Technol. 2018, 74, 56–68. [Google Scholar] [CrossRef]
  111. Nierop Groot, M.; Abee, T.; van Bokhorst-van de Veen, H. Inactivation of conidia from three Penicillium spp. isolated from fruit juices by conventional and alternative mild preservation technologies and disinfection treatments. Food Microbiol. 2019, 81, 108–114. [Google Scholar] [CrossRef]
  112. Xiang, Q.; Liu, X.; Li, J.; Liu, S.; Zhang, H.; Bai, Y. Effects of dielectric barrier discharge plasma on the inactivation of Zygosaccharomyces rouxii and quality of apple juice. Food Chem. 2018, 254, 201–207. [Google Scholar] [CrossRef] [PubMed]
  113. Starek, A.; Sagan, A.; Andrejko, D.; Chudzik, B.; Kobus, Z.; Kwiatkowski, M.; Terebun, P.; Pawłat, J. Possibility to extend the shelf life of NFC tomato juice using cold atmospheric pressure plasma. Sci. Rep. 2020, 10, 20959. [Google Scholar] [CrossRef]
  114. Muhammad, A.I.; Liao, X.; Cullen, P.J.; Liu, D.; Xiang, Q.; Wang, J.; Chen, S.; Ye, X.; Ding, T. Effects of nonthermal plasma technology on functional food components. Compr. Rev. Food Sci. Food Saf. 2018, 17, 1379–1394. [Google Scholar] [CrossRef] [PubMed]
  115. Hu, X.; Sun, H.; Yang, X.; Cui, D.; Wang, Y.; Zhuang, J.; Wang, X.; Ma, R.; Jiao, Z. Potential use of atmospheric cold plasma for postharvest preservation of blueberries. Postharvest Biol. Technol. 2021, 179, 111564. [Google Scholar] [CrossRef]
  116. Puligundla, P.; Lee, T.; Mok, C. Effect of intermittent corona discharge plasma treatment for improving microbial quality and shelf life of kumquat (Citrus japonica) fruits. LWT 2018, 91, 8–13. [Google Scholar] [CrossRef]
  117. Go, S.-M.; Park, M.-R.; Kim, H.-S.; Choi, W.S.; Jeong, R.-D. Antifungal effect of non-thermal atmospheric plasma and its application for control of postharvest Fusarium oxysporum decay of paprika. Food Control 2019, 98, 245–252. [Google Scholar] [CrossRef]
  118. Xiang, Q.; Liu, X.; Liu, S.; Ma, Y.; Xu, C.; Bai, Y. Effect of plasma-activated water on microbial quality and physicochemical characteristics of mung bean sprouts. Innov. Food Sci. Emerg. Technol. 2019, 52, 49–56. [Google Scholar] [CrossRef]
  119. Won, M.Y.; Lee, S.J.; Min, S.C. Mandarin preservation by microwave-powered cold plasma treatment. Innov. Food Sci. Emerg. Technol. 2017, 39, 25–32. [Google Scholar] [CrossRef]
  120. Liu, C.-M.; Nishida, Y.; Iwasaki, K.; Ting, K. Prolonged preservation and sterilization of fresh plants in controlled environments using high-field plasma. IEEE Trans. Plasma Sci. 2011, 39, 717–724. [Google Scholar] [CrossRef]
  121. Ambrico, P.F.; Šimek, M.; Rotolo, C.; Morano, M.; Minafra, A.; Ambrico, M.; Pollastro, S.; Gerin, D.; Faretra, F.; De Miccolis Angelini, R.M. Surface dielectric barrier discharge plasma: A suitable measure against fungal plant pathogens. Sci. Rep. 2020, 10, 3673. [Google Scholar] [CrossRef]
  122. Phan, K.T.K.; Phan, H.T.; Boonyawan, D.; Intipunya, P.; Brennan, C.S.; Regenstein, J.M.; Phimolsiripol, Y. Non-thermal plasma for elimination of pesticide residues in mango. Innov. Food Sci. Emerg. Technol. 2018, 48, 164–171. [Google Scholar] [CrossRef]
  123. Wu, Y.; Cheng, J.-H.; Sun, D.-W. Subcellular damages of Colletotrichum asianum and inhibition of mango anthracnose by dielectric barrier discharge plasma. Food Chem. 2022, 381, 132197. [Google Scholar] [CrossRef]
  124. Cui, D.; Hu, X.; Yin, Y.; Zhu, Y.; Zhuang, J.; Wang, X.; Ma, R.; Jiao, Z. Quality enhancement and microbial reduction of mung bean (Vigna radiata) sprouts by non-thermal plasma pretreatment of seeds. Plasma Sci. Technol. 2022, 24, 45504. [Google Scholar] [CrossRef]
  125. Choi, E.J.; Park, H.W.; Kim, S.B.; Ryu, S.; Lim, J.; Hong, E.J.; Byeon, Y.S.; Chun, H.H. Sequential application of plasma-activated water and mild heating improves microbiological quality of ready-to-use shredded salted kimchi cabbage (Brassica pekinensis L.). Food Control 2019, 98, 501–509. [Google Scholar] [CrossRef]
  126. Hosseini, S.I.; Farrokhi, N.; Shokri, K.; Khani, M.R.; Shokri, B. Cold low pressure O2 plasma treatment of Crocus sativus: An efficient way to eliminate toxicogenic fungi with minor effect on molecular and cellular properties of saffron. Food Chem. 2018, 257, 310–315. [Google Scholar] [CrossRef]
  127. Kim, J.E.; Oh, Y.J.; Won, M.Y.; Lee, K.-S.; Min, S.C. Microbial decontamination of onion powder using microwave-powered cold plasma treatments. Food Microbiol. 2017, 62, 112–123. [Google Scholar] [CrossRef]
  128. Sakudo, A.; Yagyu, Y. Application of a roller conveyor type plasma disinfection device with fungus-contaminated citrus fruits. AMB Express 2021, 11, 16. [Google Scholar] [CrossRef] [PubMed]
  129. Wiktor, A.; Hrycak, B.; Jasiński, M.; Rybak, K.; Kieliszek, M.; Kraśniewska, K.; Witrowa-Rajchert, D. Impact of atmospheric pressure microwave plasma treatment on quality of selected spices. Appl. Sci. 2020, 10, 6815. [Google Scholar] [CrossRef]
  130. Phan, K.T.K.; Phan, H.T.; Brennan, C.S.; Regenstein, J.M.; Jantanasakulwong, K.; Boonyawan, D.; Phimolsiripol, Y. Gliding arc discharge non-thermal plasma for retardation of mango anthracnose. LWT 2019, 105, 142–148. [Google Scholar] [CrossRef]
  131. Sen, Y.; Onal-Ulusoy, B.; Mutlu, M. Aspergillus decontamination in hazelnuts: Evaluation of atmospheric and low-pressure plasma technology. Innov. Food Sci. Emerg. Technol. 2019, 54, 235–242. [Google Scholar] [CrossRef]
  132. Dong, X.Y.; Yang, Y.L. A novel approach to enhance blueberry quality during storage using cold plasma at atmospheric air pressure. Food Bioprocess Technol. 2019, 12, 1409–1421. [Google Scholar] [CrossRef]
  133. Dasan, B.G.; Boyaci, I.H.; Mutlu, M. Nonthermal plasma treatment of Aspergillus spp. spores on hazelnuts in an atmospheric pressure fluidized bed plasma system: Impact of process parameters and surveillance of the residual viability of spores. J. Food Eng. 2017, 196, 139–149. [Google Scholar] [CrossRef]
  134. Hojnik, N.; Modic, M.; Žigon, D.; Kovač, J.; Jurov, A.; Dickenson, A.; Walsh, J.L.; Cvelbar, U. Cold atmospheric pressure plasma-assisted removal of aflatoxin B1 from contaminated corn kernels. Plasma Process. Polym. 2021, 18, 2000163. [Google Scholar] [CrossRef]
  135. Makari, M.; Hojjati, M.; Shahbazi, S.; Askari, H. Elimination of Aspergillus flavus from pistachio nuts with dielectric barrier discharge (DBD) cold plasma and its impacts on biochemical indices. J. Food Qual. 2021, 2021, 9968711. [Google Scholar] [CrossRef]
  136. Puligundla, P.; Lee, T.; Mok, C. Effect of corona discharge plasma jet treatment on the degradation of aflatoxin B1 on glass slides and in spiked food commodities. LWT 2020, 124, 108333. [Google Scholar] [CrossRef]
  137. Sen, Y.; Onal-Ulusoy, B.; Mutlu, M. Detoxification of hazelnuts by different cold plasmas and gamma irradiation treatments. Innov. Food Sci. Emerg. Technol. 2019, 54, 252–259. [Google Scholar] [CrossRef]
  138. Jerushalmi, S.; Maymon, M.; Dombrovsky, A.; Freeman, S. Effects of cold plasma, gamma and e-beam irradiations on reduction of fungal colony forming unit levels in medical cannabis inflorescences. J. Cannabis Res. 2020, 2, 12. [Google Scholar] [CrossRef]
  139. Lee, Y.; Lee, Y.Y.; Kim, Y.S.; Balaraju, K.; Mok, Y.S.; Yoo, S.J.; Jeon, Y. Enhancement of seed germination and microbial disinfection on ginseng by cold plasma treatment. J. Ginseng Res. 2021, 45, 519–526. [Google Scholar] [CrossRef] [PubMed]
  140. Świecimska, M.; Tulik, M.; Šerá, B.; Golińska, P.; Tomeková, J.; Medvecká, V.; Bujdáková, H.; Oszako, T.; Zahoranová, A.; Šerý, M. Non-thermal plasma can be used in disinfection of scots pine (Pinus sylvestris L.) seeds infected with Fusarium oxysporum. Forests 2020, 11, 837. [Google Scholar] [CrossRef]
  141. Waskow, A.; Betschart, J.; Butscher, D.; Oberbossel, G.; Klöti, D.; Büttner-Mainik, A.; Adamcik, J.; von Rohr, P.R.; Schuppler, M. Characterization of efficiency and mechanisms of cold atmospheric pressure plasma decontamination of seeds for sprout production. Front. Microbiol. 2018, 9, 3164. [Google Scholar] [CrossRef] [PubMed]
  142. Pérez Pizá, M.C.; Prevosto, L.; Zilli, C.; Cejas, E.; Kelly, H.; Balestrasse, K. Effects of non–thermal plasmas on seed-borne Diaporthe/Phomopsis complex and germination parameters of soybean seeds. Innov. Food Sci. Emerg. Technol. 2018, 49, 82–91. [Google Scholar] [CrossRef]
  143. Štěpánová, V.; Slavíček, P.; Kelar, J.; Prášil, J.; Smékal, M.; Stupavská, M.; Jurmanová, J.; Černák, M. Atmospheric pressure plasma treatment of agricultural seeds of cucumber (Cucumis sativus L.) and pepper (Capsicum annuum L.) with effect on reduction of diseases and germination improvement. Plasma Process. Polym. 2018, 15, 1700076. [Google Scholar] [CrossRef]
  144. Ambrico, P.F.; Šimek, M.; Morano, M.; De Miccolis Angelini, R.M.; Minafra, A.; Trotti, P.; Ambrico, M.; Prukner, V.; Faretra, F. Reduction of microbial contamination and improvement of germination of sweet basil (Ocimum basilicum L.) seeds via surface dielectric barrier discharge. J. Phys. D Appl. Phys. 2017, 50, 305401. [Google Scholar] [CrossRef]
  145. Ochi, A.; Konishi, H.; Ando, S.; Sato, K.; Yokoyama, K.; Tsushima, S.; Yoshida, S.; Morikawa, T.; Kaneko, T.; Takahashi, H. Management of bakanae and bacterial seedling blight diseases in nurseries by irradiating rice seeds with atmospheric plasma. Plant Pathol. 2017, 66, 67–76. [Google Scholar] [CrossRef]
  146. Devi, Y.; Thirumdas, R.; Sarangapani, C.; Deshmukh, R.R.; Annapure, U.S. Influence of cold plasma on fungal growth and aflatoxins production on groundnuts. Food Control 2017, 77, 187–191. [Google Scholar] [CrossRef]
  147. Kim, J.-W.; Puligundla, P.; Mok, C. Effect of corona discharge plasma jet on surface-borne microorganisms and sprouting of broccoli seeds. J. Sci. Food Agric. 2017, 97, 128–134. [Google Scholar] [CrossRef]
  148. Marin, S.; Ramos, A.J.; Cano-Sancho, G.; Sanchis, V. Mycotoxins: Occurrence, toxicology, and exposure assessment. Food Chem. Toxicol. 2013, 60, 218–237. [Google Scholar] [CrossRef] [PubMed]
  149. Stoev, S.D. Food safety and increasing hazard of mycotoxin occurrence in foods and feeds. Crit. Rev. Food Sci. Nutr. 2013, 53, 887–901. [Google Scholar] [CrossRef] [PubMed]
  150. Hojnik, N.; Modic, M.; Walsh, J.L.; Žigon, D.; Javornik, U.; Plavec, J.; Žegura, B.; Filipič, M.; Cvelbar, U. Unravelling the pathways of air plasma induced aflatoxin B1 degradation and detoxification. J. Hazard. Mater. 2021, 403, 123593. [Google Scholar] [CrossRef] [PubMed]
  151. Scholtz, V.; Šerá, B.; Khun, J.; Šerý, M.; Julák, J. Effects of nonthermal plasma on wheat grains and products. J. Food Qual. 2019, 2019, 7917825. [Google Scholar] [CrossRef]
  152. Švubová, R.; Slováková, Ľ.; Holubová, Ľ.; Rovňanová, D.; Gálová, E.; Tomeková, J. Evaluation of the impact of cold atmospheric pressure plasma on soybean seed germination. Plants 2021, 10, 177. [Google Scholar] [CrossRef]
  153. Peťková, M.; Švubová, R.; Kyzek, S.; Medvecká, V.; Slováková, Ľ.; Ševčovičová, A.; Gálová, E. The effects of cold atmospheric pressure plasma on germination parameters, enzyme activities and induction of DNA damage in barley. Int. J. Mol. Sci. 2021, 22, 2833. [Google Scholar] [CrossRef]
  154. Kučerová, K.; Henselová, M.; Slováková, Ľ.; Hensel, K. Effects of plasma activated water on wheat: Germination, growth parameters, photosynthetic pigments, soluble protein content, and antioxidant enzymes activity. Plasma Process. Polym. 2019, 16, 1800131. [Google Scholar] [CrossRef]
  155. Sivachandiran, L.; Khacef, A. Enhanced seed germination and plant growth by atmospheric pressure cold air plasma: Combined effect of seed and water treatment. RSC Adv. 2017, 7, 1822–1832. [Google Scholar] [CrossRef]
  156. Šerý, M.; Zahoranová, A.; Kerdík, A.; Šerá, B. Seed germination of black pine (Pinus nigra Arnold) after diffuse coplanar surface barrier discharge plasma treatment. IEEE Trans. Plasma Sci. 2020, 48, 939–945. [Google Scholar] [CrossRef]
  157. Zahoranová, A.; Henselová, M.; Hudecová, D.; Kaliňáková, B.; Kováčik, D.; Medvecká, V.; Černák, M. Effect of cold atmospheric pressure plasma on the wheat seedlings vigor and on the inactivation of microorganisms on the seeds surface. Plasma Chem. Plasma Process. 2016, 36, 397–414. [Google Scholar] [CrossRef]
  158. Hoppanová, L.; Medvecká, V.; Dylíková, J.; Hudecová, D.; Kaliňáková, B.; Kryštofová, S.; Zahoranová, A. Low-temperature plasma applications in chemical fungicide treatment reduction. Acta Chim. Slovaca 2020, 13, 26–33. [Google Scholar] [CrossRef]
  159. Todorova, Y.; Benova, E.; Marinova, P.; Yotinov, I.; Bogdanov, T.; Topalova, Y. Non-thermal atmospheric plasma for microbial decontamination and removal of hazardous chemicals: An overview in the circular economy context with data for test applications of microwave plasma torch. Processes 2022, 10, 554. [Google Scholar] [CrossRef]
  160. Sterflinger, K. Fungi: Their role in deterioration of cultural heritage. Fungal Biol. Rev. 2010, 24, 47–55. [Google Scholar] [CrossRef]
  161. Rehakova, M.; Čeppan, M.; Mikula, M. Study of stabilization of documents containing iron gall inks by treatment of atmospheric DBD N2 plasma. Chem. Listy 2008, 102, 1061–1063. [Google Scholar]
  162. Vizárová, K.; Kaliňáková, B.; Tiňo, R.; Vajová, I.; Čížová, K. Microbial decontamination of lignocellulosic materials with low-temperature atmospheric plasma. J. Cult. Herit. 2021, 47, 28–33. [Google Scholar] [CrossRef]
  163. Huang, W.W.; Ge, X.Y.; Huang, Y.; Chai, X.T.; Zhang, L.; Zhang, Y.X.; Deng, L.N.; Liu, C.Q.; Xu, H.; Gao, J. High-yield strain of fusidic acid obtained by atmospheric and room temperature plasma mutagenesis and the transcriptional changes involved in improving its production in fungus Fusidium coccineum. J. Appl. Microbiol. 2021, 130, 405–415. [Google Scholar] [CrossRef]
  164. Hu, Z.-C.; Li, W.-J.; Zou, S.-P.; Niu, K.; Zheng, Y.-G. Mutagenesis of echinocandin B overproducing Aspergillus nidulans capable of using starch as main carbon source. Prep. Biochem. Biotechnol. 2020, 50, 745–752. [Google Scholar] [CrossRef] [PubMed]
  165. Li, T.; Chen, L.; Wu, D.; Dong, G.; Chen, W.; Zhang, H.; Yang, Y.; Wu, W. The structural characteristics and biological activities of intracellular polysaccharide derived from mutagenic Sanghuangporous sanghuang strain. Molecules 2020, 25, 3693. [Google Scholar] [CrossRef]
  166. Gu, L.-S.; Tan, M.-Z.; Li, S.-H.; Zhang, T.; Zhang, Q.-Q.; Li, C.-X.; Luo, X.-M.; Feng, J.-X.; Zhao, S. ARTP/EMS-combined multiple mutagenesis efficiently improved production of raw starch-degrading enzymes in Penicillium oxalicum and characterization of the enzyme-hyperproducing mutant. Biotechnol. Biofuels 2020, 13, 187. [Google Scholar] [CrossRef]
  167. Shu, L.; Si, X.; Yang, X.; Ma, W.; Sun, J.; Zhang, J.; Xue, X.; Wang, D.; Gao, Q. Enhancement of acid protease activity of Aspergillus oryzae using atmospheric and room temperature plasma. Front. Microbiol. 2020, 11, 1418. [Google Scholar] [CrossRef]
  168. Ma, X.-J.; Zhang, H.-M.; Lu, X.-F.; Han, J.; Zhu, H.-X.; Wang, H.; Yao, R.-S. Mutant breeding of Starmerella bombicola by atmospheric and room-temperature plasma (ARTP) for improved production of specific or total sophorolipids. Bioprocess Biosyst. Eng. 2020, 43, 1869–1883. [Google Scholar] [CrossRef] [PubMed]
  169. Zheng, S.; Jiang, B.; Zhang, T.; Chen, J. Combined mutagenesis and metabolic regulation to enhance d-arabitol production from Candida parapsilosis. J. Ind. Microbiol. Biotechnol. 2020, 47, 425–435. [Google Scholar] [CrossRef] [PubMed]
  170. Zhang, C.; Qin, J.; Dai, Y.; Mu, W.; Zhang, T. Atmospheric and room temperature plasma (ARTP) mutagenesis enables xylitol over-production with yeast Candida tropicalis. J. Biotechnol. 2019, 296, 7–13. [Google Scholar] [CrossRef] [PubMed]
  171. Feng, W.; Liang, J.; Wang, B.; Chen, J. Improvement of kojic acid production in Aspergillus oryzae AR-47 mutant strain by combined mutagenesis. Bioprocess Biosyst. Eng. 2019, 42, 753–761. [Google Scholar] [CrossRef]
  172. Zhu, L.; Wu, D.; Zhang, H.; Li, Q.; Zhang, Z.; Liu, Y.; Zhou, S.; Wang, W.; Li, Z.; Yang, Y. Effects of atmospheric and room temperature plasma (ARTP) mutagenesis on physicochemical characteristics and immune activity in vitro of Hericium erinaceus polysaccharides. Molecules 2019, 24, 262. [Google Scholar] [CrossRef]
  173. Ma, Y.; Zhang, Q.; Zhang, Q.; He, H.; Chen, Z.; Zhao, Y.; Wei, D.; Kong, M.; Huang, Q. Improved production of polysaccharides in Ganoderma lingzhi mycelia by plasma mutagenesis and rapid screening of mutated strains through infrared spectroscopy. PLoS ONE 2018, 13, e0204266. [Google Scholar] [CrossRef]
  174. Zou, Z.; Zhao, Y.; Zhang, T.; Xu, J.; He, A.; Deng, Y. Efficient isolation and characterization of a cellulase hyperproducing mutant strain of Trichoderma reesei. J. Microbiol. Biotechnol. 2018, 28, 1473–1481. [Google Scholar] [CrossRef]
  175. Qi, F.; Zhao, X.; Kitahara, Y.; Li, T.; Ou, X.; Du, W.; Liu, D.; Huang, J. Integrative transcriptomic and proteomic analysis of the mutant lignocellulosic hydrolyzate-tolerant Rhodosporidium toruloides. Eng. Life Sci. 2017, 17, 249–261. [Google Scholar] [CrossRef]
  176. Luo, Z.; Zeng, W.; Du, G.; Liu, S.; Fang, F.; Zhou, J.; Chen, J. A high-throughput screening procedure for enhancing pyruvate production in Candida glabrata by random mutagenesis. Bioprocess Biosyst. Eng. 2017, 40, 693–701. [Google Scholar] [CrossRef]
  177. Zhu, X.; Arman, B.; Chu, J.; Wang, Y.; Zhuang, Y. Development of a method for efficient cost-effective screening of Aspergillus niger mutants having increased production of glucoamylase. Biotechnol. Lett. 2017, 39, 739–744. [Google Scholar] [CrossRef]
  178. Wang, Q.; Liu, D.; Yang, Q.; Wang, P. Enhancing carotenoid production in Rhodotorula mucilaginosa KC8 by combining mutation and metabolic engineering. Ann. Microbiol. 2017, 67, 425–431. [Google Scholar] [CrossRef]
  179. Liu, X.; Lv, J.; Xu, J.; Xia, J.; Dai, B.; Xu, X.; Xu, J. Erythritol production by Yarrowia lipolytica mutant strain M53 generated through atmospheric and room temperature plasma mutagenesis. Food Sci. Biotechnol. 2017, 26, 979–986. [Google Scholar] [CrossRef] [PubMed]
  180. Zhang, X.; Zhang, X.-F.; Li, H.-P.; Wang, L.-Y.; Zhang, C.; Xing, X.-H.; Bao, C.-Y. Atmospheric and room temperature plasma (ARTP) as a new powerful mutagenesis tool. Appl. Microbiol. Biotechnol. 2014, 98, 5387–5396. [Google Scholar] [CrossRef] [PubMed]
  181. Xu, W.; Jia, H.; Zhang, L.; Wang, H.; Tang, H.; Zhang, L. Effects of GSH1 and GSH2 gene mutation on glutathione synthetases activity of Saccharomyces cerevisiae. Protein J. 2017, 36, 270–277. [Google Scholar] [CrossRef] [PubMed]
  182. Farasat, M.; Arjmand, S.; Ranaei Siadat, S.O.; Sefidbakht, Y.; Ghomi, H. The effect of non-thermal atmospheric plasma on the production and activity of recombinant phytase enzyme. Sci. Rep. 2018, 8, 16647. [Google Scholar] [CrossRef]
  183. Veerana, M.; Lim, J.-S.; Choi, E.-H.; Park, G. Aspergillus oryzae spore germination is enhanced by non-thermal atmospheric pressure plasma. Sci. Rep. 2019, 9, 11184. [Google Scholar] [CrossRef]
  184. Veerana, M.; Choi, E.H.; Park, G. Influence of non-thermal atmospheric pressure plasma jet on extracellular activity of α-amylase in Aspergillus oryzae. Appl. Sci. 2021, 11, 691. [Google Scholar] [CrossRef]
Figure 1. Configuration of basic NTP systems: (A) corona discharge; (B) dielectric barrier discharge; (C) plasma jet (adapted and modified from [14]).
Figure 1. Configuration of basic NTP systems: (A) corona discharge; (B) dielectric barrier discharge; (C) plasma jet (adapted and modified from [14]).
Ijms 23 11592 g001
Figure 2. NTP and fungi: treatment setups, applications, and fungal response.
Figure 2. NTP and fungi: treatment setups, applications, and fungal response.
Ijms 23 11592 g002
Table 1. Studies about fungal inactivation, growth inhibition, and biofilm formation.
Table 1. Studies about fungal inactivation, growth inhibition, and biofilm formation.
NTP TypeProcess GasTime of
Treatment
Fungus/YeastEffectRef.
RF plasma jetA mixture of argon and oxygen1–10 minAspergillus flavus100% inhibition of growth after 10 min treatment[69]
Plasma jetArgon0–10 minCandida parapsilosis
Magnusiomyces magnusii
Saccharomyces cerevisiae
Schizosaccharomyces pombe
More than 90% inactivation of yeast cells after 10 min[70]
Plasma microjetA mixture of helium and oxygen0–5 minSaccharomyces cerevisiaeThe survival ratio of cells in water was significantly decreased from 40.2% to 1.5% after 5 min[71]
PAW with the plasma jetAir1–6 min water activation by plasmaAspergillus brasiliensisThe spore viability dropped to 15% after 30 min in the PAW with a plasma activation time of 3 min[72]
Linear micro discharge plasma jetHelium1 minCandida albicansChanges in both the genotype and phenotype[73]
DBSD plasmaAir0–480 sAspergillus flavusA 5 log reduction of spore viability after 480 s under both the low and high power [56]
Surface micro-discharge plasmaHelium0–10 minSaccharomyces cerevisiaeThe reduction in CFU was about 3.4 log after plasma treatment for 10 min[74]
DBD plasmaArgon0–60 minAureobasidium pullulansThe non-melanized cells were efficiently inactivated, and more than 60% of melanized cells were still alive after the 60 min exposure[75]
PAW with the CD plasma jetAir or 99,99% oxygen0–30 minColletotrichum gloeosporioides96% inactivation after 30 min incubation in air-PAW; 55% inactivation after 30 min incubation in oxygen-PAW[76]
Electric shock-free plasma jetAir0–6 minCordyceps pruinosa~100% inactivation of spore viability after 6 min[77]
CD plasma, DBD plasmaAir0–30 minAlternaria sp.
Aspergillus oryzae
Byssochlamys nivea
Cladosporium sphaerospermum
Spore inhibition after 10–40 min[78]
Plasma jetHelium0–180 sCandida albicans20–30 mm2 inhibition zone area after 180 s[79]
Plasma jetArgon0–180 sNeurospora crassaOnly ~5% spore viability after 3 min in water[80]
CD, corona discharge; DBD, dielectric barrier discharge; DBSD, dielectric barrier surface discharge; PAW, plasma-activated water; RF, radiofrequency.
Table 2. The application of plasma in food and agriculture.
Table 2. The application of plasma in food and agriculture.
NTP TypeProcess GasTime of TreatmentTreated SampleFungus/YeastEffectRef.
DBD plasmaAir0–9 minMangoColletotrichum asianumThe disease incidence and lesion diameter of mango treated for 9 min were decreased by 48.00% and 62.95%, respectively[123]
Plasma jetArgon, oxygen, nitrogen0–6 minMung beanNatural fungal contaminationReduction in natural fungal contamination ranging from 0.54 to 7.09 log at 96 h incubation[124]
Gliding arc plasmaNitrogen300–600 s Tomato juiceCandida albicans
Saccharomyces cerevisiae
600 s treatment—reduction in fungal viability below the limit of quantification; extension of shelf life to 10 days[113]
PAWAir30–120 sKimchi cabbageNatural fungal contaminationPAW treated with plasma for 120s caused a 1.8 log CFU/g reduction in fungal contamination[125]
RF cold plasmaOxygen0–15 minSaffronAspergillus sp.
Penicillium sp.
Rhizopus sp.
Complete reduction in contamination after 15 min of treatment[126]
Microwave plasmaHelium40 minOnion powderAspergillus brasiliensis1.6 log spores/cm2 reduction[127]
Flexible thin-layer plasmaAir10 min Beef jerky packagedAspergillus flavus2–3 log CFU/g reduction in spore viability[106]
DBD plasmaAir0–5 minCitrusPenicillium venetum
Aspergillus brasiliensis
Significantly decreased to ~1.50 log CFU/mL at 2 min; significantly decreased viable count to ~1.62 log CFU/mL at 5 min[128]
DBSD plasmaAir0–20 minBlueberry Botrytis cinerea15 and 20 min plasma treatment completely inhibited the mycelial growth [115]
Microwave plasmaAir15–60 sAllspice berry
Black pepper
Juniper berry
Aspergillus nigerPartial inactivation after 15 s treatment[129]
Gliding arc plasmaHumid argon0–7 minMangoColletotrichum gloeosporioidesSignificantly lower mycelium growth rate constant, the maximum reduction in spores was 1 log spore/mL after 7 min of NTP treatment with 5 L/min gas flux[130]
PAW with the plasma jetAir0–30 minMung bean sproutNatural yeast contamination~2.8 log CFU/g yeasts reduction after 30 min[118]
Plasma jetAir0–90 sPaprikaFusarium oxysporumComplete inhibition of mycelial growth and spore germination after 90 s of treatment but only 50% inhibition of fungal growth on the paprika surface[117]
AP plasma jet; LP RF plasmaAir, nitrogen,
Oxygen
0–30 minHazelnutAspergillus flavus
Aspergillus parasiticus
Spore reductions of 4.7 and 5.6 log CFU/g after 30 min of LP air plasma treatment; spore reductions of 5.4 and 5.5 log CFU/g after 1.7 min of AP air plasma treatment; deformation of spores and loss of spore integrity after plasma treatments[131]
DBD plasmaAir0–10 minBlueberrynatural fungal contaminationThe number of fungi decreased by 25.8%; the blueberry decay rates were reduced by 5.2% in the plasma treatment of 10 min after 20 days of storage[132]
CD plasma jetAir 0–120 sKumquatnatural yeasts contamination0.77–1.57 log CFU/g reduction after 120 s treatment[116]
Fluidized bed plasmaAir, nitrogen0–5 minHazelnutsAspergillus flavus
Aspergillus parasiticus
~4 log fungicidal effects after 5 min; the air plasma was more effective than nitrogen plasma[133]
Surface barrier dischargeAir0–8 minCorn kernels-Complete degradation of aflatoxin B1 after 6 min of treatment[134]
DBD plasmaAir0–180 sPistachio nuts, glass slidesAspergillus flavusDecrease in spore population by 4 log after 180 s of the treatment; maximum reduction in AFB1 was observed after 180 s of the treatment, which was 64.63% for glass slides and 52.42% for pistachio nuts[135]
CD plasma jetAir0–30 minRice,
Wheat
-Initial AFB1 concentration on slides was decreased maximally by 95% in 30 min; in rice and wheat, the average levels of AFB1 degradation ranged between 45 and 56% following 30 min treatment[136]
AP plasma jet; LP RF plasmaAir0–30 minHazelnuts-Both plasmas reduced 72–73% of AFB1 spiked on hazelnuts after plasma treatment[137]
RF plasmaair with H2O2 (35%)0–10 minCannabis inflorescencesBotrytis cinerea5 log reduction in viable fungal spores after 10 min[138]
DBD plasmaArgon or a mixture of 80% argon and 20% oxygen10 minGinseng seedsnatural fungal contamination from the surface of seeds~73% (Ar) and 60% (Ar/O2) inactivation of fungal spores[139]
DBSD plasmaAir0–60 sScot pine seedsFusarium oxysporum100% disinfection efficiency of seeds after 30 s treatment[140]
DCSBD plasmaAir0–10 minLentil seedsAspergillus niger
Penicillium decumbens
Maximum logarithmic reduction of 1.6 log CFU/g for A. niger and 3.1 log CFU/g for P. decumbens after 10 min[141]
DBD plasmaNitrogen,
Oxygen
1–3 minSoybean seedsDiaporthe/Phomopsis complexReduction in infection by about 49–81%[142]
DBD plasmaAir5, 20 minBarley and wheat seedsPenicillium verrucosumMaximal reduction of 2.1 log CFU/g for barley and 2.5 log CFU/g for wheat[67]
DCSBD plasmaAir0–50 sCucumber and pepper seedsCladosporium cucumerinum
Didymella bryoniae
Didymella licopersici
Total reduction in C. cucumerinum and 60–80% reduction in D. bryoniae on cucumber seeds after 20 s; 50–80% reduction in D. licopersici on pepper seeds[143]
DCSBD plasmaAir0–300 sMaizeAlternaria alternata
Aspergillus flavus
Fusarium culmorum
Reduction of 3.79 log CFU/g in F. culmorum after 60 s plasma treatment, 4.21 log CFU/g in A. flavus, and 3.22 log CFU/g in A. alternata after a 300 s plasma treatment[66]
DBSD plasmaAir0–300 sSweet basil seedsnatural fungal contamination~30% reduction of natural fungal contamination after 300 s[144]
Plasma jetHumid air10 minRice seedsFusarium fujikuroiBakanae disease severity index and the percentage of plants with symptoms were reduced to 18.1% and 7.8% of nonirradiated control, respectively, after 10 min treatment of seeds in water[145]
RF plane-type plasmaAir 0–30 minGroundnutsAspergillus flavus
Aspergillus parasiticus
High percentage of inactivation, 99.9% and 99.5% of A. parasiticus and A. flavus, respectively[146]
CD plasma jetAir0–3 minBroccoli seedsnatural fungal contamination1.5 log CFU/g reduction in natural fungal contamination after 3 min[147]
AP, atmospheric pressure; CD, corona discharge; DBD, dielectric barrier discharge; DBSD, dielectric barrier surface discharge; DCSBD, diffuse coplanar surface barrier discharge; LP, low pressure; PAW, plasma-activated water; RF, radiofrequency.
Table 3. Studies using NTPs for mutagenesis of fungi.
Table 3. Studies using NTPs for mutagenesis of fungi.
NTP TypeProcess GasTime of TreatmentFungus/YeastMutantRef.
ARTPHelium0–180 sFusidium coccineum~60%increase in fusidic acid production[163]
ARTPHelium0–350 sAspergillus nidulans1.3× higher production of echinocandin B[164]
ARTPHelium-Sanghuangporous sanghuang1.2–1.5 fold increase in polysaccharides production[165]
ARTP + etylmethanesulfonateHelium0–550 sPenicillium oxalicumEnhanced raw starch-degrading enzyme production[166]
ARTPHelium30–240 sAspergillus oryzae54.7% increase in acid protease activity, 17.3% increase in neutral protease activity, 8.5% increase in total protease activity, 8.1% decrease in alkaline protease activity[167]
ARTPHelium0–360 sStarmerella bombicola30% increase in lactonic, acidic, or total sophorolipid production[168]
ARTPHelium0–200 sCandida parapsilosis~60% increase in D-arabitol production[169]
ARTPHelium0–150 sCandida tropicalis22% increase in xylitol production[170]
ARTPHelium100–200 sAspergillus oryzae~292% increase in kojic acid production[171]
ARTPHelium30 s Hericium erinaceum22% increase in yield of fruiting body, 16% increase in polysaccharide production[172]
DBD plasmaArgon helium3–5 minGanoderma lingzhi25.6% increase in polysaccharides production[173]
ARTPHelium Trichoderma reeseiIncrease in cellulase production[174]
ARTP, atmospheric and room temperature plasma; DBD, dielectric barrier discharge.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hoppanová, L.; Kryštofová, S. Nonthermal Plasma Effects on Fungi: Applications, Fungal Responses, and Future Perspectives. Int. J. Mol. Sci. 2022, 23, 11592. https://doi.org/10.3390/ijms231911592

AMA Style

Hoppanová L, Kryštofová S. Nonthermal Plasma Effects on Fungi: Applications, Fungal Responses, and Future Perspectives. International Journal of Molecular Sciences. 2022; 23(19):11592. https://doi.org/10.3390/ijms231911592

Chicago/Turabian Style

Hoppanová, Lucia, and Svetlana Kryštofová. 2022. "Nonthermal Plasma Effects on Fungi: Applications, Fungal Responses, and Future Perspectives" International Journal of Molecular Sciences 23, no. 19: 11592. https://doi.org/10.3390/ijms231911592

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

Article Metrics

Back to TopTop