Next Article in Journal
Assessment of Mycotoxin Exposure and Associated Risk in Pregnant Dutch Women: The Human Biomonitoring Approach
Previous Article in Journal
Investigating Snake-Venom-Induced Dermonecrosis and Inflammation Using an Ex Vivo Human Skin Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Biocontrol of Occurrence Ochratoxin A in Wine: A Review

1
Department of Ecology, Agronomy and Aquaculture, University of Zadar, Trg kneza Viseslava 9, 23000 Zadar, Croatia
2
Faculty of Food Technology and Biotechnology, University of Zagreb, Pierottijeva 6, 10000 Zagreb, Croatia
3
Department of Environmental Biology, Sapienza University of Rome, Piazzale Aldo Moro 5, 00185 Rome, Italy
*
Authors to whom correspondence should be addressed.
Submission received: 6 May 2024 / Revised: 29 May 2024 / Accepted: 1 June 2024 / Published: 17 June 2024
(This article belongs to the Section Mycotoxins)

Abstract

:
Viticulture has been an important economic sector for centuries. In recent decades, global wine production has fluctuated between 250 and almost 300 million hectoliters, and in 2022, the value of wine exports reached EUR 37.6 billion. Climate change and the associated higher temperatures could favor the occurrence of ochratoxin A (OTA) in wine. OTA is a mycotoxin produced by some species of the genera Aspergillus and Penicillium and has nephrotoxic, immunotoxic, teratogenic, hepatotoxic, and carcinogenic effects on animals and humans. The presence of this toxin in wine is related to the type of wine—red wines are more frequently contaminated with OTA—and the geographical location of the vineyard. In Europe, the lower the latitude, the greater the risk of OTA contamination in wine. However, climate change could increase the risk of OTA contamination in wine in other regions. Due to their toxic effects, the development of effective and environmentally friendly methods to prevent, decontaminate, and degrade OTA is essential. This review summarises the available research on biological aspects of OTA prevention, removal, and degradation.
Key Contribution: Climate change is paving the way for pathogens to challenge crops and, specifically, grapevine more and more successfully. Regarding this, mycotoxin-producing fungi gain special attention. This review highlights the state-of-the-art on OTA occurrence in wine and the strategy based on biocontrol/bioremediating agents which actually are ready-to-use and cost-effective to preserve wine quality and safety.

1. Introduction

Archaeological finds indicate that viticulture began in the Neolithic period, around 7000 BC. Some authors believe that the first wine was produced in China, while others believe that this fermented beverage is related to rice wine rather than Vitis vinifera wine. In the period of 7000–4000 BC, intentional fermentation of Vitis vinifera grapes is documented in archaeological sites in Georgia, Azerbaijan, Armenia, Iran, and Turkey [1]. The transition from the Caucasian region to Western Europe took a very long time. In ancient Greek and Roman culture, wine was a prized beverage used for various purposes, including medicine. The importance of wine is attested to by the existence of gods associated with wine in various cultures: Osiris in Egypt, Dionysius in Greece, and Bacchus in Rome [2]. Ancient wine was different from the wine we know today. The technology of winemaking has evolved over the centuries to today’s production [3]. According to the International Organisation of Vine and Wine (OIV), an intergovernmental organization, global wine production in recent decades has exceeded 250 million hectoliters (mhl) with peaks near 300 mhl. The main producers are European countries, with Italy, France, and Spain accounting for around 50% of global wine production. However, wine production is also important in other regions of the world. The top ten wine producers include the USA, Australia, Argentina, Chile, and South Africa, while Chinese wine production has declined in recent years and is no longer in the top 10. In 2022, the global wine export value amounted to EUR 37.6 billion, and the largest wine-producing countries are also among the largest exporters [4]. Fluctuations in annual wine production are, as with all crops, dependent on changing climatic conditions. Climatic changes could have a major impact on wine production in the near future, and some traditional wine-growing regions could become less suitable for wine production. An example of this is South Africa, which, according to the OIV report, recorded a slight decline in wine-growing areas for eight consecutive years in 2022 as a result of the droughts between 2015 and 2017 [4]. Climate changes with a rise in temperature and possible periods of drought followed by heavy rainfall could have an impact on wine production but could also increase the risk of contamination of wine with the mycotoxin ochratoxin A. Generally, multiple preliminary models depict a scenario where numerous areas will experience increased contamination with mycotoxin-producing fungi across various crops. Examples include the increased risk of mycotoxin contamination in wheat (trichothecenes) and maize (aflatoxins) as a result of climate change [5,6]. According to a recent report, increased carbon dioxide concentrations in the Mediterranean region may increase the risk of OTA contamination in wine [7]. Furthermore, under warmer climate scenarios, Aspergillus niger populations may prevail over A. carbonarius as they are well adapted to higher temperatures and drier conditions. OTA production may also be increased under a climate change scenario for A. westerdijkiae, which is thermophilic, at least compared to other “black” Aspergillia [8].
Ochratoxin A (OTA) is a mycotoxin produced by some species of the genera Penicillium and Aspergillus. P. verrucosum is considered among the most important species for contamination of food and feed in temperate climates, while A. ochraceus, A. carbonarius, and A. niger are more important in areas with higher temperatures [9]. In addition, A. carbonarius is the main cause of contamination of wine [10]. OTA was first isolated from A. ochraceus in 1965 and named after the producing species [11]. OTA has various toxic effects on humans and animals, with the kidneys being the main target. It has been reported to have nephrotoxic, immunotoxic, teratogenic, hepatotoxic, and carcinogenic effects on animals [12,13]. A carcinogenic effect has been demonstrated in animals but not in humans, which is why the International Agency for Research on Cancer has classified OTA in Group 2B, a possible human carcinogen [14]. Some authors [15,16] report that OTA is a probable cause of nephropathies and urothelial tumors in humans, while other authors have linked OTA to the endemic Balkan nephropathy, BEN [12,17]. In their review of epidemiologic studies on OTA exposure and adverse health effects in humans, ref. [9] concluded that while nephrotoxicity in animals is well documented, the results in humans are inconclusive. BEN is now associated with aristolochic acid and not with OTA [18]. Chemically, OTA is a stable molecule that can persist in food and feed due to its resistance to common food processing treatments such as heating or fermentation [19,20]. In addition, OTA is persistent in humans and has a rather long half-life in blood; also, it can be present in animal tissue and human breast milk [21,22,23,24]. Due to its toxicity, the European Commission has limited the concentration of OTA in food; the limit for wine is 2 ng mg−1 [25]. This regulation is no longer in force; it was replaced by EC Regulation 2023/915 on 25 April 2023, but the limit remained the same [26].
Contamination with OTA has been reported in corn, barley, wheat, nuts, dried fruits, smoked and dried fish, coffee grains, grapes, wine, beer, dried meat products, and cheese [9,27]. Cereal grains are the main source of daily intake of OTA by humans [28]. The contribution of wine to the daily intake of OTA depends on the dietary habits of the consumer. According to [29], wine could contribute significantly to daily OTA intake in some populations, up to 15%, ref. [30] reported that at the 31st meeting of the Codex Alimentarius Commission’s Committee on Food Additives and Contaminants, wine was cited as the second most important source of daily OTA intake after cereals. In more recent publications, various values are given for the same country in different years [10]. Although the exact contribution of wine consumption to the daily intake of OTA cannot be exactly calculated, it certainly contributes in some population groups. The World Health Organization recommends reducing the presence of mycotoxins in food and feed as much as possible [31], and this also applies to wine.
Climate change will affect all areas of agriculture, including wine production [32]. The rise in atmospheric temperature, erratic rainfall patterns, and the increase in extreme climatic events such as drought, storms, and heavy rainfall could affect not only productivity but also food safety. Environmental conditions could favor the spread of pathogens, including mycotoxigenic ones, to areas where they were not previously present and/or increase their occurrence in areas where they are already present [33]. With regard specifically to the risk of contamination of wine with OTA, climate change could strongly favor the occurrence of OTA. Temperature is one of the most important factors influencing the contamination of wine with OTA, as discussed previously and in the following chapter. The increase in temperature could lead to both a greater frequency of wine contamination and a higher concentration of OTA in wine in areas already considered at risk of OTA contamination, as well as creating a more favorable environment for wine contamination in areas previously considered at low(er) risk. Heavy rainfall, hail, and storms, especially in the late veraison phase, could damage the berries and favor fungal invasion and OTA production. These scenarios indicate that even more attention must be paid to plant protection in the future. The control of plant pathogens relies on the use of chemicals, but their use is constantly being restricted by legislation worldwide due to their harmful effects on the environment and human and animal health. Many of the chemicals used in plant protection have been banned in recent decades, particularly in EU countries [34]. Biocontrol agents are seen as a viable and more environmentally friendly alternative to conventional pesticides [35,36,37,38]. This paper provides an overview of the methods used to decontaminate and control OTA occurrence in wine through low environmental impact methods.

2. OTA in Wine

The presence of OTA in wine was first reported in 1996 [38]. In the following 10–15 years, various papers were published reporting the presence of OTA in wine in different countries around the world. Table 1 lists some of the publications investigating the presence of OTA in wine in different countries. As Europe is the largest producer of wine, it should come as no surprise that most of the research has been conducted in European countries. Even though these publications found the presence of OTA in wines from all over the world, in most cases, the concentration was (far) below the EU limits (Table 1). The highest OTA concentration in wine, 9.2 µg L−1, was reported by [39] for a wine from southern Italy. The overall results show that the presence of OTA is influenced by the type of wine and its geographical origin and, obviously, by seasonality. In general, red wine is more sensitive to OTA contamination, followed by rose wine and white wine [5,12,38,40,41]. For example, ref. [12] showed in their analysis of 420 wine samples of worldwide origin that 25% of the white wines examined had a detectable OTA concentration, 40% of the rose wines, and 54% of the red wines. Furthermore, the highest concentration was found in red wine (3.31 ng L−1), followed by rosé (2.38 ng L−1) and white wine (1.36 ng L−1). This is in line with the results of other authors [30,42,43,44,45].
The occurrence of OTA in wine is related to the geographical location of the vineyards. Battilani et al. [30] reported that in Europe, the amount of OTA depends on the latitude of production; the lower the latitude, the higher and more frequent the occurrence of OTA in wine. Other authors also reported a higher occurrence of OTA in wines from southern regions [12,42,43,44,46,47,48]. According to these authors, the regions with the highest risk of OTA contamination in wine are southern Italy, south-eastern Spain, and parts of Greece and Turkey (Figure 1). For the rest of the world, such a correlation between latitude and OTA occurrence is not known, but it is generally assumed that wines from regions with higher temperatures, such as the Mediterranean, are more frequently contaminated with OTA than wines from regions with more temperate climatic conditions [10].
Other grape products, such as grape juices and dried grapes, have also been found to be contaminated with OTA [5,12,30,47,49,50].
Table 1. Presence of OTA in different types of wine in different parts of the world.
Table 1. Presence of OTA in different types of wine in different parts of the world.
CountryType of WineOTA Detected/Total SamplesSamples over the EU Limit
(2 ng mL−1)
Source
ArgentinaRed4/473[51]
Red136/1360[52]
Brazil Red18/260[53]
White7/170
ChileRed28/8690[54]
White6/3190
ChinaRed183/1830[55]
White40/400
CroatiaRed7/70[42]
White4/70
Red6/60[56]
White8/100
Francered grape most11/370[57]
GreeceRed58/780[49]
White40/620
Sweet6/100
Red33/1041[58]
White55/1180
Sweet7/100
ItalyRed14/96yes *[46]
Red553/73522
White128/2900[43]
Rose69/754
Desert18/280
Red29/200[48]
Sweet55/550[59]
PolandRed49/53yes *[60]
PortugalRed9/351[61]
White3/250
South AfricaRed15/150[62]
White9/90
SpainRed34/1402[50]
White6/601
Sweet186/18818[40]
USARed28/312[63]
White7/100
* The authors do not state the number of samples in which the OTA concentration is above the EU limits, but only the range of the minimum and maximum concentrations found.

3. Factors That Influence OTA Presence in Wines

OTA is synthesized on grape berries and contaminates the wine during winemaking. The crucial period for OTA contamination is between early veraison and harvest [30,64]. Conidia of ochratoxigenic fungi may be present on the berries, but their ability to penetrate the berries is low unless the berry is damaged, and wounds caused by biotic or abiotic factors favor their penetration [65]. One of the causes of berry bursting is excessive watering, either by irrigation or by rainfall, before harvest [66]. Insects, especially Lobesia botrana, the grape berry moth, can contribute significantly to OTA contamination by damaging the barriers [65,67,68]. As already reported in the previous chapter, climatic conditions, in particular air temperature and humidity, are the most important factors for OTA contamination of wines. Temperature and humidity influence the growth of all molds. The optimal temperature and aw for the growth of A. carbonarius is 25–30 °C and an aw value of 0.930 to 0.987, depending on the strain. However, the optimal temperature for OTA production within the same strains was different, 15–20 °C [69]. Other authors reported that the optimal temperature and water activity for A. carbonarius growth and optimal OTA biosynthesis may be different [70]. The grape variety and its intrinsic genetic characteristics (i.e., a compact grape shape is more exposed to OTA contamination than a loose one), as well as the state of ripeness, can also influence the occurrence of OTA in wine [44,45,71]. A positive correlation between the concentration of soluble solids, pectin content, total anthocyanins, and OTA content is reported. In laboratory tests using berries from three different varieties at different stages of ripeness, A. carbonarius inoculated onto the Muscat Italia variety showed greater OTA production in the early ripening stage when pH and sugar content were lower [45]. Passamani et al. [70] also found that lower pH contributes to OTA biosynthesis.
The role of oxidative stress in the regulation of OTA synthesis has been demonstrated [72]. In another experiment, prooxidants, but also some antioxidants, enhanced the OTA production of A. carbonarius [73]. This is confirmed by [45]. They observed that some antioxidants showed a positive correlation with OTA production, while for some others, the correlation was negative, suggesting that a structure-dependent signal is involved in the inhibition.
Another factor that can influence OTA production is carbon dioxide. Experiments with grape-based growing media have shown that higher CO2 concentrations stimulate OTA production. The higher CO2 concentration (1000 ppm) shortened both the lag phase of fungal growth and increased OTA production. In some fungal strains, the combined effect of higher temperature and higher CO2 concentration was a significantly shorter lag phase [7].
Finally, the winemaking process also influences the OTA contamination of the wine. Previously, a greater risk of OTA contamination was discussed for red wines. This fact can be explained by the different winemaking processes for white and red wines. White wines are pressed immediately after the harvest, whereas the grapes for red wines are first crushed and must be macerated for several days. During this time, OTA synthesis and its transition from the grapes to the must take place [12,74].

4. In Field Prevention of OTA Contamination in Wine

The problem of contamination of grapes with OTA begins in the vineyard, so the first preventive measures should be carried out there [10]. Certainly, good agricultural practice is necessary and could be helpful in prevention. The practices used to reduce fungal diseases in the vineyard, such as an optimal ratio of leaves to bunches, good ventilation of the bunches, or an optimal distance between bunches, could help prevent the appearance of OTA in wine [75]. As already mentioned, the critical period is between the beginning of the ripening period and the harvest. Some authors suggest that the critical period should be extended from early veraison to the beginning of vinification [5,10,44,45,74]. It is very important not to irrigate the vineyard too much to avoid bursting of the berries, which is especially important in the later stages of veraison. In addition, the harvest should be well planned, especially in regions where the risk of OTA contamination is higher. The time between harvest and vinification should be as short as possible [5,12,30,41,71,74,75].
The control of insects, especially Lobesia botrana, can contribute significantly to reducing the risk of contamination [71,76]. In a trial conducted in Sicily in 2006, the amount of OTA recovered from grapes infected with L. botrana was significantly higher than in uninfected grapes: 20 µg L−1 in infected grapes and 0.04 µg L−1 in uninfected grapes [65,76]. In another trial in which biological control agents of Lobesia botrana were used in the field for two years, there was an 80% reduction in the OTA content of the grapes in the year in which the meteorological conditions were favorable for insect infestation [68].

5. Biocontrol of OTA Occurrence in Wine

As an alternative to chemical and physical OTA detoxification methods, there is now a growing interest in the use of biological methods, i.e., microorganisms involved in food fermentation, as biological agents for OTA control and detoxification. These biological methods are based on non-toxic bacteria, yeasts, and even fungi and other microorganisms as main components. Biological methods have many advantages, such as low cost and minimal side effects, and with the increasing awareness of environmental protection and food safety, they are the best representatives of the methods with lower environmental impact. They are widely used to reduce the concentration of OTA in food, grain, and feed [16,77]. Biological methods can be categorized into three groups based on the mechanisms of action: (1) inhibition of fungi responsible for OTA biosynthesis, (2) adsorption of OTA, and (3) degradation or detoxification of OTA in contaminated matrices [78]. Microbiological detoxification methods are defined as methods using microorganisms and/or enzymes that can metabolize, destroy, or deactivate toxins to stable, less toxic, or harmless compounds [79]. An example of microbial OTA detoxification metabolites is shown in Figure 2.
Biological agents and their associated enzymes allow a specific, probably irreversible, environmentally friendly, and effective approach with minimal impact on the sensory and nutritional quality of food compared to chemical methods used to reduce mycotoxin concentration [16,80,81]. Yeasts, especially Saccharomyces species and lactic acid bacteria (LAB), are mainly used as part of the natural microbial population in spontaneous food fermentation and as starter cultures in the food and beverage industry [74,82,83,84]. The microbial degradation of OTA could be mediated by the hydrolysis of the amide bond to the non-toxic compounds L-β-phenylalanine and OTα [84,85,86] or by hydrolysis of the lactone ring [85,86,87,88,89], while a proposed model for mycotoxin binding consists of two processes: binding (adsorption) and release (desorption) to/from the binding sites on the surface of the microorganism [89,90,91,92,93]. It has been shown that both viable and dead cells of microorganisms bind mycotoxins. Carbohydrate and/or protein components of the cell wall play an important role in binding mycotoxins; the binding efficiency depends on the strain of the microorganism, the amount of mycotoxin, the environmental conditions (pH), and the stability of the microorganism–mycotoxin complex [74,84,94,95,96]. Many microorganisms, including bacteria, yeasts, filamentous fungi, and protozoa, can remove, reduce, or biodegrade OTA [82,84,88,89,91,93,97]. Among these potential microorganisms, oenological LAB and yeasts represent a unique group of biocontrol agents for the biodegradation, removal, or binding of OTA due to their role in winemaking (Figure 3) without affecting the organoleptic properties of wine [86,97,98,99].

5.1. Inhibition of the Fungi Responsible for OTA Biosynthesis

The production of OTA can be inhibited in different ways, depending on the type of microorganism. For example, many bacteria can produce various lipopeptides that inhibit OTA production by altering the cell structure or even destroying the cell structure of filamentous fungi, which leads to metabolic imbalance and inhibits spore germination but can also downregulate genes related to OTA synthesis [100,101]. Biocontrol can be used directly against ochratoxigenic molds. In peanut cultivation, contamination in the field was reduced when a non-aflatoxin-producing strain of Aspergillus flavus was inoculated [102]. In vitro experiments with OTA-producing A. carbonarius and two mutants, Dpsk and DveA, which cannot produce OTA, gave encouraging results [103]. Non-toxigenic strains were able to displace ochratoxigenic strains in liquid media, and the OTA concentration was reduced in all tested concentrations due to the presence of non-toxigenic strains. Even in the most unfavorable concentrations of the inoculum (10-fold higher concentration of the toxigenic strain), the OTA concentration was reduced by almost 30%. Similar results were obtained in grape berries, suggesting that the use of competitive, non-toxigenic A. carbonarius strains in vineyards could be an effective means of OTA control. In laboratory experiments, some yeast strains of Saccharomyces cerevisiae and Kloeckera apiculate showed inhibition of A. carbonarius growth and OTA synthesis on both liquid media and grape berries [104]. Yeast strains of other species, Candida intermedia, and Lachancea thermotolerance inhibited A. carbonarius growth and OTA production, both in vitro and on berries [105]. Although there is evidence that some yeasts or fungi could control the growth of A. carbonarius on grape berries, there is no large-scale product on the market based on biocontrol agents. Yeasts are being studied in more detail as biocontrol agents for the removal of OTA in winemaking and are discussed below.

5.2. Effects of OTA on Bacteria and Yeasts

5.2.1. Effect of OTA on Bacteria

The effect of OTA on LAB is also important, but the few published studies suggest that bacteria, including LAB, are microorganisms that are resistant to OTA. Little is still known about the effects of mycotoxins on LAB, as the toxicity thresholds reported by different authors vary. Therefore, Piotrowska and Zakowska [106] conducted a study in which the sensitivity of LAB to the effects of different OTA concentrations was investigated. In the range of 0.1 to 10 µg, only one LAB was sensitive to OTA, while other LAB strains, with few exceptions (Lactobacillus acidophilus Ind1 and L. acidophilus H-1), showed no sensitivity to OTA at concentrations of 0.1 to 5 µg. Their growth was inhibited by OTA at 5 µg. At the same time, high concentrations of OTA (10 µg) proved to be growth inhibitory for most strains. Only a few strains remained unaffected by the OTA concentrations (L. acidophilus B, L. acidophilus CH-5, and L. rhamnosus GG). The results obtained show that OTA has a negative effect on the growth of LAB, but these results differ from the results of [107], who showed that OTA at a concentration of 20 ppm does not inhibit the growth of L. plantarum and L. casei. According to [108], two concentrations of OTA (2 and 4 µg/mL) had minimal effect on the growth of L. plantarum B, with visible activity only during the lag phase, which was extended by two hours.

5.2.2. Effects of OTA on Yeasts

Much is known about the use of yeasts for biocontrol, degradation, removal, or binding of OTA [56,84,98,109]. However, little is known about the effects of mycotoxins as stress factors on yeasts, their morphology, their growth parameters, their metabolic activity, and their effects on fermentation [98]. Therefore, it is possible that OTA can inhibit the enzymes of ethanol fermentation, impair cell growth, and induce oxidative stress. The presence of OTA had an impact on the viability of the Saccharomyces strains studied. In particular, lower concentrations of OTA (2 µg/mL) had an inhibitory effect on the cell growth of both S. cerevisiae and S. uvarum. Although the viability of S. cerevisiae was not significantly affected, the cell diameter changed considerably. Since OTA had a major effect on the cell diameter and viability of S. uvarum, it is clear that the effect of OTA is strain-specific [110]. In the study by Jakopović et al. [98], ethanol synthesis by the yeast S. uvarum was slowed by 4 µg/mL OTA. Freire et al. [111] investigated the effects of 10, 20, and 30 µg/L OTA on three different yeast strains of S. cerevisiae and obtained similar results. The presence of ochratoxin A during fermentation did not affect the growth of the S. cerevisiae strains, but some differences were observed. For example, one of the S. cerevisiae strains tested showed a lower growth rate and a longer lag phase in the presence of OTA than the other two strains tested. In addition, the same S. cerevisiae strain had a longer exponential and stationary phase. Despite all this, the maximum population of the yeasts studied was not affected by the presence of different concentrations of OTA in the fermentation medium.

5.3. Adsorption of OTA

Ample evidence of the beneficial effect of LABs and their additional probiotic properties gives them significant potential for their application in food and feed, and they are also good candidates as mycotoxin-biotransforming bacteria [84,89,112,113,114]. The ability to reduce the amount of OTA is common in LAB but varies by species and bacterial strain.
After studying the effects of OTA on LAB, Piotrowska and Zakowska [106] investigated the phenomenon of OTA elimination by LAB. The study of the average values of LAB species showed that Lactobacillus acidophilus, L. rhamnosus, L. sanfranciscens, and L. plantarum have the greatest ability to remove OTA. However, analysis of the values obtained for specific strains within species revealed significant differences, again confirming that the ability to bind OTA is strain-specific. Del Prete et al. [115] analyzed 15 strains belonging to five relevant oenological LAB species for their sensitivity to ochratoxin A and their ability to remove this toxin from liquid media. The amount of OTA removed during bacterial growth was found to be between 8 and 28%, with Oenococcus oeni being the most effective strain, reducing OTA by 28%. OTA was not degraded by cell-free extracts of oenological LAB strains, and no degradation products were detected, suggesting that the removal of OTA by oenological LAB is a binding process. Mateo et al. have shown that several strains of O. oeni reduce OTA levels by 50–70%, while Lactobacillus acidophilus reduces OTA in the broth medium by ≥98%, while heat-treated LAB cells remove small amounts of OTA (less than 11%) [91]. In contrast, a study by Piotrowska [116] showed that heat-inactivated Lactobacillus cells reduced OTA (46.2–59.8%) more efficiently than live cells (16.9–35%). Shukla et al. [117] observed that the amount of OTA in red wine samples was reduced by 90% when treated with heat-treated B. subtilis cells. The results obtained by several researchers have shown that cell wall components such as peptidoglycan and polysaccharides play an important role in the binding of OTA. Due to changes in the cell surface, mycotoxin binding is permanent when the LAB cells are dead (heat or acid-treated), while the living bacteria can release part of the mycotoxin content over time [84,92,106,114,118,119]. The higher OTA adsorption by dead cells compared to viable cells can be explained by changes in the bacterial cell wall or by the hydrophobic nature of the bacterial cell wall. Thermal or acidic treatment of the cells leads to denaturation of the proteins and increased permeability of the outer layers of the cell wall. As a result, a larger number of active sites are formed, which are responsible for the absorption of various compounds [116]. Although the results of several researchers have shown that LAB cells can adsorb OTA and it is removed or reduced from the medium in the absence of metabolically active cells, the results of other researchers have shown that metabolic degradation by hydrolysis of the amide bond or by hydrolysis of the lactone ring may be the mechanism responsible for OTA removal [77,84,88,89,91,114].
Table 2 summarizes the microorganisms able to adsorb more than 75% of the OTA present in the media.

5.4. Degradation of OTA

In their study, Rodriguez et al. [120] investigated the ability to degrade OTA using Pseudomonas putida and various bacteria of the genera Rhodococcus and Brevibacterium. After the growth of the bacteria in a liquid synthetic medium and the addition of OTA (~10 µg/L), a small decrease in OTA concentration (8–28%) was observed in the supernatants of Rhodococcus and P. putida strains with no evidence of degradation products [120]. This suggests that OTA is not degraded but adsorbed by the cells, as previously described [114,115,121,122,123]. On the other hand, analysis of the supernatant of B. casei RM101 showed the complete absence of OTA [123]. Since Brevibacterium species differ from other bacteria in their ability to metabolize compounds with heterocyclic and polycyclic ring structures (a property also exhibited by fungi), further analyses were performed [120]. Brevibacterium strains can completely degrade OTA, even at concentrations of up to 40 mg/L, a concentration 1000 times greater than that commonly found in foodstuffs. Therefore, it is clear that OTA degradation is characteristic of the genus Brevibacterium. Abrunhosa and coworkers [89] tested the ability of 19 different LAB strains for wine to degrade OTA. Pediococcus parvulus UTAD 473 showed the most promising results: it converted 90% of OTA into less toxic OTα. Rodrigues et al. [124] reported that OTα is non-toxic or at least 500 times less toxic than OTA.
Table 3 lists the microorganisms for which OTA removal by degradation of more than 75% was reported.

Enzymatic Degradation of OTA

Enzymes are mainly derived from the production of specific microbial strains and are an important component of the biodegradation mechanisms of OTA. The enzymes produced can be both extracellular and intracellular [125,126,127], and the most common types of OTA biodegradation enzymes are mainly carboxypeptidase and amidase [16]. The review by [81] focuses on the biotransformation of mycotoxins with purified enzymes isolated from bacteria, fungi, and plants, whose activity has been validated by in vitro and in vivo tests. According to scientific studies, crude and purified enzymes are involved in the transformation of OTA. These enzymes include several classes of carboxypeptidases: carboxypeptidase A (CPA), carboxypeptidase B (CPB), and carboxypeptidase Y (CPY); metalloenzymes derived from atoxigenic Aspergillus niger strains, as well as lipase A, protease A, ochratoxinase, and amidases that hydrolyze the amide bond of OTA to generate OTα and phenylalanine [86,128,129,130,131,132,133,134,135,136,137]. The first report on the biodetoxification of OTA dates back to 1969 [138] and used bovine pancreatic carboxypeptidase A (CPA) to cleave OTA into OTα. Dobritzsch et al. [132] reported that purified recombinant ochratoxinase was about 600 times more effective than CPA, with optimal OTA degradation activity at 66 °C and pH 6. In the studies using CPY isolated from S. cerevisiae, it was shown that only 52% of OTA was converted to OTα at 37 °C and pH 5.6 [86,129] have shown that the use of commercial hydrolases, such as lipase preparations from A. niger, is also capable of hydrolyzing OTA to OTα and phenylalanine. Abrunhosa et al. [130] reported that Ancex, a crude enzyme isolated from A. niger, was able to degrade 99.8% of OTA (1 µg/mL) to OTα after 25 h of incubation at pH 7.5 and 37 °C. Commercially purified enzymes such as protease A and pancreatin were not able to degrade as much OTA (87.3 and 43.4%, respectively). Another example is Prolyve PAC, which was only able to degrade 3% of OTA (1 µg/mL) to OTα after 25 h of incubation at pH 3 and 37 °C [124]. While Bejaoui et al. (2006) [139] reported that OTα was further degraded to unknown products, several authors reported that OTA could be directly degraded to unknown products [99,140,141]. Since not every enzymatic reaction leads to real detoxification, because the metabolized mycotoxin can take on greater toxic properties than the parent compound, it is necessary to identify and characterize the degrading enzymes and understand the mechanism of degradation and test the toxicity of OTA by-products [16,137,142].
In Table 4 are reported enzymes produced by microorganisms able to degrade the OTA.
The OTA-degrading enzymes produced by microorganisms are listed in Table 4.

6. Conclusions

Climate change poses various risks for plant cultivation; for example, it can promote the occurrence and concentration of OTA in wine. In order to avoid or limit this, prevention strategies that are sustainable for the environment must be developed. Increasing fungal resistance and the challenges associated with conventional systems require the development of strategies that allow both efficient biocontrol of ochratoxigenic strains and rapid elimination of OTA from the wine with short processing times and negligible impact on quality.
This review shows that biological control is an important, environmentally friendly, and efficient tool for controlling the presence of OTA in wine. Although some yeasts and atoxigenic strains of A. carbonarius have been shown to be effective in controlling the growth of ochratoxigenic fungi and are able to inhibit OTA synthesis, no biological control agent is yet widely and commercially used in vineyards for OTA prevention. Regardless of the effectiveness of preventive measures, human exposure to this mycotoxin can also be minimized by decontamination and detoxification strategies once OTA is detected in wine. Various biological decontamination and detoxification methods have been proposed to limit the contamination of wine with OTA. In particular, the use of different microorganisms, such as LAB and S. cerevisiae, and microbial enzymes has shown that over 90% of OTA can be efficiently removed from wine. The effects of OTA degradation on the organoleptic properties and taste of wine should be investigated before the large-scale application of this method.
Further research is needed on microorganisms that can prevent the occurrence of OTA or safely remove it from the wine without affecting the organoleptic properties of the wine. In particular, research into biocontrol agents that can control the presence and growth of ochratoxigenic strains in vineyards is desirable.

Author Contributions

S.Z. and K.M. Conceptualization, writing original draft, writing-reviewing and editing; J.L., Z.J. and M.B. writing-reviewing and editing; M.R. Conceptualization, supervision, writing-reviewing and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable, no new data is generated.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Estereicher, S.K. The beginning of wine and viticulture. Phys. Status Solidi C 2017, 14, 1700008. [Google Scholar] [CrossRef]
  2. Rosso, A.M. Beer and wine in antiquity: Beneficial remedy or punishment imposed by the Gods? Acta Med.-Hist. Adriat. 2012, 10, 237–262. [Google Scholar] [PubMed]
  3. Chambers, P.J.; Pretorius, I.S. Fermenting knowledge: The history of wine making, science and yeast research. EMBO Rep. 2010, 11, 914–920. [Google Scholar] [CrossRef] [PubMed]
  4. OVI, International Organisation of Vine and Wine 2023. State of the World Vine and Wine Sector in 2022. Available online: https://www.oiv.int/sites/default/files/documents/2023_SWVWS_report_EN.pdf (accessed on 10 December 2023).
  5. Battilani, P.; Giorni, P.; Bertuzzi, T.; Formenti, S.; Pietri, A. Black Aspergilli and ochratoxin A in grapes in Italy. Int. J. Food Microbiol. 2006, 111, S53–S60. [Google Scholar] [CrossRef] [PubMed]
  6. Infantino, A.; Belocchi, A.; Quaranta, F.; Reverberi, M.; Beccaccioli, M.; Lombardi, D.; Vitale, M. Effects of climate change on the distribution of Fusarium spp. in Italy. Sci. Total Environ. 2023, 882, 163640. [Google Scholar] [CrossRef] [PubMed]
  7. Cervini, C.; Vercheecke Vassen, C.; Ferrara, M.G.; Garcia-Cela, E.; Magista, D.; Medina, A.; Gallo, A.; Perrone, G. Interracting climate change factors (CO2 and temperature cycles) on growth, secondary metabolites gene expression and phenotypic ochratoxin A by Aspergillus carbonarius strains on grape based matrix. Fungal Biol. 2021, 125, 115–122. [Google Scholar] [CrossRef] [PubMed]
  8. Akbar, A.; Medina, A.; Magan, N. Impact of interacting climate change factors on growth and ochratoxin A production by Aspergillus section Circumdati and Nigri species on coffee. World Mycotoxin J. 2016, 9, 863–874. [Google Scholar] [CrossRef]
  9. Bui-Klinke, T.R.; Wu, F. Ochratoxin A and human health risk: A review of the evidence. Crit. Rev. Food Sci. Nutr. 2015, 55, 1860–1869. [Google Scholar] [CrossRef] [PubMed]
  10. La Placa, L.; Tsitsigiannis, D.; Legieri, M.C.; Battilani, P. From grapes to wine: Impact of the vinification process on ochratoxin A contamination. Foods 2023, 12, 260. [Google Scholar] [CrossRef]
  11. van der Merwe, K.J.; Steyn, P.S.; Fourie, L.; Scott, D.B.; Theron, J.J. Ochratoxin A, a toxic metabolite produced by Aspergillus Ochraceus. Will. Nat. 1965, 205, 1112–1113. [Google Scholar] [CrossRef]
  12. Otteneder, H.; Majerus, P. Occurrence of ochratoxin A (OTA) in wines: Influence of the type of wine and its geographical origin. Food Addit. Contam. 2000, 17, 793–798. [Google Scholar] [CrossRef] [PubMed]
  13. Amezqueta, S.; Shorr-Galindo, S.; Murillo-Arbizu, M.; Gonzales-Pena, E.; Lopez de Certain, A.; Giraud, J.P. OTA producing fungi in foodstuff: A review. Food Control 2012, 26, 259–268. [Google Scholar] [CrossRef]
  14. IARC. Monographs on the Evaluation of Carcinogenic Risks to Humans: Some Naturally Occurring Substances: Food Items and Con-Stituents, Heterocyclic Aromatic Amines and Mycotoxins; IARC: Lyon, France, 1993; pp. 1–599. [Google Scholar]
  15. O’Brien, E.; Dietrich, D.R. Ochratoxin A: The continuing enigma. Crit. Rev. Toxicol. 2005, 35, 33–60. [Google Scholar] [CrossRef] [PubMed]
  16. Ding, L.; Han, M.; Wang, X.; Guo, Y. Ochratoxin A: Overview of prevention, removal and detoxification methods. Toxins 2023, 15, 565. [Google Scholar] [CrossRef] [PubMed]
  17. Stoev, S.D. Balkan Endemic Nephropathy—Still continuing enigma, risk assessment, and understanding the hazard of joint mycotoxin exposure of animals and humans. Chem. Biol. Interact. 2017, 261, 63–79. [Google Scholar] [CrossRef] [PubMed]
  18. Jelakovic, B.; Dika, Z.; Arlt, V.M.; Striborova, M.; Pavlovic, N.M.; Nikolic, J.; Colet, J.M.; Vanherwegen, J.L.; Nortier, J.L. Balkanic Endemic Nephropathy and Causative Role of Aristolochic Acid. Semin. Nephrol. 2019, 39, 284–296. [Google Scholar] [CrossRef] [PubMed]
  19. Ponsone, M.L.; Chiotta, M.L.; Combina, M.; Dalcero, A.M.; Chulze, S.N. Fate of ochratoxin A in Argentinean red wine during pilot scale vinification. Rev. Argent. 2009, 41, 245–250. [Google Scholar]
  20. Abraham, N.; Chan, E.T.S.; Zhou, T.; Seah, S.Y.K. Microbial detoxification of mycotoxins in food. Front. Microbiol. 2022, 13, 957148. [Google Scholar] [CrossRef] [PubMed]
  21. Petzinger, E.; Zeigler, K. Ochratoxin A from toxicological perspective. J. Vet. Pharmacol. Ther. 2000, 23, 91–98. [Google Scholar] [CrossRef]
  22. Duarte, S.C.; Lino, C.M.; Pena, A. Ochratoxin A in feed of food producing animals: An undesirable mycotoxin with health a performance effects. Vet. Microbiol. 2011, 1548, 1–13. [Google Scholar] [CrossRef]
  23. Munoz, K.; Blaskewicz, M.; Campos, V.; Vega, M.; Degen, G.H. Exposure of infants to ochratoxin A with breast milk. Arch. Toxicol. 2014, 88, 837–846. [Google Scholar] [CrossRef] [PubMed]
  24. Memis, E.Y.; Yalcin, S.S. Human milk mycotoxin contamination: Smoking exposure and breastfeeding problems. J. Matern.-Fetal Neonatal Med. 2021, 34, 31–40. [Google Scholar] [CrossRef] [PubMed]
  25. European Commission. Commission Regulation (EC) Number 1881/2006 of 19 December 2006 Setting Maximum Levels of Certain. Contaminants in Foodstuff; European Commission: Ispra, Italy, 2006. [Google Scholar]
  26. European Commission. Directive 2009/128/EC Aims to Achieve a Sustainable Use of Pesticides in the EU by Reducing the Risks and Impacts of Pesticide Use; European Commission: Ispra, Italy, 2009. [Google Scholar]
  27. Bhat, R.; Rai, R.V.; Karim, A.A. Mycotoxins in food and feed: Present status and future concerns. Compr. Rev. Food Sci. Food Saf. 2010, 9, 57–81. [Google Scholar] [CrossRef] [PubMed]
  28. Kumar, P.; Mahato, D.K.; Sharma, B.; Borah, R.; Haque, S.; Mahmud, M.M.C.; Shah, A.K.; Rawal, D.; Bora, H.; Bui, S. Ochratoxins in food and feed: Occurrence and its impact on human health and management strategies. Toxicon 2020, 187, 151–162. [Google Scholar] [CrossRef] [PubMed]
  29. Leblanc, J.C.; Tard, A.; Volatier, J.L.; Verger, P. Estimated dietary exposure to principal food mycotoxins from the first French total diet study. Food Addit. Contam. 2005, 22, 652–672. [Google Scholar] [CrossRef] [PubMed]
  30. Battilani, P.; Magan, N.; Logrieco, A. European research on ochratoxin A in grapes and wine. Int. J. Food Microbiol. 2006, 111, S2–S4. [Google Scholar] [CrossRef] [PubMed]
  31. World Health Organization (WHO). Mycotoxins. 2023. Available online: https://www.who.int/news-room/fact-sheets/detail/mycotoxins (accessed on 30 November 2023).
  32. Barik, S.K.; Behera, M.D.; Shrotriya, S.; Likhovskoi, V. Monitoring climate change impacts on agriculture and forests: Trends and prospects. Environ. Monit. Assess. 2022, 195, 174. [Google Scholar] [CrossRef] [PubMed]
  33. Singh, B.K.; Delgado-Baquerizo, M.; Egidi, E.; Guirado, E.; Leach, J.E.; Liu, H.; Trivedi, P. Climate change impacts on plant pathogens, food security and paths forward. Nat. Rev. Microbiol. 2023, 21, 640–656. [Google Scholar] [CrossRef] [PubMed]
  34. Marchand, P.A. EU chemical plant protection products in 2023: Current state and perspectives. Agrochemicals 2023, 2, 106–117. [Google Scholar] [CrossRef]
  35. Vekemans, M.C.; Marchand, P.A. The fate of biocontrol agents under the European phytopharmaceutical regulation: How this regulation hinders the approval of botanicals as new active substances. Environ. Sci. Pollut. Res. 2020, 27, 39879–39887. [Google Scholar] [CrossRef]
  36. Loncar, J.; Bellich, B.; Parroni, A.; Reverberi, M.; Rizzo, R.; Zjalić, S.; Cescutti, P. Oligosaccharides Derived from Tramesan: Their Structure and Activity on Mycotoxin Inhibition in Aspergillus flavus and Aspergillus carbonarius. Biomolecules 2021, 11, 243. [Google Scholar] [CrossRef] [PubMed]
  37. Thomas, G.; Rusman, Q.; Morrison, W.R., 3rd; Magalhães, D.M.; Dowell, J.A.; Ngumbi, E.; Osei-Owusu, J.; Kansman, J.; Gaffke, A.; Pagadala Damodaram, K.J.; et al. Deciphering Plant-Insect-Microorganism Signals for Sustainable Crop Production. Biomolecules 2023, 13, 997. [Google Scholar] [CrossRef] [PubMed]
  38. Zimmerli, B.; Dick, R. Ochratoxin A in grape wine and grape-juice: Occurrence and risk assessment. Food Addit. Contam. 1996, 13, 665–668. [Google Scholar] [CrossRef] [PubMed]
  39. Lucchetta, G.; Bazzo, I.; Cortivo, G.D.; Stringher, L.; Bellotto, D.; Borgo, M.; Angelini, E. Occurrence of black Aspergilli and ochratoxin A on grapes in Italy. Toxins 2010, 2, 840–855. [Google Scholar] [CrossRef] [PubMed]
  40. Burdasal, P.; Legarda, T. Occurrence of ochratoxin A in sweet wines produced in Spain and other countries. Food Addit. Contam. 2007, 24, 976–986. [Google Scholar] [CrossRef] [PubMed]
  41. Freire, L.; Braga, P.A.C.; Furtado, M.M.; Delafiori, J.; Dias-Audibert, L.; Pereira, G.E.; Reyes, F.G.; Catharino, R.R.; Sant’Ana, A.S. From grape to wine: Fate of ochratoxin A during red, rose and white wine making process and of ochratoxin derivate in the final products. Food Control 2020, 113, 107167. [Google Scholar] [CrossRef]
  42. Domijan, A.M.; Peraica, M. Ochratoxin A in wine. Arh. Hig. Rada Toksikol. 2005, 56, 17–20. [Google Scholar] [PubMed]
  43. Brera, C.; Debegnach, F.; Minardi, V.; Prantera, E.; Pannunzi, E.; Faleo, S.; de Santis, B.; Miraglia, M. Ochratoxin A contamination in Italian wine samples and evaluation of the exposure in the Italian population. J. Agric. Food Chem. 2008, 56, 10611–10618. [Google Scholar] [CrossRef] [PubMed]
  44. Freire, L.; Passamani, F.R.; Thomas, A.B.; Nassur, R.C.; Silva, L.M.; Paschoal, F.N.; Pereira, G.E.; Prado, G.; Batista, L.R. Influence of physical and chemical characteristics of wine grapes on the incidence of Penicillium and Aspergillus fungi in grapes and ochratoxin A in wines. Int. J. Food Microbiol. 2017, 241, 181–190. [Google Scholar] [CrossRef]
  45. Freire, L.; Guerreiro, T.M.; Carames, E.T.S.; Lopes, L.S.; Orlando, E.A.; Pereira, G.E.; Pallone, J.A.L.; Catharino, R.R.; Sant’Ana, A.S. Influence of maturation stages in different varieties of wine grapes (Vitis vinifera) on the production of ochratoxin A and its modified forms by Aspergillus carbonarius and Aspergillus niger. J. Agric. Food Chem. 2018, 66, 8824–8831. [Google Scholar] [CrossRef]
  46. Pietri, A.; Bertuzzi, P.; Pallaroni, L.; Piva, G. Occurrence of ochratoxin A in Italian wines. Food Addit. Contam. 2001, 18, 647–654. [Google Scholar] [CrossRef] [PubMed]
  47. Clouvel, P.; Bonvarlet, L.; Martinez, A.; Lagouarde, P.; Dieng, I.; Martin, P. Wine contamination by ochratoxin A in relation to vine environment. Int. J. Food Microbiol. 2008, 123, 74–80. [Google Scholar] [CrossRef] [PubMed]
  48. Di Stefano, V.; Avellone, G.; Pintozzo, R.; Capocchiano, V.G.; Mazza, A.; Cicero, N.; Dugo, G. Natural co-occurrence of ochratoxin A, ochratoxin B and aflatoxins in Sicilian red wines. Food Addit. Contam. Part. A 2015, 32, 1343–1351. [Google Scholar] [CrossRef] [PubMed]
  49. Stefanaki, I.; Foufa, E.; Tsatsou-Dritsa, A.; Photis, D. Ochratoxin A concentration in Greek domestic vines and dried wine fruits. Food Addit. Contam. 2003, 20, 74–83. [Google Scholar] [CrossRef] [PubMed]
  50. Belli, N.; Marin, S.; Duaigues, A.; Ramos, A.J.; Sanchis, V. Ochratoxin A in wines, musts and grape juices from Spain. J. Sci. Food Agric. 2004, 84, 591–594. [Google Scholar] [CrossRef]
  51. Ponsone, M.L.; Chiotta, M.L.; Combina, M.; Torres, A.; Knass, P.; Dalcero, A.; Chulze, S. Natural occurrence of ochratoxin A in musts, wines and grape vine fruits from grapes harvested in Argentina. Toxins 2010, 2, 1984–1996. [Google Scholar] [CrossRef] [PubMed]
  52. Marino-Repizo, L.; Gargantini, R.; Manzano, H.; Raba, J.; Cerutti, S. Assesment of ochratoxin A occurrence in Argentinian red wines using a novel sensitive quichers-solid phase extraction approach prior to ultra high-performance liquid chromatography mass spectrometry methodology. J. Sci. Food Agric. 2017, 97, 2487–2497. [Google Scholar] [CrossRef] [PubMed]
  53. Kruger, C.D.; Fernandes, A.M.; Rosa, C.A. Ochratoxin A in wines from 2002 to 2008 harvest marketed in Rio de Janeiro, Brazil. Food Addit. Contam. Part B Surveill. 2012, 5, 204–207. [Google Scholar] [CrossRef]
  54. Vega, M.; Rios, G.; von Baer, D.; Marones, C.; Tessini, C.; Herlitz, E.; Saelzer, R.; Ruiz, M.A. Ochratoxin A occurrence in wines produced in Chile. Food Control 2012, 28, 147–150. [Google Scholar] [CrossRef]
  55. Zhong, Q.D.; Li, G.H.; Wang, D.B.; Shao, Y.; Li, J.G.; Xiong, Z.H.; Wu, Y.N. Exposure assessment to ochratoxin A in Chinese wine. J. Agric. Food Chem. 2014, 62, 8908–8913. [Google Scholar] [CrossRef]
  56. Peraica, M.; Flajs, D.; Domijan, A.M.; Ivic, D.; Cvjetkovic, B. Ochratoxin A contamination in food from Croatia. Toxins 2010, 2, 2098–2105. [Google Scholar] [CrossRef]
  57. Sage, L.; Garon, D.; Seigle-Morandi, F. Fungal microflora and ochratoxin A risk in French vineyards. J. Agric. Food Chem. 2004, 52, 5764–5768. [Google Scholar] [CrossRef]
  58. Labrinea, E.P.; Natskoulis, P.I.; Spiropoulus, A.E.; Magan, N.; Tassou, C.C. A survey of ochratoxin A occurrence in Greek wines. Food Addit. Contam. Part. B Surveill. 2011, 4, 4–6. [Google Scholar] [CrossRef]
  59. Di Stefano, V.; Pitonzo, R.; Avellone, G.; Di Fiore, A.; Monte, L.; Ogorka, A.Z.T. Determination of aflatoxin and ochratoxins in Sicilian weet wines by high-perfomance liquid chromatography with fluorometric detection and immunoaffinity clean up. Food Anal. Methods 2015, 8, 569–577. [Google Scholar] [CrossRef]
  60. Czerwiecki, L.; Wilczynska, G.; Kwiecien, A. Ochratoxin A: An improvement clean-up and HPLC method used to investigate wine and juice on Polish market. Food Addit. Contam. 2005, 22, 158–162. [Google Scholar] [CrossRef] [PubMed]
  61. Pena, A.; Cerejo, F.; Silva, L.; Lino, C. Ochratoxin A survey in Portuguese wine by LC-FD with direct injection. Talanta 2010, 82, 1556–1561. [Google Scholar] [CrossRef]
  62. Shephard, G.S.; Fabiani, A.; Stockenstom, S.; Mshicileli, N.; Sewram, V. Quantitation of ochratoxin A in South African wines. J. Agric. Food Chem. 2003, 51, 1102–1106. [Google Scholar] [CrossRef] [PubMed]
  63. De Jesus, C.L.; Bartley, A.; Welch, A.Z.; Berry, J.P. High incidence and levels of Ochratoxin A in wines sourced from United States. Toxins 2018, 10, 1. [Google Scholar] [CrossRef] [PubMed]
  64. Barata, A.; Malfeito-Ferreira, M.; Loureiro, V. The microbial ecology of wine grape berries. Int. J. Food Microbiol. 2012, 153, 243–259. [Google Scholar] [CrossRef]
  65. Mondani, L.; Palumbo, R.; Tsitsigiannis, D.; Perdikis, D.; Mazzoni, E.; Battilani, P. Pest management and ochratoxin A contamination in grapes: A review. Toxins 2020, 12, 303. [Google Scholar] [CrossRef]
  66. Visconti, F.; López, R.; Olego, M.Á. The Health of Vineyard Soils: Towards a Sustainable Viticulture. Horticulturae 2024, 10, 154. [Google Scholar] [CrossRef]
  67. Cozzi, G.; Haidukowski, M.; Perrone, G.; Visconti, A.; Logrieco, A. Influence of Lobesia botrana field control on black aspergilli rot and ochratoxin A contamination in grapes. J. Food Prot. 2009, 72, 894–897. [Google Scholar] [CrossRef] [PubMed]
  68. Cozzi, G.; Somma, S.; Haidukowski, M.; Logrieco, A.F. Ochratoxin A management in vineyards by Lobesia botrana biocontrol. Toxins 2013, 5, 49–59. [Google Scholar] [CrossRef] [PubMed]
  69. Mitchell, D.; Parra, R.; Alfred, D.; Magan, N. Water and temperature relations of growth and ochratoxin A production by Aspergillus carbonarius strains from grapes in Europe and Israel. J. Appl. Microbiol. 2004, 97, 439–445. [Google Scholar] [CrossRef] [PubMed]
  70. Passamani, F.R.; Hernandes, T.; Lopes, N.A.; Bastoc, S.C.; Santiago, W.D.; Cardoso, M.; Batista, L.R. Effect of temperature, water activity and pH on growth and production of ochratoxin A by Aspergillus niger and Aspergillus carbonarius from Brazilian grapes. J. Food Prot. 2014, 77, 1947–1952. [Google Scholar] [CrossRef] [PubMed]
  71. Visconti, A.; Perrone, G.; Cozzi, G.; Solfrizzo, M. Managing ochratoxin A risk in the grape-wine food chain. Food Addit. Contam. 2008, 25, 193–202. [Google Scholar] [CrossRef]
  72. Reverberi, M.; Gazzetti, K.; Punelli, F.; Scarpari, M.; Zjalic, S.; Ricelli, A.; Fabbri, A.A.; Fanelli, C. Aopyap1 regulates OTA synthesis by controlling cell redox balance in Aspergillus ochraceus. Appl. Microbiol. Biotechnol. 2012, 95, 1293–1304. [Google Scholar] [CrossRef]
  73. Crespo-Sempere, A.; Selma Lazaro, C.; Palumbo, J.D.; Gonzales-Candelas, L.; Martinez-Culebras, P.V. Effect of oxidants and phenolic antioxidants on the ochratoxigenic fungus Aspergillus carbonarius. J. Sci. Food Agric. 2015, 96, 169–177. [Google Scholar] [CrossRef] [PubMed]
  74. Ortiz-Villeda, B.; Lobos, O.; Aguilar-Zuniga, K.; Carrasco-Sanchez, V. Ochratoxins in wine: A review of their occurrence in the last decade, toxicity and exposure risk in humans. Toxins 2021, 13, 478. [Google Scholar] [CrossRef]
  75. Zjalic, S.; Frece, J.; Pavlovic, M.; Jakopovic, Z.; Markov, K. Procjena Rizika Kontaminacije Vina Okratoksinom A. Brošura za Vinare. (Croatian); Markov, K., Zjalic, S., Frece, J., Eds.; Sveučilište u Zadru: Zadar, Croatia, 2017; p. 48. ISBN 978-953-331-173-9. [Google Scholar]
  76. Tsolakis, H.; Corona, O.; Pulizzi, A.S.; Grippi, F.; Mondello, V. Incidence of grapevine moth Lobesia botrana (Den. & Schif.) on occurrence of ochratoxin A in grapes. In Integrated Protection in Viticulture; IOBC/WPRS Bulletin: Dijon, France, 2008; pp. 363–368. [Google Scholar]
  77. Wang, L.; Cai, R.; Zhang, J.; Liu, X.; Wang, S.; Ge, Q.; Zhao, Z.; Yue, T.; Yuan, Y.; Wang, Z. Removal of ochratoxin A in wine by Cryptococcus albidus and safety assessment of degradation products. J. Sci. Food Agric. 2023, 104, 2030–2037. [Google Scholar] [CrossRef]
  78. Farbo, M.G.; Urgeghe, P.P.; Fiori, S.; Marcello, A.; Oggiano, S.; Balmas, V.; Hassan, Z.U.; Jaoua, S.; Migheli, Q. Effect of yeast volatile organic compounds on ochratoxin A-producing Aspergillus carbonarius and A. ochraceus. Int. J. Food Microbiol. 2018, 284, 1–10. [Google Scholar] [CrossRef] [PubMed]
  79. Commission regulation (EU). No 2023/915 of 25 April 2023 on Maximum Levels of Certain Contaminants in Food and Replacing Regulation 2023. (EC) No 1881/2006. Available online: https://eur-lex.europa.eu/legal-content/EN/TXT/PDF/?uri=CELEX:32023R0915 (accessed on 30 February 2024).
  80. Bhatnagar, D.E.E.P.A.K.; Ehrlich, K.C.; Cleveland, T.E. Handbook of Applied Mycology, 5th ed.; Marcel Dekker Inc.: New York, NY, USA, 1991; pp. 255–286. [Google Scholar]
  81. Loi, M.; Fanelli, F.; Liuzzi, V.C.; Logrieco, A.F.; Mulè, G. Mycotoxin Biotransformation by Native and Commercial Enzymes: Present and Future Perspectives. Toxins 2017, 9, 111. [Google Scholar] [CrossRef] [PubMed]
  82. Shetty, P.H.; Jespersen, L. Saccharomyces cerevisiae and lactic acid bacteria as potential mycotoxin decontamination agents. Trends Food Sci. Technol. 2006, 17, 48–55. [Google Scholar] [CrossRef]
  83. Petruzzi, L.; Bevilacqua, A.; Corbo, M.R.; Garofalo, C.; Baiano, A.; Sinigaglia, M. Selection of autochtonous Saccharomyces cerevisiae strains as wine starters using a polyphasic approach and ochratoxin A removal. J. Food Prot. 2014, 77, 1168–1177. [Google Scholar] [CrossRef] [PubMed]
  84. Wang, L.; Wang, Q.; Wang, S.; Cai, R.; Yuan, Y.; Yue, T.; Wang, Z. Bio-control on the contamination of Ochratoxin A in food: Current research and future prospects. Curr. Res. Food Sci. 2022, 5, 1539–1549. [Google Scholar] [CrossRef] [PubMed]
  85. Shimizu, S.; Kataoka, M.; Honda, K.; Sakamoto, K. Lactone-ring-cleaving enzymes of microorganisms: Their diversity and applications. J. Biotechnol. 2001, 92, 187–194. [Google Scholar] [CrossRef] [PubMed]
  86. Abrunhosa, L.; Paterson, R.R.M.; Venâncio, A. Biodegradation of Ochratoxin A for Food and Feed Decontamination. Toxins 2010, 2, 1078–1099. [Google Scholar] [CrossRef]
  87. Li, S.; Marquardt, R.R.; Frohlich, A.A.; Vitti, T.G.; Crow, G. Pharmacokinetics of ochratoxin A and its metabolites in rats. Toxicol. Appl. Pharmacol. 1997, 145, 82–90. [Google Scholar] [CrossRef]
  88. Abrunhosa, L.; Serra, R.; Venâncio, A. Biodegradation of ochratoxin A by fungi isolated from grapes. J. Agric. Food Chem. 2002, 50, 7493–7496. [Google Scholar] [CrossRef]
  89. Abrunhosa, L.; Inês, A.; Rodrigues, A.I.; Guimarães, A.; Pereirab, V.L.; Parpot, P.; Mendes-Faia, A.; Venâncio, A. Biodegradation of ochratoxin A by Pediococcus parvulus isolated from Douro wines. Int. J. Food Microbiol. 2014, 188, 45–52. [Google Scholar] [CrossRef]
  90. Bueno, D.J.; Casale, C.H.; Pizzolitto, R.P.; Salano, M.A.; Olivier, G. Physical adsorption of aflatoxin B1 by lactic acid bacteria and Saccharomyces cerevisiae: A theoretical model. J. Food Prot. 2007, 70, 2148–2154. [Google Scholar] [CrossRef]
  91. Fuchs, S.; Sontag, G.; Stidl, R.; Ehrlich, V.; Kundi, M.; Knasmüller, S. Detoxification of patulin and ochratoxin A, two abundant mycotoxins, by lactic acid bacteria. Food Chem. Toxicol. 2008, 46, 1398–1407. [Google Scholar] [CrossRef]
  92. Hathout, A.S.; Aly, S.E. Biological detoxification of mycotoxins: A review. Ann. Microbiol. 2014, 64, 905–919. [Google Scholar] [CrossRef]
  93. Perczak, A.; Goliński, P.; Bryła, M.; Waśkiewicz, A. The efficiency of lactic acid bacteria against pathogenic fungi and mycotoxins. Arch. Ind. Hyg. Toxicol. 2018, 69, 32–45. [Google Scholar] [CrossRef]
  94. Mateo, E.M.; Medina, A.; Mateo, F.; Valle-Algarra, F.M.; Pardo, I.; Jiménez, M. Ochratoxin A removal in synthetic media by living and heat-inactivated cells of Oenococcus oeni isolated from wines. Food Control 2010, 21, 23–28. [Google Scholar] [CrossRef]
  95. Piotrowska, M.; Masek, A. Saccharomyces cerevisiae cell wall components as tools for ochratoxin A decontamination. Toxins 2015, 7, 1151–1162. [Google Scholar] [CrossRef]
  96. Jakopović, Ž.; Hanousek Čiča, K.; Mrvčić, J.; Pucić, I.; Čanak, I.; Frece, J.; Pleadin, J.; Stanzer, D.; Zjalić, S.; Markov, K. Properties and fermentation activity of industrial yeasts Saccharomyces cerevisiae, S. uvarum, Candida utilis and Kluyveromyces marxianus exposed to AFB1, OTA and ZEA. Food Technol. Biotechnol. 2018, 56, 208–217. [Google Scholar] [CrossRef]
  97. Pfliegler, W.P.; Pusztahelyi, T.; Pócsi, I. Mycotoxins—Prevention and decontamination by yeasts. J. Basic. Microbiol. 2015, 55, 805–818. [Google Scholar] [CrossRef]
  98. Bejaoui, H.; Mathieu, F.; Taillandier, P.; Lebrihi, A. Ochratoxin A removal in synthetic and natural grape juices by selected oenological Saccharomyces strains. J. Appl. Microbiol. 2004, 97, 1038–1044. [Google Scholar] [CrossRef]
  99. Patharajan, S.; Reddy, K.R.N.; Karthikeyan, V.; Spadaro, D.; Lore, A.; Gullino, M.L.; Garibaldi, A. Potential of yeast antagonists on in vitro biodegradation of ochratoxin A. Food Control 2011, 22, 290–296. [Google Scholar] [CrossRef]
  100. Jiang, C.; Li, Z.; Shi, Y.; Guo, D.; Pang, B.; Chen, X.; Shao, D.; Liu, Y.; Shi, J. Bacillus subtilis inhibits Aspergillus carbonarius by producing iturin A, which disturbs the transport, energy metabolism, and osmotic pressure of fungal cells as revealed by transcriptomics analysis. Int. J. Food Microbiol. 2020, 330, 108783. [Google Scholar] [CrossRef] [PubMed]
  101. Liu, W.C.; Pushparaj, K.; Meyyazhagan, A.; Arumugam, V.A.; Pappuswamy, M.; Bhotla, H.K.; Baskaran, R.; Issara, U.; Balasubramanian, B.; Khaneghah, A.M. Ochratoxin A as an alarming health threat for livestock and human: A review on molecular interactions, mechanism of toxicity, detection, detoxification, and dietary prophylaxis. Toxicon 2022, 213, 59–75. [Google Scholar] [CrossRef] [PubMed]
  102. Zanon, M.S.A.; Barros, G.G.; Chulze, S.N. Non-aflatoxigenic Aspergillus flavus as a potential biocontrol agents to reduce aflatoxin contamination in the peanuts harvested in Northen Argentina. Int. J. Food Microbiol. 2016, 231, 63–68. [Google Scholar] [CrossRef] [PubMed]
  103. Llobregat, B.; Gonzales-Candelas, L.; Bellester, A.L. Ochratoxin A deffective Aspergillus carbonarius mutants as potential biocontrol agents. Toxins 2022, 14, 745. [Google Scholar] [CrossRef] [PubMed]
  104. Cubaiu, L.; Abbas, H.; Dobson, A.D.W.; Burdoni, M.; Migheli, Q. Saccharomyces cerevisiae wine strain inhibits growth and decreases ochratoxin A biosynthesis by Aspergillus carbonarius and Aspergillus ochraceus. Toxins 2013, 4, 1468–1681. [Google Scholar] [CrossRef] [PubMed]
  105. Fiori, S.; Urgheghe, P.P.; Hammami, W.; Razzu, S.; Jaoua, S.; Migheli, Q. Biocontrol of four non- and low-fermenting yeast strains against Aspergillus carbonarius and their ability to remove ochratoxin A from grape juice. Int. J. Food Microbiol. 2014, 189, 45–50. [Google Scholar] [CrossRef] [PubMed]
  106. Piotrowska, M.; Zakowska, Z. The elimination of ochratoxin A by lactic acid bacteria strains. Pol. J. Microb. 2005, 54, 279–286. [Google Scholar]
  107. Ali-Vehmas, T.; Rizzo, A.; Westermarck, T.; Atroshi, F. Measurement of antibacterial activities of T-2 toxin, deoxynivalenol, ochratoxin A, aflatoxin B1 and fumonisin B1 using microtitration tray-based turbidimetric techniques. J. Vet. Med. Ser. A 1998, 45, 453–458. [Google Scholar] [CrossRef]
  108. Šain, A. Changes in Morphological Characteristics of Selected Wine Yeasts and Lactic Acid Bacteria in the Presence of Ochratoxin A. Bachelor Thesis, University of Zagreb, Zagreb, Croatia, 19 September 2017. [Google Scholar]
  109. Pulvirenti, A.; De Vero, L.; Blaiotta, G.; Sidari, R.; Iosca, G.; Gullo, M.; Caridi, A. Selection of wine Saccharomyces cerevisiae strains and their screening for the adsorption activity of pigments, phenolics and ochratoxin A. Fermentation 2020, 6, 80. [Google Scholar] [CrossRef]
  110. Jakopović, Ž. Effect of Selected Wine Yeast Strains on Binding, Degradation and Toxicity of Ochratoxin A In Vitro. Ph.D. Thesis, University of Zagreb, Zagreb, Croatia, 8 October 2021. [Google Scholar]
  111. Freire, L.; Furtado, M.M.; Guerreiro, T.M.; da Graca, J.S.; da Silva, B.S.; Oliveira, D.N.; Catharino, R.R.; Sant’ana, A.S. The presence of ochratoxin A does not influence Saccharomyces cerevisiae growth kinetics but leads to the formation of modified ochratoxins. Food Chem. Toxicol. 2019, 133, 110756. [Google Scholar] [CrossRef]
  112. Markov, K.; Mrvčić, J.; Pleadin, J.; Delaš, F.; Frece, J. FT-IR spectroscopy applied to identification of functional groups on the surface of Gluconobacter oxydans involved in the bacteria–ochratoxin A interaction. In Proceedings of the Power of Fungi and Mycotoxins in Health and Disease, Šibenik, Croatia, 20–23 September 2015. [Google Scholar]
  113. Markov, K.; Frece, J.; Pleadin, J.; Bevardi, M.; Barišić, L.; Kljusurić, J.G.; Vulić, A.; Jakopović, Ž.; Mrvčić, J. Gluconobacter oxydans–potential biological agent for binding or biotransformation of mycotoxins. World Mycotoxin J. 2019, 12, 153–161. [Google Scholar] [CrossRef]
  114. Mwabulili, F.; Xie, Y.; Li, Q.; Sun, S.; Yang, Y.; Ma, W. Research progress of ochratoxin a bio-detoxification. Toxicon 2023, 222, 107005. [Google Scholar] [CrossRef]
  115. Del Prete, V.; Rodriguez, H.; Carrascosa, A.V.; Rivas, B.D.L.; Garcia-Moruno, E.; Munoz, R. In vitro removal of ochratoxin A by wine lactic acid bacteria. J. Food Prot. 2007, 70, 2155–2160. [Google Scholar] [CrossRef] [PubMed]
  116. Piotrowska, M. The adsorption of ochratoxin A by Lactobacillus species. Toxins 2014, 6, 2826–2839. [Google Scholar] [CrossRef]
  117. Shukla, S.; Park, J.H.; Chung, S.H.; Kim, M. Ochratoxin A reduction ability of biocontrol agent Bacillus subtilis isolated from Korean traditional fermented food Kimchi. Sci. Rep. 2018, 8, 8039. [Google Scholar] [CrossRef]
  118. Niderkorn, V.; Morgavi, D.P.; Aboab, B.; Lemaire, M.; Boudra, H. Cell wall component and mycotoxin moieties involved in the binding of fumonisin B1 and B2 by lactic acid bacteria. J. Appl. Microbiol. 2009, 106, 977–985. [Google Scholar] [CrossRef]
  119. Dalie, D.K.D.; Deschamps, A.M.; Richard-Foget, F. Lactic acid bacteria—Potential for control of mould growth and mycotoxins: A review. Food Control 2010, 21, 370–380. [Google Scholar] [CrossRef]
  120. Rodriguez, H.; Reveron, I.; Doria, F.; Constantini, A.; De Las Rivas, B.; Munoz, R.; Garcia-Moruno, E. Degradation of ochratoxin A by Brevibacterium species. J. Agric. Food Chem. 2011, 59, 10755–10760. [Google Scholar] [CrossRef]
  121. Chen, W.; Li, C.; Zhang, B.; Zhou, Z.; Shen, Y.; Liao, X.; Yang, J.; Wang, Y.; Li, X.; Li, Y.; et al. Advances in biodetoxification of ochratoxin AA review of the past five decades. Front. Microbiol. 2018, 9, 1386. [Google Scholar] [CrossRef] [PubMed]
  122. Santos, J.; Castro, T.; Venâncio, A.; Silva, C. Degradation of ochratoxins A and B by lipases: A kinetic study unraveled by molecular modeling. Heliyon 2023, 9, e19921. [Google Scholar] [CrossRef]
  123. Cecchini, F.; Morassut, M.; Garcia Moruno, E.; Di Stefano, R. Influence of yeast strain on ochratoxin A content during fermentation of white and red must. Food Microbiol. 2006, 23, 411–417. [Google Scholar] [CrossRef] [PubMed]
  124. Rodrigues, I.; Binder, E.M.; Schatzmayr, G. Microorganisms and Their Enzymes for Detoxifying Mycotoxins Posing a Risk to Livestock Animals. In Mycotoxin Prevention and Control in Agriculture; Appel, M., Kendra, D.F., Trucksess, M.W., Eds.; American Chemical Society: Washington, DC, USA, 2009; pp. 107–117. [Google Scholar]
  125. Zhang, H.; Zhang, Y.; Yin, T.; Wang, J.; Zhang, X. Heterologous expression and characterization of a novel ochratoxin a degrading enzyme, N-acyl-L-amino acid amidohydrolase, from Alcaligenes faecalis. Toxins 2019, 11, 518. [Google Scholar] [CrossRef] [PubMed]
  126. Wei, M.; Dhanasekaran, S.; Ngea, G.L.N.; Godana, E.A.; Zhang, X.; Yang, Q.; Zheng, X.; Zhang, H. Cryptococcus podzolicus Y3 degrades ochratoxin A by intracellular enzymes and simultaneously eliminates citrinin. Biocontrol 2022, 168, 104857. [Google Scholar] [CrossRef]
  127. Wei, M.; Dhanasekaran, S.; Ji, Q.; Yang, Q.; Zhang, H. Sustainable and efficient method utilizing N-acetyl-L-cysteine for complete and enhanced ochratoxin A clearance by antagonistic yeast. J. Hazard. Mater. 2023, 448, 130975. [Google Scholar] [CrossRef]
  128. Varga, J.; Rigo, K.; Teren, J. Degradation of ochratoxin A by Aspergillus species. Int. J. Food Microbiol. 2000, 59, 1–7. [Google Scholar] [CrossRef]
  129. Stander, M.A.; Bornscheuer, W.T.; Henke, E.; Steyn, P.S. Screening of commercial hydrolases for the degradation of ochratoxin A. J. Agric. Food. Chem. 2000, 48, 5736–5739. [Google Scholar] [CrossRef]
  130. Abrunhosa, L.; Santos, L.; Venancio, A. Degradation of ochratoxin A by proteases and by a crude enzyme of Aspergillus niger. Food Biotechnol. 2006, 20, 231–242. [Google Scholar] [CrossRef]
  131. Abrunhosa, L.; Venancio, A. Isolation and purification of an enzyme hydrolyzing ochratoxin A from Aspergillus niger. Biotech. Lett. 2007, 29, 1909–1914. [Google Scholar] [CrossRef]
  132. Dobritzsch, D.; Wand, H.; Schneider, G.; Yu, S. Structural and functional characterization of ochratoxinase, a novel mycotoxin-degrading enzyme. Biochem. J. 2014, 462, 441–452. [Google Scholar] [CrossRef]
  133. Liuzzi, V.C.; Fanelli, F.; Tristezza, M.; Haidukowski, M.; Picardi, E.; Manzari, C.; Lionetti, C.; Grieco, F.; Logrieco, A.F.; Thon, M.R.; et al. Transcriptional analysis of Acinetobacter sp neg1 capable of degrading Ochratoxin A. Front. Microbiol. 2016, 7, 2162. [Google Scholar] [CrossRef]
  134. Dellafiora, L.; Gonaus, C.; Streit, B.; Galaverna, G.; Moll, W.-D.; Vogtentanz, G.; Schatzmayr, G.; Dall’Asta, C.; Prasad, S. An In Silico Target Fishing Approach to Identify Novel Ochratoxin A Hydrolyzing Enzyme. Toxins 2020, 12, 258. [Google Scholar] [CrossRef]
  135. Peng, M.; Zhang, Z.; Xu, X.; Zhang, H.; Zhao, Z.; Liang, Z. Purification and characterization of the enzymes from Brevundimonas naejangsanensis that degrade ochratoxin A and B. Food Chem. 2023, 419, 135926. [Google Scholar] [CrossRef]
  136. Wei, W.; Qian, Y.; Wu, Y.; Chen, Y.; Peng, C.; Luo, M.; Xu, J.; Zhou, Y. Detoxification of ochratoxin A by Lysobacter sp. CW239 and characteristics of a novel degrading gene carboxypeptidase cp4. Environ. Pollut. 2020, 258, 113677. [Google Scholar] [CrossRef]
  137. Luo, H.; Wang, G.; Chen, N.; Fang, Z.; Xiao, Y.; Zhang, M.; Gerelt, K.; Qian, Y.; Lai, R.; Zhou, Y. A superefficient Ochratoxin A hydrolase with promising potential for industrial applications. Appl. Environ. Microbiol. 2022, 88, e01964-21. [Google Scholar] [CrossRef]
  138. Pitout, M.J. The hydrolysis of Ochratoxin A by some proteolytic enzymes. Biochem. Pharmacol. 1969, 18, 485–491. [Google Scholar] [CrossRef]
  139. Bejaoui, H.; Mathieu, F.; Taillandier, P.; Lebrihi, A. Biodegradation of Ochratoxin A by Aspergillus section Nigri species isolated from French grapes: A potential means of Ochratoxin A decontamination in grape juices and musts. FEMS Microbiol. Lett. 2006, 255, 203–208. [Google Scholar] [CrossRef]
  140. Shi, L.; Liang, Z.; Xu, S.; Zheng, H.; Huang, K. Adsorption and degradation of Ochratoxin A by Bacillus licheniformis Sl-1. J. Agric. Biotechnol. 2013, 21, 1420–1425. [Google Scholar]
  141. Shi, L.; Liang, Z.; Li, J.; Hao, J.; Xu, Y.; Huang, K.; Tian, J.; He, X.; Xu, W. Ochratoxin A biocontrol and biodegradation by Bacillus subtilis CW 14. J. Sci. Food Agric. 2014, 94, 1879–1885. [Google Scholar] [CrossRef]
  142. Böswald, C.; Engelhardt, G.; Vogel, H.; Wallnöfer, P.R. Metabolism of the Fusarium mycotoxins zearalenone and deoxynivalenol by yeast strains of technological relevance. Nat. Toxins 1995, 3, 138–144. [Google Scholar] [CrossRef]
Figure 1. European regions considered at higher (circled in red) and high (painted in red) risk of wine contamination with ochratoxin A [12,30,42,43,44,46,47,48].
Figure 1. European regions considered at higher (circled in red) and high (painted in red) risk of wine contamination with ochratoxin A [12,30,42,43,44,46,47,48].
Toxins 16 00277 g001
Figure 2. Possible metabolites of OTA as products of microbial detoxification methods (OTA—ochratoxin A; OTB—ochratoxin B; OTC—ochratoxin C; OTα—ochratoxin α; OTβ—ochratoxin β; OP-OTA—lactone-opened ochratoxin A; 4(S)-OH OTA—epimer of 4-hydroxyochratoxin A (4-OH OTA); 10-OH OTA—10-hydroxyochratoxin A; OT…—metabolites that remain to be characterized; ?—unknown metabolites).
Figure 2. Possible metabolites of OTA as products of microbial detoxification methods (OTA—ochratoxin A; OTB—ochratoxin B; OTC—ochratoxin C; OTα—ochratoxin α; OTβ—ochratoxin β; OP-OTA—lactone-opened ochratoxin A; 4(S)-OH OTA—epimer of 4-hydroxyochratoxin A (4-OH OTA); 10-OH OTA—10-hydroxyochratoxin A; OT…—metabolites that remain to be characterized; ?—unknown metabolites).
Toxins 16 00277 g002
Figure 3. Schematic representation of mechanisms of microbiological control of OTA (adsorption—OTA bounded on cell wall surfaces of bacteria/yeast; biotransformation—hydrolysis of the lactone ring of OTA; biodegradation—hydrolysis of the peptide bond of OTA).
Figure 3. Schematic representation of mechanisms of microbiological control of OTA (adsorption—OTA bounded on cell wall surfaces of bacteria/yeast; biotransformation—hydrolysis of the lactone ring of OTA; biodegradation—hydrolysis of the peptide bond of OTA).
Toxins 16 00277 g003
Table 2. An overview of the microorganisms reported to remove over 75% of OTA from media by adsorption.
Table 2. An overview of the microorganisms reported to remove over 75% of OTA from media by adsorption.
MicroorganismStrainOTA Removal (%)Source
Saccharomyces cerevisiae-82.8–85.1 in grape116
-73–9084,110,111
BS75–77116
EC 11829–9584
Y2881.8783
TP5
TT173
79–100/
47–86
109
Candida intermedia->8078
Debaryomyces hansenii->9884
Bacillus subtilis-78117
Gluconobacter oxydans->80113
Table 3. An overview of the microorganisms reported to degrade more than 75% of the OTA present in the media.
Table 3. An overview of the microorganisms reported to degrade more than 75% of the OTA present in the media.
Microorganism StrainOTA Degradation Rate (%)Source
Saccharomyces cerevisiae 83–8578, 84
Candida intermedia >8084
Debaryomyces hansenii >9884
Bacillus subtilis 7884, 117
Brevibacterium caseiRM101100120
DSM 20657 T
DSM 9657 ND
DSM 20658
RM101
100120
Brevibacterium linensDSM 20425 T100120
Brevibacterium iodinumDSM20626T100120
Brevibacterium epidermidisDSM 20660T100120
Lactobacillus kefiriKFLM381121
Pediococcus parvulusUTAD 16889 89
UTAD 33397 89
UTAD 33494 89
UTAD 33598 89
UTAD 473100 ± 089
Aspergillus niger >90%124
Rhizopus sp. 100124
Trichosporon
mycotoxinivorans
MTV100124
Table 4. Microbial enzymes able to degrade OTA, species that produce them, and degradation products.
Table 4. Microbial enzymes able to degrade OTA, species that produce them, and degradation products.
MicroorganismEnzymeDegradation Product(s)Source
Aspergillus nigerLipaseAcid L-β-phenylalanine and (PHE) and OTα129
Protease A (acid protease)OTα86
Prolyve PAC (acid protease)OTα86
Saccharomyces cerevisiaeCarboxypeptidase Y (CPY)OTα86
Stenotrophomonas acidaminiphila CW117Amido hydrolase ADH3OTα137
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zjalic, S.; Markov, K.; Loncar, J.; Jakopovic, Z.; Beccaccioli, M.; Reverberi, M. Biocontrol of Occurrence Ochratoxin A in Wine: A Review. Toxins 2024, 16, 277. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins16060277

AMA Style

Zjalic S, Markov K, Loncar J, Jakopovic Z, Beccaccioli M, Reverberi M. Biocontrol of Occurrence Ochratoxin A in Wine: A Review. Toxins. 2024; 16(6):277. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins16060277

Chicago/Turabian Style

Zjalic, Slaven, Ksenija Markov, Jelena Loncar, Zeljko Jakopovic, Marzia Beccaccioli, and Massimo Reverberi. 2024. "Biocontrol of Occurrence Ochratoxin A in Wine: A Review" Toxins 16, no. 6: 277. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins16060277

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