ArticlePDF AvailableLiterature Review

Abstract and Figures

Some fungal species of the genera Aspergillus, Penicillium, and Fusarium secretes toxic metabolites known as mycotoxins, has become a global concern that is toxic to different species of animals and humans. Biological mycotoxins detoxification has been studied by researchers around the world as a new strategy for the mycotoxin removal. Bacteria, fungi, yeast, molds, and protozoa are the main living organisms appropriate for the mycotoxin detoxification. Enzymatic and degradation sorption are the main mechanisms involved in microbiological detoxification of mycotoxins. Regardless of the method used, proper management tools that consist of before-harvest prevention and after-harvest detoxification, are required. Here, in this review we focus on the microbiological detoxification, and mechanisms involved in decontamination of mycotoxins.
Content may be subject to copyright.
Send Orders for Reprints to reprints@benthamscience.net
Infectious Disorders - Drug Targets, 2020, 20, 1-19 1
1871-5265/20 $65.00+.00 © 2020 Bentham Science Publishers
REVIEW ARTICLE
Microbiological Detoxification of Mycotoxins: Focus on Mechanisms and
Advances
Milad Abdi1, Arezoo Asadi1, Farajolah Maleki2, Ebrahim Kouhsari3,4,*, Azam Fattahi5,
Elnaz Ohadi1, Ensieh Lotfali6, Alireza Ahmadi4 and Zahra Ghafouri7
1Department of Microbiology, Faculty of Medicine, Iran University of Medical Sciences, Tehran, Iran; 2Department of
Laboratory Sciences, School of Allied Medical Sciences, Ilam University of Medical sciences, Ilam, Iran; 3Clinical Mi-
crobiology Research Center, Ilam University of Medical Sciences, Ilam, Iran; 4Laboratory Sciences Research Center,
Golestan University of Medical Sciences, Gorgan, Iran; 5Center for Research and Training in Skin Disease and Lep-
rosy, Tehran University of Medical Sciences, Tehran, Iran; 6Department of Medical Parasitology and Mycology,
School of Medicine, Shahid Beheshti University of Medical Sciences, Tehran, Iran; 7Department of Biochemistry, Bio-
physics and Genetics, Faculty of Medicine, Mazandaran University of Medical Sciences, Sari, Iran
ARTICLE HISTORY
Received: January 09, 2020
Revised: April 10, 2020
Accepted: April 10, 2020
DOI:
10.2174/1871526520666200616145150
Abstract: Some fungal species of the genera Aspergillus, Penicillium, and Fusarium secretes toxic
metabolites known as mycotoxins, have become a global concern that is toxic to different species
of animals and humans. Biological mycotoxins detoxification has been studied by researchers
around the world as a new strategy for mycotoxin removal. Bacteria, fungi, yeast, molds, and proto-
zoa are the main living organisms appropriate for the mycotoxin detoxification. Enzymatic and
degradation sorptions are the main mechanisms involved in microbiological detoxification of myco-
toxins. Regardless of the method used, proper management tools that consist of before-harvest pre-
vention and after-harvest detoxification are required. Here, in this review, we focus on the microbi-
ological detoxification and mechanisms involved in the decontamination of mycotoxins.
Keywords: Biological detoxification, microbiological decontamination, mycotoxins.
1. INTRODUCTION
The mycotoxins are secondary metabolites with low--
molecular-weight produced by some filamentous fungi or
molds, including: Aspergillus species (spp.), Alternaria spp.,
Penicillium spp., Fusarium spp., Claviceps spp., Stachy-
botrys spp., Alternaria spp., Boletus spp., Amanita spp., Co-
prinus spp., Myrothecium spp., Pithomyces spp. etc., may de-
velop on various foods and feeds, causing severe risks to hu-
man and animal health” [1-3]. Human exposure to myco-
toxins occurs via direct contact (consumption of plan-
t-derived foods and animal products) and inhalation (expo-
sure to air and dust containing toxins) [4]. Currently, more
than 300 mycotoxins are known and are estimated more than
25% of the world’s agricultural commodities to be contami-
nated with mycotoxins [1, 5, 6]. These mycotoxins are
known to be either carcinogenic (e.g. aflatoxin B1, ochra-
toxin A, fumonisin B1), estrogenic (zearalenone), neurotox-
ic (fumonisin B1), nephrotoxic (ochratoxin), dermatotoxic
(trichothecenes) or immunosuppressive (aflatoxin B1, ochra-
toxin A, and T-2 toxin) [5]. Some studies have reviewed the
health hazards of mycotoxicosis to humans or animals [2, 6].
The symptoms of a mycotoxicosis can be effected on energy
* Address correspondence to this author at the Clinical Microbiology Re-
search Center, Ilam University of Medical Sciences, Ilam, Iran;
Tel/Fax: +98(843)2227101; P.O. Box: 6939177143;
E-mail: Kouhsari-E@medilam.ac.ir
production, nervous system, inhibition of synthesis of nucle-
ic acids, hormonal activities, reproductive systems and im-
mune system [9], which depends on the type of mycotoxin,
the concentration and length of exposure, the age, health, di-
et, interactions with other toxins and sex of the exposed indi-
vidual [3, 7]. Several methods applied for mycotoxins detoxi-
fication are classified as physical, chemical and biological.
For example, the biological methods consist of the use of mi-
croorganisms to degrade mycotoxins. Summary of microbio-
logical detoxification of mycotoxins is displayed in Table 1.
In this paper, we reviewed recent advances and mechanisms
involved in microbiological detoxification of mycotoxins.
2. DETOXIFICATION BY PROBIOTIC
Probiotics are living microorganisms that put out in ade-
quate amounts can give a health benefit on the host. Some of
the protective roles of probiotics are antagonistic effects
against pathogens, anticarcinogenic and antimutagenic activi-
ties [10]. Different probiotic strains of Lactobacillus and Bifi-
dobacterium genera possess important and widely acknowl-
edged health-promoting and immunomodulatory properties.
Application of biocontrol agents (BCA) such as lactic acid
bacteria (LAB) as biological food preservatives has long
been used as a cost-effective and safe method of food stor-
age. Interaction between LAB and contaminating organisms
induces production of low molecular weight compounds:
phenyllactic acid, reuterin, cyclic dipeptides, benzoic acid,
2 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
Table 1. Summary of microbiological detoxification of mycotoxins
Type of
Mycotoxins
Type of Detoxifica-
tion Type of Microorganisms Rate of Detoxification
(%)
Mechanisms of
Action Ref.
Total afla-
toxins
Bacterial Detoxifi-
cation
Lactobacillus acidophilus ATCC 4495
Lactobacillus acidophilus ATCC 20552 61-100 extracellular
metabolites [177]
Lactobacillus acidophilus ATCC 20552
Lactobacillus rhamnosus TISTR 541
Lactobacillus sanfrasiscensis DSM 20451
Bifidobacterium angulatum DSMZ 20098
11.5-84.7 Cell-binding [178]
Bifidobacterium longum
Lactobacillus rhamnosus
Bifidobacterium species 420
Lactobacillus acidophilus
Lactobacillus acidophilus NCFM 150B
Lactobacillus casei Shirota
31.9-35.6 Cell-binding [179]
Bifidobacterium bifidum PTCC 1644 Lactobacillus fermentum
PTCC 1744 88.8-99.8
Cell-binding,
extracellular
metabolites
[28]
Fungal Detoxifica-
tion Candida parapsilosis IP1698 54.54-92.98 extracellular
metabolites [180]
Aflatoxin B1
Bacterial Detoxifi-
cation
Lactobacillus delbrueckii,
Streptococcus thermophilus 100 - [181]
Lactobacillus plantarum (PTCC 1058) 45 - [22]
Lactobacillus casei 14-49 - [23]
Lactobacillus casein 26.06 - [182]
Pleurotus ostreatus 81.7-89.4 - [183]
Bifidobacterium 18–49 - [25]
B. subtilis ANSB060 81.5 - [184]
Stenotrophomonas maltophilia 82.5 - [77]
Rhodococcus erythropolis 33 - [78]
R. erythropolis DSM 14303
Nocardia corynebacterioides DSM 12676, DSM 20151
Mycobacterium fluoranthenivorans DSM 44556
60 - ˃ 90 - [79]
Flavobacterium aurantiacum 34.5-94.5 - [81]
Bacillus subtilis 38.4 - [182]
Bacillus TUBF1 90 - [185]
Bacillus licheniformis CFR1 91.2 - [186]
Rhodococcus. globerulus AK36 95 - [85]
R. rhodochrous NI2 & ATCC 12674 96.1-99.98 - [187]
Staphylococcus warneri 47.4 - [188]
Lysinibacillus, Fusiformis, Sporosarcina sp. 46.9-61.3 - [188]
Staphylococcus sp. VGF2 56.8 - [189]
Pseudomonas Anguilliseptica VGF1 Pseudomonas fluorescens 51.7
47.7 - [189]
Brevundimonas sp. 78.1-76.8 [77]
L. paracasei BEJ01 30- 91.2
Decrease of
immunotoxicity
and oxidative
stress
[190]
lactic acid bacteria 85
Production of
organic acids;
acetic, citric and
lactic
acids
[191]
L. plantarum,
L. rhamnosus,
L. brevis,
Lactobacillus delbrueckii subsp. lactis
48.9-79.5 Cell-binding [192]
L. rhamnosus GG
L. rhamnosusLC-705
L. acidophilusLC1
L. lactissubsp.lactis
L. acidophilusATCC 4356
L. plantarum
L. caseiShirota
L. delbrueckiisubsp.bulgaricus
L. helveticus
P. freudenreichiisubsp.shermaniiJS
87- 97.3 Cell-binding [193]
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 3
(Table 1) contd....
Fungal detoxifica-
tion
Lactococcus lactissubsp.cremoris
Streptococcus thermophilus
Lactobacillus spp
Bifidobacteria spp
Lactococcus spp
5.6 - 59.7 Cell-binding
[25]
Lactobacillus plantarum 13-54 [194]
Lactobacillus pentosus
Lactobacillus beveris 17.4-34.7 Cell-binding
- [195]
Pseudomonas putida MTCC 1274
and 2445 10-90 extracellular
metabolites [164]
Aspergillus
niger D15-
Lcc2#3
55 enzymatic
degradation [103]
Peniophora sp SCC0152 52.4 enzymatic
degradation [103]
Pleurotus
treatus St2-3 76 enzymatic
degradation [103]
S. cerevisiae 10-53 - [18]
A. niger (ND-1) 58.2 - [107]
S. cerevisiae
Rhizopus oligosporus 22.6-64.5 - [106]
Candida utilis 21.08 -
[182]
Streptomyces sp. 1.6-88.4 - [196]
Alternaria sp. 83 - [197]
Trichoderma. Reesi QM9414 100 - [198]
Absidia repens 50-60 - [199]
Rhizopus
oryzae
NRRL 395
58 - [200]
Saccharomyces cerevisiae 90.4-99.3 Surface binding [201]
Phoma glomerata PG41 66-78 extracellular
metabolites [202]
R. oryzae
CCT7560 100 - [198]
Aspergillus
niger D15-
Lcc2#3
55 Enzymatic
degradation [103]
Peniophora sp SCC0152 52.4 Enzymatic
degradation [103]
Pleurotus
treatus St2-3 76 Enzymatic
degradation [103]
Aflatoxin
M1
Bacterial detoxifica-
tion
Lactobacillus acidophilus, Lactobacillus rhamnosus, Lactobacillus
casei
Bifidobacterium lactis
24.2-98.8 Cell-binding
[203]
Lactobacillus 29 - [26]
Bifidobacterium spp
Lactobacillus spp. 4.5-73.1 - [19]
LAB 12 - 46 - [20]
Lactobacillus rhamnosus 18.8 - [173]
Streptococcus thermophilus 14.8-39 -
[26]
Lactobacillus, Bifidobacterium 7.85−25.94 - [168]
Lactobacillus bulgaricus 87.6 - [204]
Lactobacillus. acidophilus 27 - [203]
ZEA Bacterial detoxifica-
tion
Lactobacillus casein 19.05 - [182]
Lactobacillus pentosus 81.7- 83.2 - [205]
Pediococcus pentosaceus 29.4-68.13 - [206]
Pseudomonas alcaliphila TH-C1
Pseudomonas plecoglossicida TH-L1
68
57 - [86]
Bacillus subtilis 42.18 - [182]
Lactobacillus plantarum 38.06-47.8 binding [207]
ZEA Fungal detoxifica-
tion
Gliocladiumroseum NRRL 1859 80–90 - [110]
Aspergillus nigri
Streptomyces spp 0-96 - [112]
4 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
(Table 1) contd....
Streptomyces griseus (ATCC 13273) Streptomyces rutgersensus
(NRRL-B 1256) Rhizopus arrhizus (IFO-6155)
Aspergillus niger (ATCC 111394)
S. rutgersensus (NRRL-B 1256)
F. oxysporum S-F3
10-40 - [114]
Trichoderma spp. 9 - 97 - [208]
Candida utilis 40.7 - [182]
Streptomyces sp. 35.4-99.6 - [196]
Gliocladiumroseum NRRL 1859 80–90 - [110]
Aspergillus nigri
Streptomyces spp 0-96 - [112]
Streptomyces griseus (ATCC 13273) Streptomyces rutgersensus
(NRRL-B 1256) Rhizopus arrhizus (IFO-6155)
Aspergillus niger (ATCC 111394)
S. rutgersensus (NRRL-B 1256)
F. oxysporum S-F3
10-40 - [114]
Trichoderma spp. 9 - 97 - [208]
Candida utilis 40.7 - [182]
Fumonisin
B1, B2
Bacterial and Fungal
detoxification
Lactic acid bacteria
Propionic acid bacteria 55-100 - [39]
L. paracasei BEJ01 30- 91.2
Decrease of
immunotoxicity
and oxidative
stress
[190]
Lactobacillus paraplantarum CNRZ 1885
Streptococcus thermophiles RAR1
Streptococcus thermophilus CNRZ 1066
Strep.
thermophilus JIM 8752
Lactococcus lactis subsp. cremoris MG1363
2-99
Cell-binding
(Peptidoglycan
and tricarballylic
acid)
[175]
S. cerevisiae 25->75 - [124]
Pleurotus eryngii 0-61 - [209]
Lactobacillus brevis. N195
Lactobacillus brevisN197
Klebsiella variicolaVA26
Bacillus sp. N339
Klebsiella oxytocaN186
Candida glabrrataN353
Candida glabrrataN345
Candida glabrrataN346
Piichia fermenttansN183
6-33 - [210]
PAT
Bacterial detoxifica-
tion
Fungal detoxifica-
tion
Lactobacillus plantarum 39 - [45]
Lactobacillus rhamnosus 80.4 - [47]
B. bifidum 671 52.9-54.1 - [51]
lactic
acid bacteria 47 - 80 biosorption [47]
Alicyclobacillus spp 11.4-60.2 [211]
LAB 2-97 Cell-binding [45]
Gluconobacter oxydans > 96 - [130, 212,
213]
P. ohmeri 158 > 83- >99 - [132]
Deoxyniva-
lenol
Bacterial detoxifica-
tion
Lactic acid bacteria
Propionic acid bacteria 55-100 - [39]
LAB 55 - [71]
lactic acid bacteria 16.41- 71.19 enzymatic
degradation [214]
L. plantarum (ATCC14917T),
L. brevis (ATCC14869T),
L. brevis (ATCC 8287),
Oenococcus oeni (RM8,RM11,RM1),
L. plantarum (RM28, RM35),
L. brevis (RM273), Leuconostoc mesenteroides (RM54), Pediococ-
cus acidilactici (RM86)
8 - 28 Cell-binding [57]
LAB 2-97 Cell-binding [45]
Fungal detoxifica-
tion S. cerevisiae > 37 - [124]
Beauvericin
Bacterial detoxifica-
tion
Probiotic
Bifidobacteria, lactobacilli, Eubacterium crispatus, Salmonella
fecalis, Salmonella termofilus
66-83 - [215]
OTA Fungal detoxifica-
tion
Saccharomyces sensu stricto 66 - 100 - [123]
Yeast 4.67- 99.87 extracellular
metabolites [216]
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 5
hydroxylated fatty acids, methylhydantoin, mevalonolac-
tone, lactic acid [11] and antagonistic compounds, such as
hydrogen peroxide, ethanol, and bacteriocins [12]. LAB pre-
vent germination of mould spores and reduce mycotoxin pro-
duction in food products [7]. The LAB could detoxify myco-
toxins through penetration of bacterial structure or the degra-
dation of its metabolism [8]. It is estimated that 25% of the
world’s crop production is contaminated by mycotoxins dur-
ing the pre-harvest period, transport, processing or storage
[9]. New preservation systems including antifungal LAB
strains could be used in the future as strategies for bio-con-
trolling mould growth [10].
2.1. Aflatoxins
Aflatoxins (AF) are secondary metabolites of fungal spe-
cies, including Aspergillus flavus, A. snomius and A. parasiti-
cus. AF is a group of potent mycotoxins with mutagenic, ter-
atogenic, carcinogenic, hepatotoxic and immunosuppressive
properties [11]. AFB1, AFB2, AFG1 and AFG2 are natural-
ly occurring aflatoxins. Aflatoxin structure permits the for-
mation of DNA adducts with guanine, inducing cancerous
cell formation [12]. Aspergillus flavus produces the B group
whereas A. parasiticus produces both B and G groups of afla-
toxins. AFB1 is the most common and toxic molecule
among aflatoxin’s family. Hydroxylation of AFB1, leading
to the formation of AFM1 (4-hydroxy-AFB1) [13]. This hy-
droxylation process appears to be realized in the rumen of
dairy mammals as the direct intake by these animals of afla-
toxin B1 (AFB1) contaminated feed leading to the accumula-
tion of AFM1 in milk. Thus, the AFM1 content in milk is
strongly correlated with the level of AFB1 in raw feedstuff.
AFM1 mycotoxin exerts a direct cytotoxic effect on human
cells that do not require prior metabolic activation. It is resis-
tant to pasteurization or sterilization [14]. The European
Community and Codex Alimentarius (ECCA) recommended
that the maximum limit of AFM1 in liquid milk is 0.05
mg/L [15]. According to USA regulations, the level of
AFM1 in milk should not be higher than 0.5 mg/L [16].
Some strains of LAB; Lactobacillus bulgaricus and Strepto-
coccus thermophilus have been reported to be effective in re-
moving AFB1 and AFM1 from contaminated liquid media
and milk [17]. Different binding sites may be present in dif-
ferent strains, and there are differences between binding
sites in each bacteria and variation, among them. AF binding
ability to both the microorganisms enhances by heating.
Heat treatments are also significantly enhancing the ability
of bacteria to degrade AFM1 [18]. Some researchers as-
sessed the ability of Bifidobacterium spp., Lactobacillus
spp. and Lactococcus spp. to bind AFM1 in solution; they
observed that the percentages of AFM1 binding to these
strains were 0 to 14.6% after 24 hours and 4.5% to 73.1% af-
ter 96 hours [19]. Also, previous data evaluated the ability
of LAB to remove AFM1 in PBS and skimmed milk. These
authors found that tested strains bound AFM1 within a range
from 12.42% to 45.67% for heat-inactivated bacteria and
within 5.6% to 33.54% for viable bacteria exposed to the
aflatoxin for 15 min or 24 h, respectively [20]. L. reuteri NR-
RL B-14171 strain bound up to 85.3% of AFB1 in PBS after
4 h. According to the study conducted by El-Nezami et al.,
binding of AFM1 is different to binding of AFB1 due to the
presence of an additional MeOH group in the AFM1
molecule, resulting in increasing polarity of the AFM1
molecule, making it more hydrophilic and thus increasing its
tendency to dissolve in aqueous solutions [21]. Khanafari et
al. (2007) studied the efficacy of L. plantarum PTCC 1058
to bind AFB1 and reported that 45% of AFB1 was removed
from the solution after 1 hour [22]. Mendoza et al. (2009)
screened eight strains of Lactobacillus casei for their ability
to bind AFB1, and the percentage bound ranged from 14 to
49%. They mentioned that Lactobacillus caseiShirota (LcS)
could bind AFB1 into the bacterial cell envelope and that
the documents also revealed that aflatoxin binding produced
structural changes that modified the bacterial cell surface
[23]. Shah and Wu assessed the antimutagenic activity of
killed and living cells of six strains of Lactobacillus aci-
dophilus on eight chemical mutagens and promutagens [24].
The destruction of specific components of the bacterial cell
wall, (carbohydrates and proteins), resulted in a reduction in
AFB1 binding. The amount of AFM1 is increased by reduc-
ing the acidity of the environment. Peltonen et al. found that
Bifidobacterium strains are bound to 18%–48.7% of AFB1.
They showed that the acid and heat treatment of the cells sig-
nificantly increased their ability to capture aflatoxin B1 in
the liquid medium [25]. Sarimehmetoğlu et al., reported per-
centages of AFM1 binding near to 29.42% for strains of the
genus Lactobacillus [26]. Also, high-performance liquid
chromatography (HPLC) analysis showed that quantitation
of AFB1 in supernatant samples of Lactobacillus spp., Bifi-
dobacterium spp. and Lactococcus spp. were 17.3 to 59.7%,
18.0 to 48.7%, and 5.6 to 41.1% of mentioned mycotoxin, re-
spectively [27]. Ghazvini et al. [28] assessed antifungal ef-
fects of two LAB against growth and aflatoxin production of
Aspergillus parasiticus. Their results showed that B. bifidum
and L. fermentum extensively decreased aflatoxin produc-
tion and growth rate in comparison with the controls
(p≤0.05). The most significant reduction (> 99%) by LAB in
total aflatoxins, B1, B2, G1 and G2 fractions was also report-
ed. Furthermore, bacterial metabolites reduced the level of
standard AFB1, B2, G1, and G2 from 88.8% to 99.8%
(p≤0.05).
2.2. Zearalenone (ZEA or ZEN)
The Fusarium family, including Fusarium cereals, F.
graminearum, F. semitectu, F. culmorum, and F. equiseti
produced ZEA. It can produce toxic secondary metabolites
(mycotoxins) that interfere with grain quality by destroying
its amylose, cellulose, and protein reserves, and reducing
productivity, consequently causing significant losses to in-
dustrial production and crops [29]. ZEA or ZEN is described
as 6-[10-hydroxy-6-oxotrans-1-undecenyl]-B-resorcyclic
acid lactone, similar to the natural estrogens [30]. It can
change their mechanism of metabolism and synthesis, and in-
terfere with the synthesis of the receptor, which contributes
to change neoplastic prostate cancer or breast cancer [31].
ZEA biotransformation in animals leads to the formation of
two essential metabolites, α-zearalenol and β-zearalenol (α--
6 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
ZOL and β-ZOL); alpha zearalenol is higher estrogenicity
than ZEA [32]. Streptomyces rutgersensus, Rhizopus ar-
rhizus., and Streptomyces griseus were able to metabolize
zearalenone to α- zearalenol. Although, L. plantarum., L.
rhamnosus and L. acidophilus can reduce the concentration
of mycotoxins in food [8]. Other microorganisms such as Ba-
cillus subtilis, Saccharomyces cerevisiae, Pseudomonas and
Rhizopus spp. are toxin neutralizing agents [33]. The mech-
anism of the toxin is not entirely understood, but there is
some evidence of cytogenic effect and apoptosis caused by
ZEA [19]. Lioi et al. [34] found that ZEA increased DNA
fragmentation in three cell lines (DOK, Caco-2, and Vero)
after 24 h [35]. ZEA and its metabolization can cause dam-
ages of bacterial cells and lead to death. Peptidoglycan struc-
ture and amino acid composition are important for ZEA bio-
control [36]. The authors mentioned that both heat-treated
and acid-treated bacteria were able to remove ZEA and α-
zearalenol. ZEA predominantly binds to carbohydrate of the
cell wall of LAB by using hydrophobic interactions. This
bacterium-mycotoxin is unstable because hydrophobic inter-
actions are relatively weak [37]. Fourier-transform infrared
(FTIR) spectroscopy indicates that in the neutralization of
ZEA by LAB, the deprotonated carboxyl group (mainly
from glutamine and asparagine) of bacterial proteins and
peptidoglycan are mainly involved. Finally, LAB seem to be
promising alternatives for the development of new anti-my-
cotoxin agents [38].
2.3. Fumonisins
The fumonisins are derived from Fusarium. They have
substantial structural analogue to sphinganine. Inhibit ce-
ramide synthase, inducing an adverse effect on the sphinga-
nine/sphingosine ratio. Amine groups play a vital role in fu-
monisin toxicity, so deamination of FB1 induces a loss of
toxicity. Niderkorn et al. showed that ability of L. paraplan-
tarum to bind fumonisins B1 and B2 (FB1, FB2) in foods
and feeds. Peptidoglycan and tricarballylic acid chains of
LAB and FB play a significant role in binding interactions
[39]. They revealed that FB2 was removed more efficiently
than FB1. The ability of three strains of Propionibacterium
to remove FB1 and FB2 from acidified MRS broth samples
(pH 4.0) was evaluated and demonstrated that FB1 was not
as effectively removed as FB2. The acidity of the environ-
ment affects the process of binding since, at pH 7, LAB is
not able to trap FB1 and FB2. Niderkorn suggested that pep-
tidoglycans were the most plausible fumonisin binding sites
[39].
One study showed that F. graminearum IDM623 was
sensitive to 25 different strains of LAB tested. These results
are following those reported by Corsetti et al. [40]. They
evaluated the inhibition of F. culmorum, F. avenaceum, F.
oxysporum, and F. graminearum by viable cell extracts of L.
plantarum and did not find any relation or influence among
the lactic acid concentration.
2.4. Patulin (PAT)
PAT (4-hydroxy-4H-furo [3, 2c] pirano- 2(6H)-one), is
produced during colonisation by ubiquitous mould and has
been shown to induce toxicosis in the rat and genotoxicity in
Chinese hamster V79 cells [41]- [42]. PAT is the major my-
cotoxin from P. expansum. The maximum accepted levels of
patulin in the EU is 50 mg/L for fruit juice [43]. PAT is pro-
duced by a large number of fungi within several genera such
as Byssochlamys, Eupenicillium, Penicillium, Paecilomyces,
and Aspergillus [44] and predominantly found in vegetables
and fruits (in particular in apples and apple-derived product-
s) [45]. The PAT causes oxidative damage, has a negative
impact on reproduction in males via interaction with hor-
mone production and affects the immune system. PAT
caused potent genotoxic effects in experiments with mam-
malian cells but is devoid of activity in bacterial mutagenici-
ty assays [46]. Since DNA damage is causally related to can-
cer induction and some other diseases, it was suspected that
PAT might possess carcinogenic properties, LAB has been
reported to reduce the amount of PAT in liquid media [47].
Other studies using S. cerevisiae, Rhodosporidium kra-
tochvilovae and Gluconobacter oxydans demonstrated detox-
ification of patulin to potentially less toxic compounds such
as estradiol and desoxypatulinic acid [48]. Both L. plan-
tarum B1 and S1 strains were able to degrade PAT, and the
degradation seemed to follow the well-documented pathway
to estradiol [49]. The degradation activity is performed by a
heat-labile factor that is present in the LAB and secreted and
is therefore likely to be protein based [47]. Fuchs et al.
studied the effect of the LAB on PAT and noticed that the
most substantial decline of PAT (39%) was detected by L.
plantarum [45]. Hateb et al. (2012) [50] reported that the
maximum PAT uptake from aqueous solution was achieved
by L. rhamnosus 6149 strains after 24-h incubation. Results
revealed that removal of PAT contamination from apple
juice by using 10 different inactivated LAB, showed that L.
rhamnosus caused a decrease of PAT by 80.4%. Recently,
they reported that the maximum uptake of PAT was
achieved by B. bifidum .671 (52.9%) for viable and (54.1%)
for killed cells after 24-h incubation [51].
2.5. Ochratoxins (OTAs)
OTAs are produced as secondary metabolites by several
fungi of Penicillium spp. or Aspergillus spp. Ochratoxin-A
(OTA) is the carcinogenic fungal toxin found in a variety of
food commodities. OTA causes severe nephrotoxic effects
in animals and humans, and its genotoxic and carcinogenic
effects are well documented [52]. Ochratoxins affect protein
synthesis and inhibit ATP production. It has been shown in a
few earlier investigations that LAB cause removal of OTA
from liquid medium [53, 54]. Lactobacillus spp. and Strepto-
coccus spp. have the ability to degrade of OTA in milk. Hei-
dler and Schatzmayr [55] demonstrated that Lactobacillus vi-
tulinus was able to cleave OTA into the amino acid pheny-
lalanine and the nontoxic metabolite ochratoxin α (OTα)
[56]. Del Prete et al. [57] performed a study on the interac-
tion between OTA and wine LAB (Lactobacillus spp., Oeno-
coccus oeni spp., Pediococcus spp.). They found that a de-
crease in OTA concentration by wine LAB strains. Lević et
al. indicated that the greatest adsorption (more than 50% of
the initial OTA content) was obtained with L. rhamnosus
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 7
GG, L. acidophilus, L. plantarum BS, L. sanfranciscensis
and L. brevis [58].
2.6. Trichothecene
Trichothecenes are one of the most important of myco-
toxins, leading to a significant economic impact on grain
crops and cereal each year [59]. Trichothecenes are pro-
duced by fungi of genera Fusarium spp., Cephalosporium
spp., Trichoderma spp., Spicellum spp., Myrothecium spp.,
Stachybotrys spp., and Trichothecum spp. Amphipathic tri-
chothecenes can move passively across cell membranes
[60]. They are easily absorbed via gastrointestinal systems
and have a rapid effect on proliferating tissues [61]. These
toxins can induce vomiting, hemorrhagic lesions, immuno-
logical problems, and skin dermatitis. They are also phyto-
toxic and can cause inhibition of root elongation, chlorosis,
and dwarfism [62]. Studies showed that trichothecenes had
prevented peptide bond formation at the peptidyl transferase
center of the 60S ribosomal subunit and inhibiting eukaryot-
ic protein synthesis [63]. This inhibition typically affects
elongation and polypeptide chain termination may also be in-
hibited. Trichothecenes inhibit mitochondrial protein synthe-
sis and interact with protein sulfhydryl groups [61]. The ac-
tivity of trichothecenes eventually produces harmful levels
of oxidative stress due to the production of free radicals.
Based on the substitution pattern of tricyclic 12,13-epoxytri-
chothec-9-ene (EPT), trichothecenes classified into four
groups (Types A, B, C, and D) [64]. Trichothecenes have
12,13-epoxytrichothecene group responsible for their cyto-
toxicity and 9,10-double bond with various side-chain substi-
tutions [65]. The position of the acetyl-ester and hydroxyl
groups can influence the relative toxicity within eukaryotic
cells. Their relative capacity to interfere with protein synthe-
sis has been attributed to a combination of different factors;
the rate of transport into cells, changes in affinity for the ac-
tive binding site, metabolism by cytosol enzymes or the abili-
ty to interfere with protein synthesis [66].
Deoxynivalenol (DON) is produced by certain Fusarium
spp. that infect corn, barley, rice, wheat, oats, and other
grains in the field or during storage. The risk for human is di-
rectly through foods of plant origin or indirectly through
foods of animal origin. DON affects animal and human
health causing acute temporary headache, nausea, vomiting,
abdominal pain, diarrhea, dizziness, and fever [67]. Tri-
chothecenes can be removed by a mechanism other than bio-
transformation or biodegradation. Polysaccharide and pepti-
doglycan of the cell wall are the two primary moieties re-
sponsible for the binding to Lactobacillus spp.. The thick-
ness of peptidoglycan may change by heating [68]. This per-
turbation of the bacterial cell wall may allow the tri-
chothecenes bind to the cell wall and plasma membrane con-
stituents that are not available when the bacterial cell is in-
tact [69]. DON and T-2 toxin induce apoptosis in hemopoiet-
ic progenitor cells and immune cells, and they also inhibit
the protein, ADN and ARN synthesis, whereas their toxicity
is induced by the epoxy structure [70]. Many of LAB strains
such as L. plantarum., L. acidophilus, L. helveticus., L. rham-
nosus, S. thermophilus, Lactobacillus casei strain GG have
the capacity for removing DON in acidified liquid medium
(pH 4) and the results indicated removal levels up to 55%
for DON [71]. Although the molecular structure of tri-
chothecenes is entirely different from aflatoxins, the mech-
anisms involved in binding aflatoxins and trichothecenes are
similar. These results indicate that the binding is a physical
phenomenon and that deoxynivalenol (DON), nivalenol
(NIV), and fusarenon-X (FX) are bound to the bacteria by
weak non-covalent interactions such as those associated
with hydrophobic pockets on the bacterial surface. A recent
study confirmed to reduce the intestinal absorption of myco-
toxins from the human diet and animal feeds [71]. DON-in-
duced significant increases in apoptosis rates compared to
controls in mouse thymocytes in vivo. Also, DON signifi-
cantly decreased the proliferation indexes of the treated
cells. Preliminary cultures had shown that DON induced ne-
crosis of the explants after 4 h of incubation [72].
3. OTHER BACTERIA
Sangare et al., 2014 [73] reported that Pseudomonas
aeruginosa N17-1 could degrade AFB1, AFB2, and AFM1,
82.8%, 46.8% and 31.9%, respectively. Some species of Ba-
cillus such as B. subtilis and B. licheniformis displayed afla-
toxin reduction activity [74, 75]. Gao et al. evaluated the
ability of B. subtilis (isolated from a different source) to
transform aflatoxins [76]. The results showed that B. subtilis
ANSB060 from the fish gut had the most potent ability to de-
toxify aflatoxins B1, M1, and G1 with 81.5%, 60%, and
80.7%, respectively. Guan et al. [77] found that Stenotropho-
monas maltophilia could reduce AFB1 by 82.5% in the
liquid medium at 37 °C for 72 hrs. Alberts et al., (2006) [78]
investigated the biological degradation of aflatoxins B1 (AF-
B1) in cell-free, extracellular extracts of Rhodococcus ery-
thropolis liquid cultures. They reported a significant reduc-
tion (33.2%) of AFB1 from 72 h when treated with R. ery-
thropolis extracellular extracts. Also, studies from Teniola et
al., (2005) [79] were conducted on biological degradation of
AFB1 by cell-free extracts of four bacterial strains: R. ery-
thropolis DSM 14303, Nocardia corynebacterioides DSM
12676, N. corynebacterioides DSM 20151, and Mycobacteri-
um fluoranthenivorans nov. DSM 44556 in cell-free extracts
under different incubation conditions. A significant detoxifi-
cation activity rates of aflatoxin B1 after 24 h by cell-free ex-
tracts of N. corynebacterioides DSM 12676 and DSM 20151
were 60%, ˃ 90%, respectively. Also, R. erythropolis and
M. fluoranthenivorans nov. DSM 44556 T have shown ˃
90% degradation of AFB1 within 4 h at 30 °C. Biological
degradation of AFB1 was optimized under a range of temper-
atures from 25°C to 40 °C and pH values of 4 to 8 by using
three Actinomycete, Rhodococcus erythropolis ATCC 4277,
Streptomyces lividans TK 24, and Streptomyces aureofa-
ciens ATCC 10762, in liquid cultures [80]. All species were
able to degrade the AFB1 at the end of the first 24 h in a
range 0.81 to 2.78 µg/mL. A study evaluated the ability of
aqueus protein extract from Flavobacterium aurantiacum to
degradation of AFB1 [81]. Degradation of AFB1 after incu-
bation with heat-treated, non-heat-treated, DNase I–treated
and Proteinase K–treated of crude protein extracts were
8 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
94.5%, 74.5%, 80.5%, and 34.5% respectively. Some
studies on the bio-transformation of ZEA by the bacterial mi-
crobiome of pigs [82], mixed culture (Alcaligenes, Bacillus,
Achromobacter., Flavobacterium, and Pseudomonas spp.)
[83], Acinetobacter spp [84]. have been released. Cserh ati
et al. (2013) have reported that Rhodococcus. Erythropolis,
Rhodococcus ruber, and Rhodococcus pyridinivorans had
ZEA degrading capacity [85]. Tan et al., (2014) determined
Pseudomonas alcaliphila TH-C1 and Pseudomonas pleco-
glossicida TH-L1 had degradation rate of ZEA about 68%
and 57%, respectively [86]. Altalhi (2007) [87] described
Pseudomonas putida ZEA-1 isolated from rhizosphere of
corn plant, can utilize ZEA as energy source and transform
ZEA along with a and b zearalenol, without generating harm-
ful metabolites. Later, Altalhi and El-Deeb (2009) [88] local-
ized and cloned the fragment containing genes encoding for
ZEA detoxification in the plasmid of P. Putida ZEA-1, and
the cloned genes were actively expressed in E. coli. The re-
sults showed that ZEA degradation by recombinant E. coli
was relatively rapid and effective, leaving no detectable
ZEA after 24 hours. The efficient abilities of B. lichenifor-
mis CK1 and B. subtilis to degrade ZEA in the culture media
were also reported [33, 89]. To date, several microorganisms
from various sources such as soils, animal guts, and plants
have been reported to have the ability to degrade DON [90].
4. BIOLOGICAL DETOXIFICATION BY FUNGI
Miazzo et al. removed mycotoxins via structural modifi-
cations in the mycotoxin molecule without the generation of
any toxic by-products or metabolites [91]. This is generally
implemented by using some yeasts, bacteria, or fungal en-
zymes which decompose mycotoxins into non-toxic com-
pounds [92]. However, because yeast is particularly suitable
for post-harvest use, has fast growth and has been proven in
decreasing the incidence of fungal pathogens, the applica-
tion of yeasts has great potential in reducing the economic
damage caused by toxigenic fungi in the agriculture [93,
94]. Furthermore, it has been reported that by binding the
mycotoxins in the gastrointestinal tract, yeast cell wall ex-
tract, containing polysaccharides and protein, was able to de-
crease adverse effects of mycotoxins in chickens, cows, and
pigs [95-98]. Some experimental studies indicated that the
isolates of phyllosphere yeast have the capability of con-
trolling sour rot that is caused by Aspergillus carbonarius in
wine-producing vineyards [99]. Also, Cryptococcus spp. is
very active against Fusarium spp. head blight (up to 59% re-
duction) ondurum wheat in the field [[100]].
4.1. Aflatoxin (AF)
The role of fungal species in the degradation of AF has
been investigated in several research in recent years [101].
This finding is based on microbial processes involved in the
degradation of complex organic aromatic compounds such
as lignin. Regarding polyphenolic compounds in nature,
lignin is undeniably the most plentiful and probably the
most heterogeneous and recalcitrant compound to be degrad-
ed microbially [102]. The degradation of AFB1 by white rot
fungi could be a crucial bio-control measure to reduce the
level of this mycotoxin in food commodities. As investigat-
ed in the Salmonella typhimurium mutagenicity assay, degra-
dation of AFB1 by laccase enzyme from T. versicolor and re-
combinant laccase enzyme produced by A. niger D15-Lc-
c2#3 and production of recombinant laccase enzyme by A.
niger D15-Lcc2# resulted in loss of mutagenicity of AFB1
[103]. Moreover, it was found that a live yeast, S. cerevisiae
could reduce the negative effects of aflatoxicosis in poultry
[104]. Later, Huynh et al. employ (NH4)2SO4 at 80% to
100% precipitation to extract a crude mycelial protein from
a 16-day-old culture of A. parasiticus in which AFB1could
be detoxified. In their study, the main breakdown product of
AFB1 was isolated and confirmed to be non-fluorescent,
non-mutagenic, and non-toxic for ducklings. The result
showed that the dialyzedmycelial protein destructed AF by
degradation of the cyclopentenone moiety, principally the
lactone ring, via infrared spectralanalysis of the main break-
down product [105]. To avoid aflatoxin contamination in
feed, Kusumaningtyas et al. introduced S. cerevisiae and Rhi-
zopus oligosporus. In this approach, A. flavus is cultured to-
gether with S. cerevisiae and Rhizopus oligosporus and their
combination (SCRO) in chicken feed and the content of AF-
B1 is observed at day 0, 5, 10, and 15. The results of this
study indicated that AFB1 contaminations in feed were re-
duced by SC, RO and SCRO addition. The highest and best
decline of AFB1 content was at day five by Ro [106]. Based
on Shetty et al. study, even at the highest concentration test-
ed (20 µgmL -1), S. cerevisiae cells were able to bind high
amounts of AFB1. They also indicated that the ability of S.
cerevisiae to bind AFB1 was strain-specific with seven
strains binding 10%–20%, eight strains binding 20–40% and
three strains binding more than 40% of the added AFB [18].
Recently, Zhang et al. reported that screened strain of A.
niger (ND-1) could degrade 58.2% of AFB1 after 48 h of fer-
mentation and this degradation was meaningfully powerful
in culture supernatant compare to cells and cell extracts
[107].
4.2. Zearalenone
Transformation of ZEA by fungi and actinomycetes was
initially introduced by El-Sharkawy and Abul-Hajj in 1987,
in which ZEA-4-O-b-glucoside substance was characterized
[108]. In another study, they suggested that the most impor-
tant detoxification was the lactone bond cleavage of ZEA by
the mycoparasite Gliocladium roseum [109]. El-Sharkawy et
al. reported that screening with 150 fungal species showed
that ZEA was transformed by many microbes leading to a re-
duction of the C-6’ carbonyl, saturation of the 1’-,2’- double
bond and hydroxylation at 3’ and 8’. They also showed that
one microbe, namely Gliocladium roseum NRRL 1859, was
capable of metabolizing ZEA in 80%–90% yields. The prod-
uct consisting of a mixture of two isomeric hydroxyl ketones
and far less oestrogenic compare to ZEA, rendered the reac-
tion irreversible [110]. Molnar et al. introduced a new yeast
strain, Trichosporon mycotoxinivorans that was capable in
ZEA degrading to carbon oxide, and other non-toxic metabo-
lites, neither a- nor b-ZEA were detected. They indicated
that Trichosporon mycotoxinivorans was useful in detoxifica-
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 9
tion of the respective mycotoxins in animal feeds [111]. Jard
et al. showed the ability of A. niger to transform ZEA over a
broad range of ZEA concentrations (5 to 150 μg/mL) and
that sulfonation could lead to a less toxic compound [112].
Also, Vekiru et al. indicated that ZOM-1, a new ZEA
metabolite, was characterized by an opening of the macro-
cyclic ring of ZEA at the ketone group at C60. This new
ZEA metabolite did not interact in vitro with the estrogen re-
ceptor protein and was not estrogenic in vivo, indicating the
loss of ZOM-1 (novel ZEA metabolite) estrogenicity. They
also showed that ZEA was found to be entirely degraded by
several Rhizopus including R. stolonifer, R. oryzae and R. mi-
crosporus [113]. Moreover, one another study showed that
Rhizopus arrhizus catalyzes sulfation of ZEA at the C-4 hy-
droxyl group [114]. Based on the study of Utermark J et al.,
the Gliocladium roseum (mycoparasitic fungus) produces a
ZEA -specific lactonase which catalyzes the hydrolysis of
ZEA, followed by a spontaneous decarboxylation. The
growth of G. roseum was not inhibited by ZEA, and the lac-
tonase may protect G. roseum from the toxic effects of this
mycotoxin [115]. Using strains of S. cerevisiae, G. fermen-
tans, K. marxianus, and M. pulcherrima yeasts, J.
Repecˇkiene et al. eliminated ZEA from wheat flour and fod-
der [116].
4.3. Ochratoxins
Based on some research, reduction of OTA production
of Penicillium verrucosum can be caught by several S. cere-
visiae and also some non-Saccharomyces yeasts such as
Hanseniaspora spp., Trichosporon spp., Rhodotorula spp.,
and Cryptococcus spp., Trichosporon spp., Rhodotorula
spp., and Cryptococcus spp. were able to split the amide
bond of the OTA molecule and release non-toxin ochratoxin
an (OTα) [117, 118]. In addition to the effects of Pichia ano-
mala, Pichia kluyveri and Hanseniaspora uvarum predomi-
nant during coffee processing on the growth of Aspergillus
ochraceus and production of OTA on malt extract agar
(MEA) and coffee agar (CA) were studied by Utermark J et
al. [115] when three yeasts co-cultured in MEA and CA,
they showed capability to inhibit growth of A. ochraceus.
Moreover, Molnar et al. also isolated, characterized, and re-
searched the ability of T. mycotoxinivorans, to detoxify
OTA in a mineral solution [111]. Ochratoxin is detoxified
by the cleavage of the phenylalanine moiety to form the deri-
vate ochratoxin-A, a virtually nontoxic metabolite compared
to the parent compound [111]. Velmourougane et al., in
their study, employed yeast suspension (S. cerevisiae) in cof-
fee post-harvest. In result, total mould incidence and ochra-
toxin-A contamination were mitigated meaningfully without
any effects on cup quality. Also, the use of yeast culture in
coffee processing was found to be an affordable and cost-ef-
fective approach in the management of A. ochraceus and
OTA in parchment and cherry coffee preparation [119].
Based on the study of Bejaoui et al., S. cerevisiae and
Saccharomyces bayanus. showed OTA absorption ability in
synthetic grape juice [120]. There is some evidence on the
destroyable effects of some baker's yeast (S. cerevisiae)
strains on the growth of Aspergillus carbonarius and Fusari-
um graminearum effectively. Based on another study, total
mold incidence (A. ochraceus and A. niger) and contamina-
tion of OTA on a coffee can be decreased by S. cerevisia
[119]. Furthermore, the growth rate of ochratoxin-producing
Aspergillus spp. and the amount of the accumulated OTA in
vitro and on detached grape berries can be reduced by
Lachancea thermotolerans (formerly Kluyveromyces)
[121]. One investigation about biocontrol activity of the na-
tive yeast flora indicated that OTA biosynthesis could be in-
hibited by Debaryomyces spp., Candida spp., and Hy-
phopichia spp. Based on this study, the indigenous yeasts
had an antagonistic effect on the Penicillium nordicumas.
well growth. Also, the growth and biosynthesis of OTA can
be effectively reduced by H. burtonii and C. zeylanoides
[122]. In another study, detoxification capabilities of S. cere-
visiae, Geotrichum fermentans, Kluyveromyces marxianus,
and Metschnikowia pulcherrima have been tested on wheat
flour and composite fodder contaminated with different my-
cotoxins, and the DON content was significantly reduced in
the samples with greater effectiveness in the fodder [116].
Based on the study of Caridi et al., wine yeast (S. sensus-
tricto) could remove OTA from the synthetic medium. They
resulted that seven out of 20 yeast strains removed high lev-
els of OTA, ranging from 66% to 100% [123].
4.4. Deoxynivalenol
Styriak et al. indicated that after incubation with yeast
culture, the concentration of DON reduced up to 37% [124].
Also, Paskevicius et al. reported that in fodder inoculated
with several yeast genera resulted in the loss of DON [125].
P. Karlovsky et al. supposed that these results might have
been caused by adsorption [126]. Chenghua He et al. [127]
used the biotransformation system to decrease DON toxici-
ty. They isolated a Aspergillus spp. (NJA-1) from soil and
cultured it in an inorganic salt medium containing DON. An
unknown factor in NJA-1 caused DON toconvertinto
another product. Recent investigations indicated that NJA-1
could hydrolyze DON [127].
4.5. Fumonisins
There are not reliable and available documents about the
effects of yeast and fungal strains on fumonisins degrada-
tion. However, the ability of 12 S. cerevisiae strains has
been tested for the degradation of fumonisins and
zearalenone in Sabouraud broth. In the result, one strain de-
creased the concentration of mycotoxin up to 25% and one
strain up to 75% of the original amount. On the contrary, the
effects of two strains on fumonisins degradation were partial
[124].
4.6. Patulin
The investigation of yeast effects on PAT degradation
has been developed properly. For example, Stinson et al. in-
dicated that the PAT content in contaminated apple juice
with yeast reduced during alcoholic fermentation [128].
Moss et al. [129] found that three commercial cider strains
of S. cerevisiae were able to degrade PAT, but this was
achieved during active fermentative growth, but not when
10 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
growing aerobically. The polarity of the PAT degradation
products was more than PAT itself and remained in the clari-
fied fermented cider. Richelli et al. [130] showed that ac-
cording to the change of chemical structure and the degrada-
tion product (ascladiol), the capability of Gluconobacter oxy-
dans to PAT degradation was more than 96% after twelve-
hour treatment. In another study and in order to ferment ap-
ple juice including 15 mg of patulin/L, eight yeast strains
were used in three typical American processes. In the result,
except for two cases, patulin was decreased to less than the
minimum detectable level of 50 µg/L. In all cases and dur-
ing alcoholic fermentation, the level of PAT was diminished
by over 99% [131]. The culture supernatant of P. ohmeri
158 was able to inhibit mycelial growth (66.17%). The ini-
tial PAT concentration of 223 µg in the presence of P. oh-
meri 158 cells was decreased over 83% of the original con-
centration, when incubated at 25ºC for 2 days and > 99% af-
ter 5 days incubation time, with undetectable PAT level af-
ter 15 days. The results of this study showed that P. ohmeri
158 could be a suitable and efficient alternative for the inhi-
bition of P. expansum growth and PAT degradation [132].
[137] According to the study of Ianiri et al. [133] Basi-
diomycete yeasts, which have shown capability for degrad-
ing PAT effectively, could be explored by various methods
to realize this degradation process.
Based on this study, Sporobolomyces strain IAM 13481
was able to degrade PAT (which produced by the mold) to
form two various breakdown products, desoxypatulinic acid,
and (Z)-ascladiol. Compare to PAT; these products have less
toxicity. Based on their study, pre-incubated cells in a low
concentration of PAT had stronger resistance to PAT toxici-
ty and faster kinetics of degradation. When extracts of
Sporobolomyces spp. were prepared from the cells pre-treat-
ed with mycotoxin, the degradation of PAT was faster [133].
Moreover, the yeast species that showed their effects on my-
cotoxins, there are some other yeast species containing
Sporobolomyces roseus, P. ohmeri, M. pulcherrima, and R.
kratochvilovae analysed for their ability to degrade PAT
[132-135].
5. MECHANISMS INVOLVED IN DETOXIFICATION
Biological detoxification of mycotoxins occurs mainly
via two primary mechanisms, enzymatic and degradation
sorption. Both of these mechanisms can be performed by mi-
croorganisms.
Living microorganisms can absorb mycotoxins by attach-
ing or actively internalizing them. Enzymatic degradation
can be occurred by either extra or intracellular enzymes. Dif-
ferent chemical structures of mycotoxin groups make it im-
possible to equally deactivate all mycotoxins using only one
single strategy [136].
5.1. Degradation
5.1.1. Fumonisins
Degradation of the fumonisins mostly occurs through the
removal of TCA groups and a free amino group. It has been
observed that Sphingomonas spp. ATCC 55552 can degrade
fumonisin B1 through the action of an aminotransferase
[137] and a carboxylesterase [138]. Also, in Sphingopyxis
spp. MTA 144, a similar pathway was found [139].
Exophiala spp. can degrade the fumonisins through the ac-
tion of a carboxylesterase and an oxidative deaminase. Two
products that result from fumonisin degradation include a
new compound, 2-oxo-12,16-dimethyl-3,5,10,14,15- icosane-
pentol hemiketal, and in fewer amounts the N-acetylated
aminopentol backbone (N-acetylAP1). Strain NCB 1492,
gave rise to four tentative degradation products of fumonisin
B1: heptadecanone (C17H34O), isononadecene (C19H38),
octadecenal (C18H34O) and eicosane (C20H42) [140]. It
seems that the first steps of degradation occur extracellular-
ly, with deaminase (and possibly esterase) activity after a
slower degradation of the aliphatic chain. Due to structural
similarity of Alternaria toxins AAL-T to Fumonisins, it is
possible that fumonisin degrading organisms also be able to
degrade AAL-toxin [141].
5.1.2. Zearalenone
Recently, two significant mechanisms of Zearalenone
(ZEN) inactivation were discovered. Both mechanisms of ac-
tion inactivate through the cleavage in a ring structure.
The first mechanism is the cleavage of the lactone ring,
which is the one that appears in some fungal species. In the
case of Gliocladiumroseum NRRL1859, two degradation
products of ZEN toxins were obtained: by obtained 1-(3,5-di-
hydroxyphenyl)-10′-hydroxy-1-undecen- 6′-one and 1-(3,5-
dihydroxyphenyl)-6′-hydroxy-1-undecen-10′-one [110].
Matthies et al. showed that derivatives zearalanol and α-
zearalanol induced production of the ZEN-degrading en-
zyme in G. roseum DSM 62726 [142]. A near isogenic
strain, Clonostachys rosea degrade ZEN through the action
of a ZEN lactonohydrolase enzyme (zhd101) which cata-
lyzes the hydrolysis of ZEN at the ester bond in the lactone
ring, followed by spontaneous decarboxylation. Only the
first metabolite (1-(3,5-dihydroxyphenyl)-10′-hydroxy-1-un-
decen- 6′-one) was observed after degradation [143]. Based
on this knowledge, Popiel et al. studied on a collection of
Trichoderma spp., and Clonostachys spp., to find a function-
al ZEN lactonohydrolase in mycoparasitic Trichoderma ag-
gressivum [144]. A similar process might also exist in Bacil-
lus spp. such as B. natto CICC 24640 and B. subtilis 168
that cause complete degradation of ZEN [145]. A second
cleavage pathway is shown by the yeast Trichosporum myco-
toxinivorans to ZOM-1 intermediate (cleavage at the C6-ke-
tone group) [111], suggested taking place through a lactone
intermediate and subsequent activity by unspecified a/b-hy-
drolase. Similar pathway (but without the decarboxylation)
was seen in C. rosea strain FS10 convert ZEN into two inter-
mediate products, ZEN-A, and ZEN-B [146]. Similarly,
Acinetobacter spp. SM04 produces two degradation prod-
ucts (ZEN-1 and ZEN-2) from ZEN [147]. Pseudomonas
spp. ZEA-1 was shown to harbor the responsible genes for
degradation of ZEN and its derivatives on a 120 kb plasmid.
Rhodococcus spp., as a metabolically highly versatile genus,
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 11
showed considerable potential for biodegradation and to the
complete loss of ZEN estrogenic activity and also simultane-
ously degradation AFB1, ZEN, and T2-toxin [85, 148].
5.1.3. Trichothecenes
Trichothecenes toxicity results from their epoxide-group
and depends on their acylated side chains. Two main groups
of trichothecenes are; acylated (e.g., T-2 toxin) and non-acy-
lated trichothecenes (e.g., DON) [141]. The first step in dis-
abling the acylated trichothecenes is de-acylation. Curtobac-
terium spp. strain 114-2 degrade T-2 toxin to HT-2 toxin
and subsequently to T-2 triol, which is less toxic [149]. The
next step of trichothecenes detoxification is de-epoxidation.
Eubacterium spp. BBSH 797 can degrade several tri-
chothecenes [150, 151]. This strain has been developed into
a commercial product (Biomin BBSH 797) for detoxifying
trichothecenes in animal feed [152]. It is known for its abili-
ty to detoxify DON into deepoxy-deoxynivalenol (DOM-1)
and de-epoxidization of NIV, T-2 tetraol, scirpentriol, and
HT-2 toxin. Further, 4-acetyl NIV and 3-acetyl NIV was de-
acetylated and/or de-epoxidized [150]. Degradation of DON
occurs through de-epoxidation, oxidation, or isomerization.
Microbial culture C133 of fish guts transformed DON to
DOM-1 [153]. Citrobacter freundii could transform DON in-
to DOM-1 aerobically [154]. DON can also be oxidized to
3-keto DON which is 10 times less toxic than DON evaluat-
ed with a bioassay based on mitogen-induced and mito-
gen-free proliferation of mouse spleen lymphocytes [155].
The bacterium strain E3-39 which degraded DON to 3-keto--
DON, is belonging to the Agrobacterium-Rhizobium spp.
group. A mixed culture from environmental sources could
degrade DON into 3-keto-DON, whereas 15-acetyl DON, 3-
acetyl DON, and fusarenon-X were also transformed [156].
The metabolite 3-epi-DON was also formed by degradation
of DON through Nocardioides spp. strain WSN05-2 [157].
Moreover, lastly, nine Nocardioides spp. (Gram-positive)
and four Devosia spp. (Gram-negative) produced 3-epi--
DON aerobically. The Gram-positive strains showed DON
assimilation, whereas the Gram-negatives did not showed
[158].
5.1.4. Aflatoxins
For the first time, the biological detoxification of AFB1
mediated by Flavobacterium aurantiacum (now called No-
cardia corynebacterioides) has been reported in 1966 [79,
159].
After this first report, while many studies focused on the
detoxification of AFB1, only a few studies detected the
degradation products and analyzed their toxicity. There are
two major pathways of detoxification of aflatoxins which in-
clude modification of the difuran ring and the coumarin
structure. Several studies reported modification of the difu-
ran ring moiety. Degradation of AFB1 into AFB1-8,9- dihy-
drodiol was performed by manganese peroxidase from the
white rot fungi Phanerochaete sordida [160] and the “afla-
toxin-detoxifizyme (ADTZ)” of fungus Armillariella tabes-
cens [161]. The authors suggested that AFB1 degradation ini-
tially involves the formation of AFB1-8,9-epoxide, after
which hydrolysis resulted in a dihydrodiol-derivate. Another
metabolite was detected with the white rot fungus Pleurotus
ostreatus GHBBF10 which degraded 91.76% of AFB1 into
a component which could be a hydrolyte of AFB1, namely
dihydro hydroxy aflatoxin B1 (AFB2a) [162]. Many studies
have shown a modification of the lactone ring in the cou-
marin moiety of AFB1. In a study, a Pseudomonas putida
strain has shown the ability of degradation of AFB1 into
AFD1 and subsequently into AFD2 that have lower muta-
genicity and toxicity [163, 164]. In some studies, no degrada-
tion product was identified, but toxicity tests were per-
formed on the treated AFB1. Similarly to F. aurantiacum, as
mentioned before, a pure laccase enzyme from Trametes ver-
sicolor and a recombinant laccase enzyme produced by A.
niger degraded AFB1 with a significant loss of toxin muta-
genicity [103]. Also, in Rhodococcus eryhtropolis extracellu-
lar enzymes were able to degrade AFB1 with a loss of muta-
genicity [78].
5.1.5. Ochratoxins
The primary detoxification pathway of OTA is the hydro-
lyzation of the amide bond between the isocoumarin residue
and phenylalanine by a carboxypeptidase. Two classes of
carboxypeptidases have been associated with degradation of
OTA which include carboxypeptidase A (CPA) [103] and
carboxypeptidase Y (CPY) [165]. Peptidase enzyme has an
essential role in the efficient degradation of OTA. Other en-
zymes are also able to carry out this reaction: Dioxygenases,
lipases, amidases, and several commercial proteases [166].
Although depending on the enzyme, intermediates may be
different, but the end product is always OTα. Trichosporon
mycotoxinivorans was able to convert OTA into the nontox-
ic OTα. Also, Phenylobacterium immobile can convert OTA
to ochratoxin α (OTα) through a dioxygenase step on the
phenylalanine moiety, a dehydrogenation to catechol, a ring
cleavage and the final formation of OTα via ahydrolase
[141].
5.2. Adsorption
The most common technique used for reducing exposure
to mycotoxins is the use of mycotoxin-binding agents or ad-
sorbents, which reduce mycotoxin uptake. Lactic acid bacte-
ria and Bifidobacterium species are the two most important
groups of microorganisms that have been much studied.
5.2.1. Bifidobacterium Species
Peltonen et al. studied the AFB1 binding ability of five
Bifidobacterium strains in phosphate buffered saline (PBS),
and they noticed that Bifidobacterium spp. bound
18%–48.7% of AFB1 [25]. Recently, Fuchs et al. investigat-
ed the detoxification of two abundant mycotoxins, namely
OTA and PAT, by Bifidobacterium and they found that two
Bifidobacterium longum (LA 02, VM 14) strains were high-
ly effective and caused a decrease of OTA by approximately
50%, whereas with PAT, the most potent effect (ca. 80% de-
crease) was seen with a B. animalis (VM 12) strain [45]. Re-
cently, Hatab et al. reported that the maximum PAT uptake
12 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
was achieved by B. bifidum 671 by (52.9%) for viable and
(54.1%) for nonviable cells after 24-h incubation [167]. On
studying the ability of two dairy strains of B. bifidum to re-
move AFM1 from PBS and reconstituted milk, results re-
vealed that viable (108 CFU mL−1) and heat-killed Bifi-
dobacterium spp. ranged from 14.04% to 28.07%, and from
12.85 to 27.31% in PBS and reconstituted milk, respectively
[168]. Previous studies also revealed that the binding abili-
ties of AFM1 by B. longum and B. bifidum reached 26.7 and
32.5%, respectively [169].
5.2.2. Lactic Acid Bacteria
5.2.2.1. Aflatoxins
Specific dairy species of LAB such as Lactobacillus and
Lactococcus have been shown to bind and remove AFB1,
the most common aflatoxin [170, 171]. The toxin-removing
capacity of a combination of strains of LAB is not the sum
of their capacities. Thus suggesting that pure, single strains
should be used when the goal is to remove single com-
pounds and that the use of combinations of strains may be
beneficial when several compounds are removed together
[172].
The binding abilities of AFM1 by viable Lactobacillus
spp. at 108 CFU mL−1 in PBS ranged from 10.22 to 26.55%
depending on the contamination level and incubation period
[168]. They added that the percentage of AFM1 removal by
heat-killed bacteria from PBS ranged from 14.04 to 28.97%.
Similarly, Kabak and Var found that the binding abilities of
AFM1, by Lactobacillus spp. in PBS and reconstituted milk,
ranged from 25.7 to 30.5% [169]. These results were consid-
ered to be lower than those reported by Pierides et al., who
found that the AFM1 binding abilities of viable Lactobacil-
lus spp. within 15–16 h ranged from 18.1 to 50.7% [173].
5.2.2.2. Ochratoxin A
Degradation of OTA was observed in milk due to the ac-
tion of Lactobacillus and Streptococcus spp [174]. Lactoba-
cillus spp. such as L. rhamnosus spp., L. bulgaricus spp., L.
helveticus spp., and L. acidophilus were efficiently able to re-
move OTA [51]. Wine LAB (Lactobacillus spp., Pediococ-
cus spp., and Oenococcus oeni spp.) caused a decrease in
OTA concentration [57].
5.2.2.3. Patulin
The effect of LAB on PAT, was shown by the decline of
PAT (39%) by Lactobacillus plantarum [45]. Hatab et al. it
has been reported that the high PAT uptake from aqueous so-
lution was achieved by Lactobacillus rhamnosus 6149 strain
[167]. At pH 4.0 and 37 ° C, the highest removal of PAT
was seen Lactobacillus rhamnosus caused 80.4% decrease
of PAT in apple juice [47].
5.2.2.4. Fusarium Toxins
Certain LAB strains have been the detoxifying effect on
other mycotoxins such as fusarium toxins (T-2 toxin, DON,
NIV, HT-2 toxin) [51]. Binding is the mode of action of re-
moval of Fusarium mycotoxins by fermentative bacteria
[39]. The consistency of the mycotoxin-LAB interaction
was due to the amino acid composition of the peptidoglycan
structure [175]. The ability of Lactobacillus rhamnosus
strains to remove ZEN, and its derivative α-zearalenol, from
a liquid medium, was reported by El-Nezami et al. [176].
Haskard et al. studied the mechanism of binding between
AFs and Lactobacillus rhamnosus and found that binding oc-
curs predominantly with the carbohydrate and to some ex-
tent, the protein components in the cell wall. These results
were based on the inhibitory effects of periodate and pro-
nase E on AFB1 binding by Lactobacillus rhamnosus Perio-
date and pronase E can be used to degrade relatively non-
specifically both carbohydrates and proteins, respectively.
They also described the primary role of hydrophobic interac-
tions in binding and found that electrostatic interactions play
only a minor role [69].
CONCLUSION
Microbiological decontamination of mycotoxins has
been considered as a main and effective approach for myco-
toxin biodegradation. Regardless of the method used, proper
management tools that consist of before-harvest prevention
and after-harvest detoxification, are required. However, the
use of this approach depends on its effectiveness from both
a practical and economic perspective.
CONSENT FOR PUBLICATION
Not applicable.
FUNDING
None.
CONFLICT OF INTEREST
The authors have no conflicts of interest, financial or
otherwise.
ACKNOWLEDGEMENTS
We are grateful for the useful comments and suggestions
from anonymous referees.
REFERENCES
Didwania, N.; Trivedi, P. Mycotoxins: A review of toxicity,[1]
metabolism and biological approaches to counteract the produc-
tion in food. MR International Journal of Engineering & Tech-
nology, 2018, 6(2), 38-42.
Zain, M.E. Impact of mycotoxins on humans and animals. J. Sau-[2]
di Chem. Soc., 2011, 15(2), 129-144.
[http://dx.doi.org/10.1016/j.jscs.2010.06.006]
Becker-Algeri, T.A.; Castagnaro, D.; de Bortoli, K.; de Souza, C.;[3]
Drunkler, D.A.; Badiale-Furlong, E. Mycotoxins in bovine milk
and dairy products: a review. J. Food Sci., 2016, 81(3), R544-
R552.
[http://dx.doi.org/10.1111/1750-3841.13204] [PMID: 26799355]
Jarvis, B.B. Chemistry and toxicology of molds isolated from wa-[4]
ter-damaged buildings.Mycotoxins and Food Safety; Springer,
2002, pp. 43-52.
[http://dx.doi.org/10.1007/978-1-4615-0629-4_5]
Khatun, S.; Chakraborty, M.; Islam, A.; Cakilcioglu, U.; Chatter-[5]
jee, N.C. Mycotoxins as health hazard. Biological Diversity and
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 13
Conservation, 2012, 5(3), 123-133.
Trucksess, M.W.; Scott, P.M. Mycotoxins in botanicals and dried[6]
fruits: a review. Food Addit. Contam. Part A Chem. Anal. Control
Expo. Risk Assess., 2008, 25(2), 181-192.
[http://dx.doi.org/10.1080/02652030701567459] [PMID:
18286408]
Trias, R.; Bañeras, L.; Montesinos, E.; Badosa, E. Lactic acid bac-[7]
teria from fresh fruit and vegetables as biocontrol agents of phyto-
pathogenic bacteria and fungi. Int. Microbiol., 2008, 11(4),
231-236.
[PMID: 19204894]
Dalié, D.; Deschamps, A.; Richard-Forget, F. Lactic acid bacteri-[8]
a–Potential for control of mould growth and mycotoxins: A re-
view. Food Control, 2010, 21(4), 370-380.
[http://dx.doi.org/10.1016/j.foodcont.2009.07.011]
Vogelmann, S.A.; Seitter, M.; Singer, U.; Brandt, M.J.; Hertel, C.[9]
Adaptability of lactic acid bacteria and yeasts to sourdoughs pre-
pared from cereals, pseudocereals and cassava and use of competi-
tive strains as starters. Int. J. Food Microbiol., 2009, 130(3),
205-212.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2009.01.020] [PMID:
19239979]
Oelschlaeger, T.A. Mechanisms of probiotic actions - A review.[10]
Int. J. Med. Microbiol., 2010, 300(1), 57-62.
[http://dx.doi.org/10.1016/j.ijmm.2009.08.005] [PMID:
19783474]
Kesarcodi-Watson, A.; Kaspar, H.; Lategan, M.J.; Gibson, L. Pro-[11]
biotics in aquaculture: the need, principles and mechanisms of ac-
tion and screening processes. Aquaculture, 2008, 274(1), 1-14.
[http://dx.doi.org/10.1016/j.aquaculture.2007.11.019]
Prakash, B.; Shukla, R.; Singh, P.; Kumar, A.; Mishra, P.K.;[12]
Dubey, N.K. Efficacy of chemically characterized Piper betle L.
essential oil against fungal and aflatoxin contamination of some
edible commodities and its antioxidant activity. Int. J. Food Micro-
biol., 2010, 142(1-2), 114-119.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2010.06.011] [PMID:
20621374]
Cavaliere, C.; Foglia, P.; Pastorini, E.; Samperi, R.; Laganà, A.[13]
Liquid chromatography/tandem mass spectrometric confirmatory
method for determining aflatoxin M1 in cow milk: comparison be-
tween electrospray and atmospheric pressure photoionization
sources. J. Chromatogr. A, 2006, 1101(1-2), 69-78.
[http://dx.doi.org/10.1016/j.chroma.2005.09.060] [PMID:
16221477]
Cavaliere, C.; Foglia, P.; Guarino, C.; Marzioni, F.; Nazzari, M.;[14]
Samperi, R.; Laganà, A. Aflatoxin M1 determination in cheese by
liquid chromatography-tandem mass spectrometry. J. Chromato-
gr. A, 2006, 1135(2), 135-141.
[http://dx.doi.org/10.1016/j.chroma.2006.07.048] [PMID:
17056052]
Zinedine, A.; Brera, C.; Elakhdari, S.; Catano, C.; Debegnach, F.;[15]
Angelini, S. Natural occurrence of mycotoxins in cereals and
spices commercialized in Morocco. Food Control, 2006, 17(11),
868-874.
[http://dx.doi.org/10.1016/j.foodcont.2005.06.001]
Santini, A.; Ritieni, A. Aflatoxins: risk, exposure and remediation.[16]
In: Aflatoxins-Recent Advances and Future Prospects: ; InTech,
2013.
[http://dx.doi.org/10.5772/52866]
Shahin, A. Removal of aflatoxin B1 from contaminated liquid me-[17]
dia by dairy lactic acid bacteria. Int. J. Agric. Biol., 2007, 9(1),
71-75.
Shetty, P.H.; Hald, B.; Jespersen, L. Surface binding of aflatoxin[18]
B1 by Saccharomyces cerevisiae strains with potential decontami-
nating abilities in indigenous fermented foods. Int. J. Food Micro-
biol., 2007, 113(1), 41-46.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2006.07.013] [PMID:
16996157]
Shetty, P.H.; Jespersen, L. Saccharomyces cerevisiae and lactic[19]
acid bacteria as potential mycotoxin decontaminating agents.
Trends Food Sci. Technol., 2006, 17(2), 48-55.
[http://dx.doi.org/10.1016/j.tifs.2005.10.004]
Bovo, F.; Corassin, C.H.; Rosim, R.E.; de Oliveira, C.A. Efficien-[20]
cy of lactic acid bacteria strains for decontamination of aflatoxin
M 1 in phosphate buffer saline solution and in skimmed milk.
Food Bioprocess Technol., 2013, 6(8), 2230-2234.
[http://dx.doi.org/10.1007/s11947-011-0770-9]
El-Nezami, H.; Kankaanpaa, P.; Salminen, S.; Ahokas, J. Ability[21]
of dairy strains of lactic acid bacteria to bind a common food car-
cinogen, aflatoxin B1. Food Chem. Toxicol., 1998, 36(4),
321-326.
[http://dx.doi.org/10.1016/S0278-6915(97)00160-9] [PMID:
9651049]
Khanafari, A.; Soudi, H.; Miraboulfathi, M. Biocontrol of As-[22]
pergillus flavus and aflatoxin B1 production in corn. J. Environ.
Health Sci. Eng., 2007, 4(3), 163-168.
Hernandez-Mendoza, A.; Garcia, H.S.; Steele, J.L. Screening of[23]
Lactobacillus casei strains for their ability to bind aflatoxin B1.
Food Chem. Toxicol., 2009, 47(6), 1064-1068.
[http://dx.doi.org/10.1016/j.fct.2009.01.042] [PMID: 19425181]
Shah, N.; Wu, X. Aflatoxin B1 binding abilities of probiotic bacte-[24]
ria. Biosci. Microflora, 1999, 18(1), 43-48.
[http://dx.doi.org/10.12938/bifidus1996.18.43]
Peltonen, K.; el-Nezami, H.; Haskard, C.; Ahokas, J.; Salminen,[25]
S. Aflatoxin B1 binding by dairy strains of lactic acid bacteria and
bifidobacteria. J. Dairy Sci., 2001, 84(10), 2152-2156.
[http://dx.doi.org/10.3168/jds.S0022-0302(01)74660-7] [PMID:
11699445]
Sarimehmetoğlu, B.; Küplülü, Ö. Binding ability of aflatoxin M1[26]
to yoghurt bacteria. Ankara Univ. Vet. Fak. Derg., 2004, 51(3),
195-198.
Peltonen, K.D.; El‐Nezami, H.S.; Salminen, S.J.; Ahokas, J.T.[27]
Binding of aflatoxin B1 by probiotic bacteria. J. Sci. Food Agric.,
2000, 80(13), 1942-1945.
[http://dx.doi.org/10.1002/1097-0010(200010)80:13<1942::AID-J
SFA741>3.0.CO;2-7]
Ghazvini, R.D.; Kouhsari, E.; Zibafar, E.; Hashemi, S.J.; Amini,[28]
A.; Niknejad, F. Antifungal activity and aflatoxin degradation of
bifidobacterium bifidum and lactobacillus fermentum against toxi-
genic aspergillus parasiticus. Open Microbiol. J., 2016, 10,
197-201.
[http://dx.doi.org/10.2174/1874285801610010197] [PMID:
28077976]
Schnürer, J.; Magnusson, J. Antifungal lactic acid bacteria as biop-[29]
reservatives. Trends Food Sci. Technol., 2005, 16(1-3), 70-78.
[http://dx.doi.org/10.1016/j.tifs.2004.02.014]
Urry, W.; Wehrmeister, H.; Hodge, E.; Hidy, P. The structure of[30]
zearalenone. Tetrahedron Lett., 1966, 7(27), 3109-3114.
[http://dx.doi.org/10.1016/S0040-4039(01)99923-X]
Malekinejad, H.; Maas-Bakker, R.F.; Fink-Gremmels, J. Bioactiva-[31]
tion of zearalenone by porcine hepatic biotransformation. Vet.
Res., 2005, 36(5-6), 799-810.
[http://dx.doi.org/10.1051/vetres:2005034] [PMID: 16120254]
Shier, W.T.; Shier, A.C.; Xie, W.; Mirocha, C.J. Structure-activity[32]
relationships for human estrogenic activity in zearalenone myco-
toxins. Toxicon, 2001, 39(9), 1435-1438.
[http://dx.doi.org/10.1016/S0041-0101(00)00259-2] [PMID:
11384734]
Cho, K.J.; Kang, J.S.; Cho, W.T.; Lee, C.H.; Ha, J.K.; Song, K.B.[33]
in vitro degradation of zearalenone by Bacillus subtilis. Biotech-
nol. Lett., 2010, 32(12), 1921-1924.
[http://dx.doi.org/10.1007/s10529-010-0373-y] [PMID:
20697929]
Lioi, M.B.; Santoro, A.; Barbieri, R.; Salzano, S.; Ursini, M.V.[34]
Ochratoxin A and zearalenone: a comparative study on genotoxic
effects and cell death induced in bovine lymphocytes. Mutat. Res.,
2004, 557(1), 19-27.
[http://dx.doi.org/10.1016/j.mrgentox.2003.09.009] [PMID:
14706515]
Abid-Essefi, S.; Ouanes, Z.; Hassen, W.; Baudrimont, I.; Creppy,[35]
E.; Bacha, H. Cytotoxicity, inhibition of DNA and protein synthes-
es and oxidative damage in cultured cells exposed to zearalenone.
Toxicol. in vitro, 2004, 18(4), 467-474.
[http://dx.doi.org/10.1016/j.tiv.2003.12.011] [PMID: 15130604]
Niderkorn, V.; Morgavi, D.P.; Pujos, E.; Tissandier, A.; Boudra,[36]
H. Screening of fermentative bacteria for their ability to bind and
14 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
biotransform deoxynivalenol, zearalenone and fumonisins in an in
vitro simulated corn silage model. Food Addit. Contam., 2007,
24(4), 406-415.
[http://dx.doi.org/10.1080/02652030601101110] [PMID:
17454114]
Čvek, D; Markov, K; Frece, J; Friganović, M; Duraković, L; De-[37]
laš, F Adhesion of zearalenone to the surface of lactic acid bacte-
ria cells. Hrvatski časopis za prehrambenu tehnologiju, bioteh-
nologiju i nutricionizam , 2012, 7(SPECIAL ISSUE-7th), 49-52.
Król, A.; Pomastowski, P.; Rafińska, K.; Railean-Plugaru, V.;[38]
Walczak, J.; Buszewski, B. Microbiology neutralization of
zearalenone using Lactococcus lactis and Bifidobacterium sp.
Anal. Bioanal. Chem., 2018, 410(3), 943-952.
[http://dx.doi.org/10.1007/s00216-017-0555-8] [PMID:
28852794]
Niderkorn, V.; Boudra, H.; Morgavi, D.P. Binding of Fusarium[39]
mycotoxins by fermentative bacteria in vitro. J. Appl. Microbiol.,
2006, 101(4), 849-856.
[http://dx.doi.org/10.1111/j.1365-2672.2006.02958.x] [PMID:
16968296]
Corsetti, A.; Gobbetti, M.; Rossi, J.; Damiani, P. Antimould activi-[40]
ty of sourdough lactic acid bacteria: identification of a mixture of
organic acids produced by Lactobacillus sanfrancisco CB1. Appl.
Microbiol. Biotechnol., 1998, 50(2), 253-256.
[http://dx.doi.org/10.1007/s002530051285] [PMID: 9763693]
Glaser, N.; Stopper, H. Patulin: Mechanism of genotoxicity. Food[41]
Chem. Toxicol., 2012, 50(5), 1796-1801.
[http://dx.doi.org/10.1016/j.fct.2012.02.096] [PMID: 22425938]
Speijers, G.J.; Franken, M.A.; van Leeuwen, F.X. Subacute toxici-[42]
ty study of patulin in the rat: effects on the kidney and the gas-
tro-intestinal tract. Food Chem. Toxicol., 1988, 26(1), 23-30.
[http://dx.doi.org/10.1016/0278-6915(88)90037-3] [PMID:
3345966]
Karaca, H.; Nas, S. Aflatoxins, patulin and ergosterol contents of[43]
dried figs in Turkey. Food Addit. Contam., 2006, 23(5), 502-508.
[http://dx.doi.org/10.1080/02652030600550739] [PMID:
16644598]
Speijers, G; Magan, N; Olsen, M. Patulin. Mycotoxins in food: de-[44]
tection and control , 2004, 339-352.
Fuchs, S.; Sontag, G.; Stidl, R.; Ehrlich, V.; Kundi, M.; Knas-[45]
müller, S. Detoxification of patulin and ochratoxin A, two abun-
dant mycotoxins, by lactic acid bacteria. Food Chem. Toxicol.,
2008, 46(4), 1398-1407.
[http://dx.doi.org/10.1016/j.fct.2007.10.008] [PMID: 18061329]
Ferrer, E.; Juan-García, A.; Font, G.; Ruiz, M.J. Reactive oxygen[46]
species induced by beauvericin, patulin and zearalenone in CHO-
K1 cells. Toxicol. in vitro, 2009, 23(8), 1504-1509.
[http://dx.doi.org/10.1016/j.tiv.2009.07.009] [PMID: 19596061]
Hatab, S.; Yue, T.; Mohamad, O. Removal of patulin from apple[47]
juice using inactivated lactic acid bacteria. J. Appl. Microbiol.,
2012, 112(5), 892-899.
[http://dx.doi.org/10.1111/j.1365-2672.2012.05279.x] [PMID:
22394257]
Wang, L.; Yue, T.; Yuan, Y.; Wang, Z.; Ye, M.; Cai, R. A new in-[48]
sight into the adsorption mechanism of patulin by the heat-inac-
tive lactic acid bacteria cells. Food Control, 2015, 50, 104-110.
[http://dx.doi.org/10.1016/j.foodcont.2014.08.041]
Hawar, S.; Vevers, W.; Karieb, S.; Ali, B.K.; Billington, R.; Beal,[49]
J. Biotransformation of patulin to hydroascladiol by Lactobacillus
plantarum. Food Control, 2013, 34(2), 502-508.
[http://dx.doi.org/10.1016/j.foodcont.2013.05.023]
Hateb, S.; Yue, T.; Mohaned, O. Reduction of patulin in aqueous[50]
solution using inactivated lactic acid bacteria. J. Appl. Microbiol.,
2012, 112(5), 892-899.
[PMID: 22394257]
Hathout, A.S.; Aly, S.E. Biological detoxification of mycotoxins:[51]
a review. Ann. Microbiol., 2014, 64(3), 905-919.
[http://dx.doi.org/10.1007/s13213-014-0899-7]
Battilani, P.; Giorni, P.; Bertuzzi, T.; Formenti, S.; Pietri, A.[52]
Black aspergilli and ochratoxin A in grapes in Italy. Int. J. Food
Microbiol., 2006, 111(Suppl. 1), S53-S60.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2006.03.006] [PMID:
16713645]
Piotrowska, M.; Zakowska, Z. The elimination of ochratoxin A by[53]
lactic acid bacteria strains. Pol. J. Microbiol., 2005, 54(4),
279-286.
[PMID: 16599298]
Amézqueta, S.; González-Peñas, E.; Murillo-Arbizu, M.; de[54]
Cerain, A.L. Ochratoxin A decontamination: A review. Food Con-
trol, 2009, 20(4), 326-333.
[http://dx.doi.org/10.1016/j.foodcont.2008.05.017]
Heidler, D.; Schatzmayr, G. A new approach to managing myco-[55]
toxins. Feed Mix, 2003, 11(1), 31-34.
Varga, J.; Tóth, B. Novel strategies to control mycotoxins in[56]
feeds: a review. Acta Vet. Hung., 2005, 53(2), 189-203.
[http://dx.doi.org/10.1556/AVet.53.2005.2.4] [PMID: 15959977]
Del Prete, V.; Rodriguez, H.; Carrascosa, A.V.; de las Rivas, B.;[57]
Garcia-Moruno, E.; Muñoz, R. In vitro removal of ochratoxin A
by wine lactic acid bacteria. J. Food Prot., 2007, 70(9),
2155-2160.
[http://dx.doi.org/10.4315/0362-028X-70.9.2155] [PMID:
17900096]
Lević, J.; Stanković, S.; Bočarov-Stančić, A.; Škrinjar, M.; Mašić,[58]
Z. The overview on toxigenic fungi and mycotoxins in Serbia and
Montenegro.An overview on toxigenic fungi and mycotoxins in Eu-
rope; Springer, 2004, pp. 201-218.
[http://dx.doi.org/10.1007/978-1-4020-2646-1_15]
McMullen, M.; Jones, R.; Gallenberg, D. Scab of wheat and bar-[59]
ley: a re-emerging disease of devastating impact. Plant Dis., 1997,
81(12), 1340-1348.
[http://dx.doi.org/10.1094/PDIS.1997.81.12.1340] [PMID:
30861784]
McCormick, S.P.; Stanley, A.M.; Stover, N.A.; Alexander, N.J.[60]
Trichothecenes: from simple to complex mycotoxins. Toxins
(Basel), 2011, 3(7), 802-814.
[http://dx.doi.org/10.3390/toxins3070802] [PMID: 22069741]
Pestka, J.J. Mechanisms of deoxynivalenol-induced gene expres-[61]
sion and apoptosis. Food Addit. Contam. Part A Chem. Anal. Con-
trol Expo. Risk Assess., 2008, 25(9), 1128-1140.
[http://dx.doi.org/10.1080/02652030802056626] [PMID:
19238623]
McLean, M. The phytotoxicity ofFusarium metabolites: An up-[62]
date since 1989. Mycopathologia, 1996, 133(3), 163-179.
[http://dx.doi.org/10.1007/BF02373024] [PMID: 20882471]
Kimura, M.; Tokai, T.; Takahashi-Ando, N.; Ohsato, S.; Fujimura,[63]
M. Molecular and genetic studies of fusarium trichothecene
biosynthesis: pathways, genes, and evolution. Biosci. Biotechnol.
Biochem., 2007, 71(9), 2105-2123.
[http://dx.doi.org/10.1271/bbb.70183] [PMID: 17827683]
Nathanail, A.V.; Syvähuoko, J.; Malachová, A.; Jestoi, M.; Varga,[64]
E.; Michlmayr, H.; Adam, G.; Sieviläinen, E.; Berthiller, F.; Pelto-
nen, K. Simultaneous determination of major type A and B tri-
chothecenes, zearalenone and certain modified metabolites in Fin-
nish cereal grains with a novel liquid chromatography-tandem
mass spectrometric method. Anal. Bioanal. Chem., 2015, 407(16),
4745-4755.
[http://dx.doi.org/10.1007/s00216-015-8676-4] [PMID:
25935671]
Gottschalk, C.; Barthel, J.; Engelhardt, G.; Bauer, J.; Meyer, K. Si-[65]
multaneous determination of type A, B and D trichothecenes and
their occurrence in cereals and cereal products. Food Addit. Con-
tam., 2009, 26(9), 1273-1289.
[http://dx.doi.org/10.1080/02652030903013260]
Pestka, J.J.; Smolinski, A.T. Deoxynivalenol: toxicology and po-[66]
tential effects on humans. Journal of Toxicology and Environmen-
tal Health. Part B, 2005, 8(1), 39-69.
Flannery, B.M.; Wu, W.; Pestka, J.J. Characterization of deoxyni-[67]
valenol-induced anorexia using mouse bioassay. Food Chem. Toxi-
col., 2011, 49(8), 1863-1869.
[http://dx.doi.org/10.1016/j.fct.2011.05.004] [PMID: 21575669]
El-Nezami, H.S.; Chrevatidis, A.; Auriola, S.; Salminen, S.;[68]
Mykkänen, H. Removal of common Fusarium toxins in vitro by
strains of Lactobacillus and Propionibacterium. Food Addit. Con-
tam., 2002, 19(7), 680-686.
[http://dx.doi.org/10.1080/02652030210134236] [PMID:
12113664]
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 15
Haskard, C.; Binnion, C.; Ahokas, J. Factors affecting the seques-[69]
tration of aflatoxin by Lactobacillus rhamnosus strain GG. Chem.
Biol. Interact., 2000, 128(1), 39-49.
[http://dx.doi.org/10.1016/S0009-2797(00)00186-1] [PMID:
10996299]
Richard, J.L. Some major mycotoxins and their mycotoxicos-[70]
es--an overview. Int. J. Food Microbiol., 2007, 119(1-2), 3-10.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2007.07.019] [PMID:
17719115]
Michlmayr, H.; Varga, E.; Malachova, A.; Nguyen, N.T.; Lorenz,[71]
C.; Haltrich, D.; Berthiller, F.; Adam, G. A versatile family 3 Gly-
coside Hydrolase from Bifidobacterium adolescentis hydrolyzes
β-glucosides of the Fusarium mycotoxins deoxynivalenol, ni-
valenol, and HT-2 toxin in cereal matrices. Appl. Environ. Micro-
biol., 2015, 81(15), 4885-4893.
[http://dx.doi.org/10.1128/AEM.01061-15] [PMID: 25979885]
Sobrova, P.; Adam, V.; Vasatkova, A.; Beklova, M.; Zeman, L.;[72]
Kizek, R. Deoxynivalenol and its toxicity. Interdiscip. Toxicol.,
2010, 3(3), 94-99.
[http://dx.doi.org/10.2478/v10102-010-0019-x] [PMID:
21217881]
Sangare, L.; Zhao, Y.; Folly, Y.M.E.; Chang, J.; Li, J.; Selvaraj,[73]
J.N.; Xing, F.; Zhou, L.; Wang, Y.; Liu, Y. Aflatoxin B₁ degrada-
tion by a Pseudomonas strain. Toxins (Basel), 2014, 6(10),
3028-3040.
[http://dx.doi.org/10.3390/toxins6103028] [PMID: 25341538]
Farzaneh, M.; Shi, Z-Q.; Ghassempour, A.; Sedaghat, N.; Ah-[74]
madzadeh, M.; Mirabolfathy, M. Aflatoxin B1 degradation by Ba-
cillus subtilis UTBSP1 isolated from pistachio nuts of Iran. Food
Control, 2012, 23(1), 100-106.
[http://dx.doi.org/10.1016/j.foodcont.2011.06.018]
Petchkongkaew, A.; Taillandier, P.; Gasaluck, P.; Lebrihi, A. Iso-[75]
lation of Bacillus spp. from Thai fermented soybean (Thua-nao):
screening for aflatoxin B1 and ochratoxin A detoxification. J. Ap-
pl. Microbiol., 2008, 104(5), 1495-1502.
[http://dx.doi.org/10.1111/j.1365-2672.2007.03700.x] [PMID:
18194245]
Gao, X.; Ma, Q.; Zhao, L.; Lei, Y.; Shan, Y.; Ji, C. Isolation of Ba-[76]
cillus subtilis: screening for aflatoxins B1, M1, and G1 detoxifica-
tion. Eur. Food Res. Technol., 2011, 232(6), 957.
[http://dx.doi.org/10.1007/s00217-011-1463-3]
Guan, S.; Ji, C.; Zhou, T.; Li, J.; Ma, Q.; Niu, T. Aflatoxin B(1)[77]
degradation by Stenotrophomonas maltophilia and other microbes
selected using coumarin medium. Int. J. Mol. Sci., 2008, 9(8),
1489-1503.
[http://dx.doi.org/10.3390/ijms9081489] [PMID: 19325817]
Alberts, J.F.; Engelbrecht, Y.; Steyn, P.S.; Holzapfel, W.H.; van[78]
Zyl, W.H. Biological degradation of aflatoxin B1 by Rhodococcus
erythropolis cultures. Int. J. Food Microbiol., 2006, 109(1-2),
121-126.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2006.01.019] [PMID:
16504326]
Teniola, O.D.; Addo, P.A.; Brost, I.M.; Färber, P.; Jany, K-D.; Al-[79]
berts, J.F.; van Zyl, W.H.; Steyn, P.S.; Holzapfel, W.H. Degrada-
tion of aflatoxin B(1) by cell-free extracts of Rhodococcus erythro-
polis and Mycobacterium fluoranthenivorans sp. nov.
DSM44556(T). Int. J. Food Microbiol., 2005, 105(2), 111-117.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2005.05.004] [PMID:
16061299]
Eshelli, M.; Harvey, L.; Edrada-Ebel, R.; McNeil, B.[80]
Metabolomics of the bio-degradation process of aflatoxin B1 by
actinomycetes at an initial pH of 6.0. Toxins (Basel), 2015, 7(2),
439-456.
[http://dx.doi.org/10.3390/toxins7020439] [PMID: 25658510]
Smiley, R.D.; Draughon, F.A. Preliminary evidence that degrada-[81]
tion of aflatoxin B1 by Flavobacterium aurantiacum is enzymatic.
J. Food Prot., 2000, 63(3), 415-418.
[http://dx.doi.org/10.4315/0362-028X-63.3.415] [PMID:
10716576]
Kollarczik, B.; Gareis, M.; Hanelt, M. In vitro transformation of[82]
the Fusarium mycotoxins deoxynivalenol and zearalenone by the
normal gut microflora of pigs. Nat. Toxins, 1994, 2(3), 105-110.
[http://dx.doi.org/10.1002/nt.2620020303] [PMID: 8087428]
Megharaj, M.; Garthwaite, I.; Thiele, J.H. Total biodegradation of[83]
the oestrogenic mycotoxin zearalenone by a bacterial culture. Lett.
Appl. Microbiol., 1997, 24(5), 329-333.
[http://dx.doi.org/10.1046/j.1472-765X.1997.00053.x] [PMID:
9172437]
Yu, Y.; Qiu, L.; Wu, H.; Tang, Y.; Yu, Y.; Li, X.; Liu, D. Degra-[84]
dation of zearalenone by the extracellular extracts of Acinetobac-
ter sp. SM04 liquid cultures. Biodegradation, 2011, 22(3),
613-622.
[http://dx.doi.org/10.1007/s10532-010-9435-z] [PMID:
21082331]
Cserháti, M.; Kriszt, B.; Krifaton, C.; Szoboszlay, S.; Háhn, J.;[85]
Tóth, S.; Nagy, I.; Kukolya, J. Mycotoxin-degradation profile of
Rhodococcus strains. Int. J. Food Microbiol., 2013, 166(1),
176-185.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2013.06.002] [PMID:
23891865]
Dixit, SM; Johura, FT; Manandhar, S; Sadique, A; Rajbhandari,[86]
RM; Mannan, SB Cholera outbreaks (2012) in three districts of
Nepal reveal clonal transmission of multi-drug resistant Vibrio
cholerae O1, 2014.
Altalhi, A.D. Plasmid-inediated detoxification of mycotoxin[87]
zearalenone in Pseudomonas sp. ZEA-1. Am. J. Biochem. Biotech-
nol., 2007, 3, 150-158.
[http://dx.doi.org/10.3844/ajbbsp.2007.150.158]
Altalhi, A.D.; El-Deeb, B. Localization of zearalenone detoxifica-[88]
tion gene(s) in pZEA-1 plasmid of Pseudomonas putida ZEA-1
and expressed in Escherichia coli. J. Hazard. Mater., 2009,
161(2-3), 1166-1172.
[http://dx.doi.org/10.1016/j.jhazmat.2008.04.068] [PMID:
18513857]
Yi, P-J.; Pai, C-K.; Liu, J-R. Isolation and characterization of a Ba-[89]
cillus licheniformis strain capable of degrading zearalenone.
World J. Microbiol. Biotechnol., 2011, 27(5), 1035-1043.
[http://dx.doi.org/10.1007/s11274-010-0548-7]
Ji, C.; Fan, Y.; Zhao, L. Review on biological degradation of my-[90]
cotoxins. Anim Nutr, 2016, 2(3), 127-133.
[http://dx.doi.org/10.1016/j.aninu.2016.07.003] [PMID:
29767078]
Miazzo, R.; Rosa, C.A.; De Queiroz Carvalho, E.C.; Magnoli, C.;[91]
Chiacchiera, S.M.; Palacio, G.; Saenz, M.; Kikot, A.; Basaldella,
E.; Dalcero, A. Efficacy of synthetic zeolite to reduce the toxicity
of aflatoxin in broiler chicks. Poult. Sci., 2000, 79(1), 1-6.
[http://dx.doi.org/10.1093/ps/79.1.1] [PMID: 10685881]
Boudergue, C; Burel, C; Dragacci, S Review of mycotoxin‐detox-[92]
ifying agents used as feed additives: mode of action, efficacy and
feed/food safety EFSA Supporting Publications, 2009, 6(9 ), 22E.
Usall, J.; Teixido, N.; Torres, R.; de Eribe, X.O.; Viñas, I. Pilot[93]
tests of Candida sake (CPA-1) applications to control postharvest
blue mold on apple fruit. Postharvest Biol. Technol., 2001, 21(2),
147-156.
[http://dx.doi.org/10.1016/S0925-5214(00)00131-9]
Jijakli, M.; Grevesse, C.; Lepoivre, P. Modes of action of biocon-[94]
trol agents of postharvest diseases: challenges and difficulties. Bul-
letin OILB/SROP= IOBC. Bull. SROP, 2001, 24(3), 317-318.
Kogan, G.; Kocher, A. Role of yeast cell wall polysaccharides in[95]
pig nutrition and health protection. Livest. Sci., 2007, 109(1-3),
161-165.
[http://dx.doi.org/10.1016/j.livsci.2007.01.134]
Franklin, S.T.; Newman, M.C.; Newman, K.E.; Meek, K.I. Im-[96]
mune parameters of dry cows fed mannan oligosaccharide and sub-
sequent transfer of immunity to calves. J. Dairy Sci., 2005, 88(2),
766-775.
[http://dx.doi.org/10.3168/jds.S0022-0302(05)72740-5] [PMID:
15653543]
Nochta, I.; Tuboly, T.; Halas, V.; Babinszky, L. Effect of different[97]
levels of mannan-oligosaccharide supplementation on some im-
munological variables in weaned piglets. J. Anim. Physiol. Anim.
Nutr. (Berl.), 2009, 93(4), 496-504.
[http://dx.doi.org/10.1111/j.1439-0396.2008.00835.x] [PMID:
18700854]
Swamy, H.V.; Smith, T.K.; Cotter, P.F.; Boermans, H.J.; Sefton,[98]
A.E. Effects of feeding blends of grains naturally contaminated
16 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
with Fusarium mycotoxins on production and metabolism in broil-
ers. Poult. Sci., 2002, 81(7), 966-975.
[http://dx.doi.org/10.1093/ps/81.7.966] [PMID: 12162357]
Tsitsigiannis, D.I.; Dimakopoulou, M.; Antoniou, P.P.; Tjamos,[99]
E.C. Biological control strategies of mycotoxigenic fungi and asso-
ciated mycotoxins in Mediterranean basin crops. Phytopathol.
Mediterr., 2012, 51(1), 158-174.
Khan, N.I.; Schisler, D.A.; Boehm, M.J.; Slininger, P.J.; Bothast,[100]
R.J. Selection and evaluation of microorganisms for biocontrol of
Fusarium head blight of wheat incited by Gibberella zeae. Plant
Dis., 2001, 85(12), 1253-1258.
[http://dx.doi.org/10.1094/PDIS.2001.85.12.1253] [PMID:
30831786]
Adebo, O.A.; Njobeh, P.B.; Gbashi, S.; Nwinyi, O.C.; Mavumeng-[101]
wana, V. Review on microbial degradation of aflatoxins. Crit.
Rev. Food Sci. Nutr., 2017, 57(15), 3208-3217.
[http://dx.doi.org/10.1080/10408398.2015.1106440] [PMID:
26517507]
De Jong, E.; Field, J.A.; de Bont, J.A. Aryl alcohols in the physiol-[102]
ogy of ligninolytic fungi. FEMS Microbiol. Rev., 1994, 13(2-3),
153-187.
[http://dx.doi.org/10.1111/j.1574-6976.1994.tb00041.x]
Alberts, J.F.; Gelderblom, W.C.; Botha, A.; van Zyl, W.H. Degra-[103]
dation of aflatoxin B(1) by fungal laccase enzymes. Int. J. Food
Microbiol., 2009, 135(1), 47-52.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2009.07.022] [PMID:
19683355]
Stanley, V.G.; Ojo, R.; Woldesenbet, S.; Hutchinson, D.H.; Kube-[104]
na, L.F. The use of Saccharomyces cerevisiae to suppress the ef-
fects of aflatoxicosis in broiler chicks. Poult. Sci., 1993, 72(10),
1867-1872.
[http://dx.doi.org/10.3382/ps.0721867] [PMID: 8415359]
Huynh, V.L.; Lloyd, A.B. Synthesis and degradation of aflatoxins[105]
by Aspergillus parasiticus. I. Synthesis of aflatoxin B1 by young
mycelium and its subsequent degradation in aging mycelium.
Aust. J. Biol. Sci., 1984, 37(1-2), 37-43.
[http://dx.doi.org/10.1071/BI9840037] [PMID: 6439179]
Kusumaningtyas, E.; Widiastuti, R.; Maryam, R. Reduction of[106]
aflatoxin B1 in chicken feed by using Saccharomyces cerevisiae,
Rhizopus oligosporus and their combination. Mycopathologia,
2006, 162(4), 307-311.
[http://dx.doi.org/10.1007/s11046-006-0047-4] [PMID:
17039279]
Zhang, W.; Xue, B.; Li, M.; Mu, Y.; Chen, Z.; Li, J.; Shan, A.[107]
Screening a strain of Aspergillus niger and optimization of fermen-
tation conditions for degradation of aflatoxin B₁. Toxins (Basel),
2014, 6(11), 3157-3172.
[http://dx.doi.org/10.3390/toxins6113157] [PMID: 25401962]
el-Sharkawy, S.; Abul-Hajj, Y. Microbial transformation of[108]
zearalenone, I. Formation of zearalenone-4-O-β-glucoside. J. Nat.
Prod., 1987, 50(3), 520-521.
[http://dx.doi.org/10.1021/np50051a038]
El-Sharkawy, S.H.; Abul-Hajj, Y.J. Microbial transformation of[109]
zearalenone. 2. Reduction, hydroxylation, and methylation prod-
ucts. J. Org. Chem., 1988, 53(3), 515-519.
[http://dx.doi.org/10.1021/jo00238a008]
el-Sharkawy, S.; Abul-Hajj, Y.J. Microbial cleavage of[110]
zearalenone. Xenobiotica, 1988, 18(4), 365-371.
[http://dx.doi.org/10.3109/00498258809041672] [PMID:
2969647]
Molnar, O.; Schatzmayr, G.; Fuchs, E.; Prillinger, H. Trichos-[111]
poron mycotoxinivorans sp. nov., a new yeast species useful in bi-
ological detoxification of various mycotoxins. Syst. Appl. Microbi-
ol., 2004, 27(6), 661-671.
[http://dx.doi.org/10.1078/0723202042369947] [PMID:
15612623]
Jard, G.; Liboz, T.; Mathieu, F.; Guyonvarc’h, A.; André, F.; Dela-[112]
forge, M. Transformation of zearalenone to zearalenone-sulfate by
Aspergillus spp. World Mycotoxin J., 2010, 3(2), 183-191.
[http://dx.doi.org/10.3920/WMJ2009.1184]
Vekiru, E.; Hametner, C.; Mitterbauer, R.; Rechthaler, J.; Adam,[113]
G.; Schatzmayr, G.; Krska, R.; Schuhmacher, R. Cleavage of
zearalenone by Trichosporon mycotoxinivorans to a novel none-
strogenic metabolite. Appl. Environ. Microbiol., 2010, 76(7),
2353-2359.
[http://dx.doi.org/10.1128/AEM.01438-09] [PMID: 20118365]
el-Sharkaway, S.H.; Selim, M.I.; Afifi, M.S.; Halaweish, F.T. Mi-[114]
crobial transformation of zearalenone to a zearalenone sulfate. Ap-
pl. Environ. Microbiol., 1991, 57(2), 549-552.
[http://dx.doi.org/10.1128/AEM.57.2.549-552.1991] [PMID:
1826596]
Utermark, J.; Karlovsky, P. Role of zearalenone lactonase in pro-[115]
tection of Gliocladium roseum from fungitoxic effects of the my-
cotoxin zearalenone. Appl. Environ. Microbiol., 2007, 73(2),
637-642.
[http://dx.doi.org/10.1128/AEM.01440-06] [PMID: 17114328]
Repedkiene, J.; Levinskaitė, L.; Paškevičius, A.; Raudonienė, V.[116]
Toxin-producing fungi on feed grains and application of yeasts for
their detoxification. Pol. J. Vet. Sci., 2013, 16(2), 391-393.
[http://dx.doi.org/10.2478/pjvs-2013-0054] [PMID: 23971211]
Angioni, A.; Caboni, P.; Garau, A.; Farris, A.; Orro, D.; Budroni,[117]
M.; Cabras, P. In vitro interaction between ochratoxin A and dif-
ferent strains of Saccharomyces cerevisiae and Kloeckera apicula-
ta. J. Agric. Food Chem., 2007, 55(5), 2043-2048.
[http://dx.doi.org/10.1021/jf062768u] [PMID: 17279767]
Schatzmayr, G.; Heidler, D.; Fuchs, E.; Nitsch, S.; Mohnl, M.;[118]
Täubel, M.; Loibner, A.P.; Braun, R.; Binder, E.M. Investigation
of different yeast strains for the detoxification of ochratoxin A.
Mycotoxin Res., 2003, 19(2), 124-128.
[http://dx.doi.org/10.1007/BF02942950] [PMID: 23604763]
Velmourougane, K.; Bhat, R.; Gopinandhan, T.; Panneerselvam,[119]
P. Management of Aspergillus ochraceus and Ochratoxin-A con-
tamination in coffee during on-farm processing through commer-
cial yeast inoculation. Biol. Control, 2011, 57(3), 215-221.
[http://dx.doi.org/10.1016/j.biocontrol.2011.03.003]
Bejaoui, H.; Mathieu, F.; Taillandier, P.; Lebrihi, A. Ochratoxin A[120]
removal in synthetic and natural grape juices by selected oenologi-
cal Saccharomyces strains. J. Appl. Microbiol., 2004, 97(5),
1038-1044.
[http://dx.doi.org/10.1111/j.1365-2672.2004.02385.x] [PMID:
15479420]
Ponsone, M.L.; Chiotta, M.L.; Combina, M.; Dalcero, A.; Chulze,[121]
S. Biocontrol as a strategy to reduce the impact of ochratoxin A
and Aspergillus section Nigri in grapes. Int. J. Food Microbiol.,
2011, 151(1), 70-77.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2011.08.005] [PMID:
21893359]
Virgili, R.; Simoncini, N.; Toscani, T.; Camardo Leggieri, M.; For-[122]
menti, S.; Battilani, P. Biocontrol of Penicillium nordicum growth
and ochratoxin A production by native yeasts of dry cured ham.
Toxins (Basel), 2012, 4(2), 68-82.
[http://dx.doi.org/10.3390/toxins4020068] [PMID: 22474567]
Caridi, A.; Galvano, F.; Tafuri, A.; Ritieni, A. In-vitro screening[123]
ofSaccharomyces strains for ochratoxin A removal from liquid
medium. Ann. Microbiol., 2006, 56(2), 135.
[http://dx.doi.org/10.1007/BF03174994]
Štyriak, I.; Conková, E.; Kmec, V.; Böhm, J.; Razzazi, E. The use[124]
of yeast for microbial degradation of some selected mycotoxins.
Mycotoxin Res., 2001, 17(1)(Suppl. 1), 24-27.
[http://dx.doi.org/10.1007/BF03036705] [PMID: 23605753]
Paškevičius, A.; Bakutis, B.; Baliukonienė, V.; Šakalytė, J. The[125]
search for ecologically safe means of mycotoxin detoxification in
fodder. Ekologija (Liet. Moksl. Akad.), 2006, (3), 128-131.
Karlovsky, P. Biological detoxification of fungal toxins and its[126]
use in plant breeding, feed and food production. Nat. Toxins,
1999, 7(1), 1-23.
[http://dx.doi.org/10.1002/(SICI)1522-7189(199902)7:1<1::AID-
NT37>3.0.CO;2-9] [PMID: 10441033]
He, C.; Fan, Y.; Liu, G.; Zhang, H. Isolation and identification of[127]
a strain of Aspergillus tubingensis with deoxynivalenol biotrans-
formation capability. Int. J. Mol. Sci., 2008, 9(12), 2366-2375.
[http://dx.doi.org/10.3390/ijms9122366] [PMID: 19330081]
Stinson, E.; Osman, S.; Bills, D. Water‐soluble products from pat-[128]
ulin during alcoholic fermentation of apple juice. J. Food Sci.,
1979, 44(3), 788-789.
[http://dx.doi.org/10.1111/j.1365-2621.1979.tb08502.x]
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 17
Moss, M.O.; Long, M.T. Fate of patulin in the presence of the[129]
yeast Saccharomyces cerevisiae. Food Addit. Contam., 2002,
19(4), 387-399.
[http://dx.doi.org/10.1080/02652030110091163] [PMID:
11962697]
Ricelli, A.; Baruzzi, F.; Solfrizzo, M.; Morea, M.; Fanizzi, F.P.[130]
Biotransformation of patulin by Gluconobacter oxydans. Appl. En-
viron. Microbiol., 2007, 73(3), 785-792.
[http://dx.doi.org/10.1128/AEM.02032-06] [PMID: 17114325]
Stinson, E.E.; Osman, S.F.; Huhtanen, C.N.; Bills, D.D. Disappear-[131]
ance of patulin during alcoholic fermentation of apple juice. Appl.
Environ. Microbiol., 1978, 36(4), 620-622.
[http://dx.doi.org/10.1128/AEM.36.4.620-622.1978] [PMID:
360989]
Coelho, A.R.; Celli, M.G.; Ono, E.Y.S.; Wosiacki, G.; Hoffmann,[132]
F.L.; Pagnocca, F.C. Penicillium expansum versus antagonist
yeasts and patulin degradation in vitro. Braz. Arch. Biol. Technol.,
2007, 50(4), 725-733.
[http://dx.doi.org/10.1590/S1516-89132007000400019]
Ianiri, G.; Idnurm, A.; Wright, S.A.; Durán-Patrón, R.; Mannina,[133]
L.; Ferracane, R.; Ritieni, A.; Castoria, R. Searching for genes re-
sponsible for patulin degradation in a biocontrol yeast provides in-
sight into the basis for resistance to this mycotoxin. Appl. Environ.
Microbiol., 2013, 79(9), 3101-3115.
[http://dx.doi.org/10.1128/AEM.03851-12] [PMID: 23455346]
Castoria, R.; Mannina, L.; Durán-Patrón, R.; Maffei, F.; Sobolev,[134]
A.P.; De Felice, D.V.; Pinedo-Rivilla, C.; Ritieni, A.; Ferracane,
R.; Wright, S.A. Conversion of the mycotoxin patulin to the less
toxic desoxypatulinic acid by the biocontrol yeast Rhodosporidi-
um kratochvilovae strain LS11. J. Agric. Food Chem., 2011,
59(21), 11571-11578.
[http://dx.doi.org/10.1021/jf203098v] [PMID: 21928828]
Reddy, K.R.; Spadaro, D.; Gullino, M.L.; Garibaldi, A. Potential[135]
of two Metschnikowia pulcherrima (yeast) strains for in vitro
biodegradation of patulin. J. Food Prot., 2011, 74(1), 154-156.
[http://dx.doi.org/10.4315/0362-028X.JFP-10-331] [PMID:
21219780]
Aliabadi, M.A.; Alikhani, F.E.; Mohammadi, M.; Darsanaki, R.K.[136]
Biological control of aflatoxins. Eur. J. Exp. Biol., 2013, 3(2),
162-166.
Heinl, S.; Hartinger, D.; Thamhesl, M.; Schatzmayr, G.; Moll, W-[137]
D.; Grabherr, R. An aminotransferase from bacterium ATCC
55552 deaminates hydrolyzed fumonisin B₁. Biodegradation,
2011, 22(1), 25-30.
[http://dx.doi.org/10.1007/s10532-010-9371-y] [PMID:
20567881]
Duvick, J.; Maddox, J.; Gilliam, J. Compositions and methods for[138]
fumonisin detoxification. In: Google Patents, 2003.
Heinl, S.; Hartinger, D.; Moll, W.; Schatzmayr, G.; Grabherr, R.[139]
Identification of a fumonisin B1 degrading gene cluster in Sphin-
gomonas spp. MTA144. N. Biotechnol., 2009, 25, S61-S62.
[http://dx.doi.org/10.1016/j.nbt.2009.06.290]
Benedetti, R.; Nazzi, F.; Locci, R.; Firrao, G. Degradation of fu-[140]
monisin B1 by a bacterial strain isolated from soil. Biodegrada-
tion, 2006, 17(1), 31-38.
[http://dx.doi.org/10.1007/s10532-005-2797-y] [PMID:
16453169]
Vanhoutte, I.; Audenaert, K.; De Gelder, L. Biodegradation of my-[141]
cotoxins: Tales from known and unexplored worlds. Front. Micro-
biol., 2016, 7, 561.
[http://dx.doi.org/10.3389/fmicb.2016.00561] [PMID: 27199907]
Matthies, I.; Woerfel, G.; Karlovsky, P. Induction of a[142]
zearalenone degrading enzyme caused by the substrate and its deri-
vatives. Mycotoxin Res., 2001, 17(Suppl. 1), 28-31.
[http://dx.doi.org/10.1007/BF03036706] [PMID: 23605754]
Kakeya, H.; Takahashi-Ando, N.; Kimura, M.; Onose, R.; Yam-[143]
aguchi, I.; Osada, H. Biotransformation of the mycotoxin,
zearalenone, to a non-estrogenic compound by a fungal strain of
Clonostachys sp. Biosci. Biotechnol. Biochem., 2002, 66(12),
2723-2726.
[http://dx.doi.org/10.1271/bbb.66.2723] [PMID: 12596876]
Popiel, D.; Koczyk, G.; Dawidziuk, A.; Gromadzka, K.;[144]
Blaszczyk, L.; Chelkowski, J. Zearalenone lactonohydrolase activi-
ty in Hypocreales and its evolutionary relationships within the
epoxide hydrolase subset of a/b-hydrolases. BMC Microbiol.,
2014, 14(1), 82.
[http://dx.doi.org/10.1186/1471-2180-14-82] [PMID: 24708405]
Tinyiro, S.E.; Wokadala, C.; Xu, D.; Yao, W. Adsorption and[145]
degradation of zearalenone by bacillus strains. Folia Microbiol.
(Praha), 2011, 56(4), 321-327.
[http://dx.doi.org/10.1007/s12223-011-0047-8] [PMID:
21647705]
Sun, X.; He, X.; Xue, Ks.; Li, Y.; Xu, D.; Qian, H. Biological de-[146]
toxification of zearalenone by Aspergillus niger strain FS10. Food
Chem. Toxicol., 2014, 72, 76-82.
[http://dx.doi.org/10.1016/j.fct.2014.06.021] [PMID: 25007785]
Yu, Y.; Qiu, L.; Wu, H.; Tang, Y.; Lai, F.; Yu, Y. Oxidation of[147]
zearalenone by extracellular enzymes from Acinetobacter sp.
SM04 into smaller estrogenic products. World J. Microbiol.
Biotechnol., 2011, 27(11), 2675-2681.
[http://dx.doi.org/10.1007/s11274-011-0741-3]
Kriszt, R.; Krifaton, C.; Szoboszlay, S.; Cserháti, M.; Kriszt, B.;[148]
Kukolya, J.; Czéh, A.; Fehér-Tóth, S.; Török, L.; Szőke, Z.;
Kovács, K.J.; Barna, T.; Ferenczi, S. A new zearalenone biodegra-
dation strategy using non-pathogenic Rhodococcus pyridinivorans
K408 strain. PLoS One, 2012, 7(9)e43608
[http://dx.doi.org/10.1371/journal.pone.0043608] [PMID:
23049739]
Ueno, Y.; Nakayama, K.; Ishii, K.; Tashiro, F.; Minoda, Y.;[149]
Omori, T.; Komagata, K. Metabolism of T-2 toxin in Curtobacteri-
um sp. strain 114-2. Appl. Environ. Microbiol., 1983, 46(1),
120-127.
[http://dx.doi.org/10.1128/AEM.46.1.120-127.1983] [PMID:
6614901]
Fuchs, E.; Binder, E.; Heidler, D.; Krska, R. Characterisation of[150]
metabolites after the microbial degradation of A- and B-tri-
chothecenes by BBSH 797. Mycotoxin Res., 2000, 16(1)(Suppl.
1), 66-69.
[http://dx.doi.org/10.1007/BF02942984] [PMID: 23605418]
Fuchs, E.; Binder, E.M.; Heidler, D.; Krska, R. Structural charac-[151]
terization of metabolites after the microbial degradation of type A
trichothecenes by the bacterial strain BBSH 797. Food Addit. Con-
tam., 2002, 19(4), 379-386.
[http://dx.doi.org/10.1080/02652030110091154] [PMID:
11962696]
He, J.; Zhou, T.; Young, J.C.; Boland, G.J.; Scott, P.M. Chemical[152]
and biological transformations for detoxification of trichothecene
mycotoxins in human and animal food chains: a review. Trends
Food Sci. Technol., 2010, 21(2), 67-76.
[http://dx.doi.org/10.1016/j.tifs.2009.08.002]
Guan, S.; He, J.; Young, J.C.; Zhu, H.; Li, X-Z.; Ji, C. Transforma-[153]
tion of trichothecene mycotoxins by microorganisms from fish di-
gesta. Aquaculture, 2009, 290(3), 290-295.
[http://dx.doi.org/10.1016/j.aquaculture.2009.02.037]
Rafiqul, I. Isolation, characterization and genome sequencing of a[154]
soil-borne Citrobacter freundii strain capable of detoxifying tri-
chothecene mycotoxins: SpringerLink, 2012.
Shima, J.; Takase, S.; Takahashi, Y.; Iwai, Y.; Fujimoto, H.; Ya-[155]
mazaki, M.; Ochi, K. Novel detoxification of the trichothecene my-
cotoxin deoxynivalenol by a soil bacterium isolated by enrichment
culture. Appl. Environ. Microbiol., 1997, 63(10), 3825-3830.
[http://dx.doi.org/10.1128/AEM.63.10.3825-3830.1997] [PMID:
9327545]
Völkl, A.; Vogler, B.; Schollenberger, M.; Karlovsky, P. Micro-[156]
bial detoxification of mycotoxin deoxynivalenol. J. Basic Microbi-
ol., 2004, 44(2), 147-156.
[http://dx.doi.org/10.1002/jobm.200310353] [PMID: 15069674]
Ikunaga, Y.; Sato, I.; Grond, S.; Numaziri, N.; Yoshida, S.; Ya-[157]
maya, H.; Hiradate, S.; Hasegawa, M.; Toshima, H.; Koitabashi,
M.; Ito, M.; Karlovsky, P.; Tsushima, S. Nocardioides sp. strain
WSN05-2, isolated from a wheat field, degrades deoxynivalenol,
producing the novel intermediate 3-epi-deoxynivalenol. Appl. Mi-
crobiol. Biotechnol., 2011, 89(2), 419-427.
[http://dx.doi.org/10.1007/s00253-010-2857-z] [PMID:
20857291]
Sato, I.; Ito, M.; Ishizaka, M.; Ikunaga, Y.; Sato, Y.; Yoshida, S.;[158]
18 Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 Abdi et al.
Koitabashi, M.; Tsushima, S. Thirteen novel deoxynivalenol-de-
grading bacteria are classified within two genera with distinct
degradation mechanisms. FEMS Microbiol. Lett., 2012, 327(2),
110-117.
[http://dx.doi.org/10.1111/j.1574-6968.2011.02461.x] [PMID:
22098388]
Ciegler, A.; Lillehoj, E.B.; Peterson, R.E.; Hall, H.H. Microbial[159]
detoxification of aflatoxin. Appl. Microbiol., 1966, 14(6),
934-939.
[http://dx.doi.org/10.1128/AEM.14.6.934-939.1966] [PMID:
16349699]
Wang, J.; Ogata, M.; Hirai, H.; Kawagishi, H. Detoxification of[160]
aflatoxin B1 by manganese peroxidase from the white-rot fungus
Phanerochaete sordida YK-624. FEMS Microbiol. Lett., 2011,
314(2), 164-169.
[http://dx.doi.org/10.1111/j.1574-6968.2010.02158.x] [PMID:
21118293]
Liu, D-L.; Yao, D-S.; Liang, R.; Ma, L.; Cheng, W-Q.; Gu, L-Q.[161]
Detoxification of aflatoxin B1 by enzymes isolated from Armil-
lariella tabescens. Food Chem. Toxicol., 1998, 36(7), 563-574.
[http://dx.doi.org/10.1016/S0278-6915(98)00017-9] [PMID:
9687963]
Das, A.; Bhattacharya, S.; Palaniswamy, M.; Angayarkanni, J.[162]
Biodegradation of aflatoxin B1 in contaminated rice straw by Pleu-
rotus ostreatus MTCC 142 and Pleurotus ostreatus GHBBF10 in
the presence of metal salts and surfactants. World J. Microbiol.
Biotechnol., 2014, 30(8), 2315-2324.
[http://dx.doi.org/10.1007/s11274-014-1657-5] [PMID:
24770873]
Méndez-Albores, A.; Arámbula-Villa, G.; Loarca-Piña, M.G.; Cas-[163]
taño-Tostado, E.; Moreno-Martínez, E. Safety and efficacy evalua-
tion of aqueous citric acid to degrade B-aflatoxins in maize. Food
Chem. Toxicol., 2005, 43(2), 233-238.
[http://dx.doi.org/10.1016/j.fct.2004.09.009] [PMID: 15621335]
Samuel, M.S.; Sivaramakrishna, A.; Mehta, A. Degradation and[164]
detoxification of aflatoxin B1 by Pseudomonas putida. Int. Biode-
terior. Biodegradation, 2014, 86, 202-209.
[http://dx.doi.org/10.1016/j.ibiod.2013.08.026]
Dridi, F.; Marrakchi, M.; Gargouri, M.; Saulnier, J.; Jaffrezic-Re-[165]
nault, N.; Lagarde, F. Comparison of carboxypeptidase Y and ther-
molysin for ochratoxin A electrochemical biosensing. Anal.
Methods, 2015, 7(20), 8954-8960.
[http://dx.doi.org/10.1039/C5AY01905B]
Abrunhosa, L.; Santos, L.; Venâncio, A. Degradation of ochra-[166]
toxin A by proteases and by a crude enzyme of Aspergillus niger.
Food Biotechnol., 2006, 20(3), 231-242.
[http://dx.doi.org/10.1080/08905430600904369]
Hatab, S.; Yue, T.; Mohamad, O. Reduction of patulin in aqueous[167]
solution by lactic acid bacteria. J. Food Sci., 2012, 77(4), M238-
M241.
[http://dx.doi.org/10.1111/j.1750-3841.2011.02615.x] [PMID:
22394296]
Kabak, B.; Var, I. Factors affecting the removal of aflatoxin M1[168]
from food model by Lactobacillus and Bifidobacterium strains. J.
Environ. Sci. Health B, 2008, 43(7), 617-624.
[http://dx.doi.org/10.1080/03601230802234740] [PMID:
18803117]
KABAK, B.; VAR, I. Binding of aflatoxin M1 by Lactobacillus[169]
and Bifidobacterium strains. Milchwissenschaft, 2004, 59(5-6),
301-303.
El-Nezami, H. Biologic control of food carcinogen using Lactoba-[170]
cillus GG. Nutr. Today, 1996, 31, 41S-43S.
[http://dx.doi.org/10.1097/00017285-199611001-00013]
Hernandez-Mendoza, A.; Guzman-De-Peña, D.; González-Córdo-[171]
va, A.F.; Vallejo-Córdoba, B.; Garcia, H.S. In vivo assessment of
the potential protective effect of Lactobacillus casei Shirota
against aflatoxin B1. Dairy Sci. Technol., 2010, 90(6), 729-740.
[http://dx.doi.org/10.1051/dst/2010030]
Halttunen, T.; Collado, M.C.; El-Nezami, H.; Meriluoto, J.; Salmi-[172]
nen, S. Combining strains of lactic acid bacteria may reduce their
toxin and heavy metal removal efficiency from aqueous solution.
Lett. Appl. Microbiol., 2008, 46(2), 160-165.
[http://dx.doi.org/10.1111/j.1472-765X.2007.02276.x] [PMID:
18028332]
Pierides, M.; El-Nezami, H.; Peltonen, K.; Salminen, S.; Ahokas,[173]
J. Ability of dairy strains of lactic acid bacteria to bind aflatoxin
M1 in a food model. J. Food Prot., 2000, 63(5), 645-650.
[http://dx.doi.org/10.4315/0362-028X-63.5.645] [PMID:
10826723]
Škrinjar, M.; Rasić, J.L.; Stojicić, V. Lowering of ochratoxin A[174]
level in milk by yoghurt bacteria and bifidobacteria. Folia Micro-
biol. (Praha), 1996, 41(1), 26-28.
[http://dx.doi.org/10.1007/BF02816335] [PMID: 9090820]
Niderkorn, V.; Morgavi, D.P.; Aboab, B.; Lemaire, M.; Boudra,[175]
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(3), 977-985.
[http://dx.doi.org/10.1111/j.1365-2672.2008.04065.x] [PMID:
19187153]
El-Nezami, H.; Polychronaki, N.; Salminen, S.; Mykkänen, H.[176]
Binding rather than metabolism may explain the interaction of two
food-Grade Lactobacillus strains with zearalenone and its deriva-
tive (')alpha-earalenol. Appl. Environ. Microbiol., 2002, 68(7),
3545-3549.
[http://dx.doi.org/10.1128/AEM.68.7.3545-3549.2002] [PMID:
12089040]
G.A., G.; A.A.M. Y, Abol-ElaM.F. inhibition of aspergillus flavus[177]
and aspergillus parasiticus fungal growth and its aflatoxins (B1,
B2, G1AND G2) production by lactobacillus acidophilus. J.
Egypt. Soc. Toxicol., 2007, 37, 53-60.
ELSANHOTY, RM; AZEKE, MA ASSESSMENT OF THE[178]
ABILITY OF SOME PROBIOTIC BACTERIA TO BIND AND
REMOVE AFLATOXIN FROM CONTAMINATED WHEAT
DURING BALADI BREAD MAKING. NIGERIAN ANNALS OF
NATURAL SCIENCES, 2009, 9(1), 49-59.
Kabak, B.; Ozbey, F. Assessment of the bioaccessibility of afla-[179]
toxins from various food matrices using an in vitro digestion mod-
el, and the efficacy of probiotic bacteria in reducing bioaccessibili-
ty. J. Food Compos. Anal., 2012, 27(1), 21-31.
[http://dx.doi.org/10.1016/j.jfca.2012.04.006]
Niknejad, F.; Zaini, F.; Faramarzi, M.; Amini, M.; Kordbacheh,[180]
P.; Mahmoudi, M.; Safara, M. Candida parapsilosis as a potent bio-
control agent against growth and aflatoxin production by Aspergil-
lus species. Iran. J. Public Health, 2012, 41(10), 72-80.
[PMID: 23308351]
Chen, Y.; Kong, Q.; Chi, C.; Shan, S.; Guan, B. Biotransforma-[181]
tion of aflatoxin B1 and aflatoxin G1 in peanut meal by anaerobic
solid fermentation of Streptococcus thermophilus and Lactobacil-
lus delbrueckii subsp. bulgaricus. Int. J. Food Microbiol., 2015,
211, 1-5.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2015.06.021] [PMID:
26143229]
Huang, W.; Chang, J.; Wang, P.; Liu, C.; Yin, Q.; Zhu, Q.; Lu, F.;[182]
Gao, T. Effect of the combined compound probiotics with myco-
toxin-degradation enzyme on detoxifying aflatoxin B1 and
zearalenone. J. Toxicol. Sci., 2018, 43(6), 377-385.
[http://dx.doi.org/10.2131/jts.43.377] [PMID: 29877214]
Das, K.M.; Lee, E.Y.; Al Jawder, S.E.; Enani, M.A.; Singh, R.;[183]
Skakni, L.; Al-Nakshabandi, N.; AlDossari, K.; Larsson, S.G.
Acute Middle East respiratory syndrome coronavirus: temporal
lung changes observed on the chest radiographs of 55 patients.
AJR Am. J. Roentgenol., 2015, 205(3)W267-74
[http://dx.doi.org/10.2214/AJR.15.14445] [PMID: 26102309]
Gao, X.; Ma, Q.; Zhao, L.; Lei, Y.; Shan, Y.; Ji, C. Isolation of Ba-[184]
cillus subtilis: screening for aflatoxins B 1, M 1, and G 1 detoxifi-
cation. Eur. Food Res. Technol., 2011, 232(6), 957.
[http://dx.doi.org/10.1007/s00217-011-1463-3]
El-Deeb, B.; Altalhi, A.; Khiralla, G.; Hassan, S.; Gherbawy, Y.[185]
Isolation and characterization of endophytic Bacilli bacterium
from maize grains able to detoxify aflatoxin B1. Food
Biotechnol., 2013, 27(3), 199-212.
[http://dx.doi.org/10.1080/08905436.2013.811083]
Rao, K.R.; Vipin, A.; Hariprasad, P.; Appaiah, K.A.;[186]
Venkateswaran, G. Biological detoxification of Aflatoxin B1 by
Bacillus licheniformis CFR1. Food Control, 2017, 71, 234-241.
[http://dx.doi.org/10.1016/j.foodcont.2016.06.040]
Microbiological Detoxification of Mycotoxins Infectious Disorders - Drug Targets, 2020, Vol. 20, No. 0 19
Krifaton, C.; Kriszt, B.; Szoboszlay, S.; Cserháti, M.; Szűcs, A.;[187]
Kukolya, J. Analysis of aflatoxin-B1-degrading microbes by use
of a combined toxicity-profiling method. Mutat. Res., 2011,
726(1), 1-7.
[http://dx.doi.org/10.1016/j.mrgentox.2011.07.011] [PMID:
21871580]
Adebo, O.A.; Njobeh, P.B.; Mavumengwana, V. Degradation and[188]
detoxification of AFB1 by Staphylocococcus warneri, Sporosarci-
na sp. and Lysinibacillus fusiformis. Food Control, 2016, 68,
92-96.
[http://dx.doi.org/10.1016/j.foodcont.2016.03.021]
Adebo, O.A.; Njobeh, P.B.; Sidu, S.; Tlou, M.G.; Mavumengwa-[189]
na, V. Aflatoxin B1 degradation by liquid cultures and lysates of
three bacterial strains. Int. J. Food Microbiol., 2016, 233, 11-19.
[http://dx.doi.org/10.1016/j.ijfoodmicro.2016.06.007] [PMID:
27294556]
Abbès, S.; Ben Salah-Abbès, J.; Jebali, R.; Younes, R.B.; Oues-[190]
lati, R. Interaction of aflatoxin B1 and fumonisin B1 in mice caus-
es immunotoxicity and oxidative stress: Possible protective role us-
ing lactic acid bacteria. J. Immunotoxicol., 2016, 13(1), 46-54.
[http://dx.doi.org/10.3109/1547691X.2014.997905] [PMID:
25585958]
Aiko, V.; Edamana, P.; Mehta, A. Decomposition and detoxifica-[191]
tion of aflatoxin B1 by lactic acid. J. Sci. Food Agric., 2016,
96(6), 1959-1966.
[http://dx.doi.org/10.1002/jsfa.7304] [PMID: 26095453]
PB, P; Selvam, P Determination of Antiaflatoxigenic Effect of pro-[192]
biotic strains in Sorghum bicolour. Biosci. Biotechnol. Res. Asia,
2016, 13(2), 1095-1100.
[http://dx.doi.org/10.13005/bbra/2138]
Haskard, C.A.; El-Nezami, H.S.; Kankaanpää, P.E.; Salminen, S.;[193]
Ahokas, J.T. Surface binding of aflatoxin B(1) by lactic acid bacte-
ria. Appl. Environ. Microbiol., 2001, 67(7), 3086-3091.
[http://dx.doi.org/10.1128/AEM.67.7.3086-3091.2001] [PMID:
11425726]
Kurhan, Ş; Çakir, I. DNA-bioprotective effects of lactic acid bacte-[194]
ria against aflatoxin B1. Current Research in Nutrition and Food
Science Journal, 2016, 4(Special Issue Nutrition in Conference
October), 87-91.
Hamidi, A.; Mirnejad, R.; Majd, N.S.; Yahaghi, E.; Behnod, V.;[195]
Darian, E.K. The survey of potential of Aflatoxin B1 isolation by
lactic acid bacteria. Asian Pac. J. Trop. Biomed., 2012, 1-4.
Harkai, P.; Szabó, I.; Cserháti, M.; Krifaton, C.; Risa, A.; Radó, J.[196]
Biodegradation of aflatoxin-B1 and zearalenone by Streptomyces
sp. collection. Int. Biodeterior. Biodegradation, 2016, 108, 48-56.
[http://dx.doi.org/10.1016/j.ibiod.2015.12.007]
Shantha, T. Fungal degradation of aflatoxin B1. Nat. Toxins,[197]
1999, 7(5), 175-178.
[http://dx.doi.org/10.1002/1522-7189(200009/10)7:5<175::AID-N
T63>3.0.CO;2-M] [PMID: 10945479]
Hackbart, H.C.; Machado, A.R.; Christ-Ribeiro, A.; Prietto, L.; Ba-[198]
diale-Furlong, E. Reduction of aflatoxins by Rhizopus oryzae and
Trichoderma reesei. Mycotoxin Res., 2014, 30(3), 141-149.
[http://dx.doi.org/10.1007/s12550-014-0202-6] [PMID:
24925827]
Detroy, R.W.; Hesseltine, C.W. Transformation of aflatoxin B1[199]
by steroid-hydroxylating fungi. Can. J. Microbiol., 1969, 15(6),
495-500.
[http://dx.doi.org/10.1139/m69-086] [PMID: 5816335]
Cole, R.J.; Kirksey, J.W.; Blankenship, B.R. Conversion of afla-[200]
toxin B 1 to isomeric hydroxy compounds by Rhizopus spp. J.
Agric. Food Chem., 1972, 20(6), 1100-1102.
[http://dx.doi.org/10.1021/jf60184a040] [PMID: 5083520]
Gonçalves, B.L.; Rosim, R.E.; de Oliveira, C.A.F.; Corassin, C.H.[201]
The in vitro ability of different Saccharomyces cerevisiae–based
products to bind aflatoxin B1. Food Control, 2015, 47, 298-300.
[http://dx.doi.org/10.1016/j.foodcont.2014.07.024]
Shcherbakova, L.; Statsyuk, N.; Mikityuk, O.; Nazarova, T.; Dzha-[202]
vakhiya, V. Aflatoxin B1 degradation by metabolites of Phoma
glomerata PG41 isolated from natural substrate colonized by afla-
toxigenic Aspergillus flavus. Jundishapur J. Microbiol., 2015,
8(1)e24324
[http://dx.doi.org/10.5812/jjm.24324] [PMID: 25789135]
Sarlak, Z.; Rouhi, M.; Mohammadi, R.; Khaksar, R.; Mortazavian,[203]
A.M.; Sohrabvandi, S. Probiotic biological strategies to decontami-
nate aflatoxin M1 in a traditional Iranian fermented milk drink
(Doogh). Food Control, 2017, 71, 152-159.
[http://dx.doi.org/10.1016/j.foodcont.2016.06.037]
El Khoury, A.; Atoui, A.; Yaghi, J. Analysis of aflatoxin M1 in[204]
milk and yogurt and AFM1 reduction by lactic acid bacteria used
in Lebanese industry. Food Control, 2011, 22(10), 1695-1699.
[http://dx.doi.org/10.1016/j.foodcont.2011.04.001]
Sangsila, A.; Faucet-Marquis, V.; Pfohl-Leszkowicz, A.; It-[205]
saranuwat, P. Detoxification of zearalenone by Lactobacillus pen-
tosus strains. Food Control, 2016, 62, 187-192.
[http://dx.doi.org/10.1016/j.foodcont.2015.10.031]
Sellamani, M.; Kalagatur, N.K.; Siddaiah, C.; Mudili, V.; Krishna,[206]
K.; Natarajan, G.; Rao Putcha, V.L. Antifungal and zearalenone in-
hibitory activity of Pediococcus pentosaceus isolated from dairy
products on Fusarium graminearum. Front. Microbiol., 2016, 7,
890.
[http://dx.doi.org/10.3389/fmicb.2016.00890] [PMID: 27379035]
Zhao, L.; Jin, H.; Lan, J.; Zhang, R.; Ren, H.; Zhang, X. Detoxifi-[207]
cation of zearalenone by three strains of Lactobacillus plantarum
from fermented food in vitro. Food Control, 2015, 54, 158-164.
[http://dx.doi.org/10.1016/j.foodcont.2015.02.003]
Tian, Y.; Tan, Y.; Yan, Z.; Liao, Y.; Chen, J.; De Boevre, M.; De[208]
Saeger, S.; Wu, A. Antagonistic and detoxification potentials of
Trichoderma isolates for control of Zearalenone (ZEN) producing
Fusarium graminearum. Front. Microbiol., 2018, 8, 2710.
[http://dx.doi.org/10.3389/fmicb.2017.02710] [PMID: 29403455]
Haidukowski, M.; Cozzi, G.; Dipierro, N.; Bavaro, S.L.; Logrieco,[209]
A.F.; Paciolla, C. Decontamination of Fumonisin B 1 in maize
grain by Pleurotus eryngii and antioxidant enzymes. Phytopathol.
Mediterr., 2017, 56(1)
Martinez Tuppia, C.; Atanasova-Penichon, V.; Chéreau, S.; Ferr-[210]
er, N.; Marchegay, G.; Savoie, J.M.; Richard-Forget, F. Yeast and
bacteria from ensiled high moisture maize grains as potential miti-
gation agents of fumonisin B1. J. Sci. Food Agric., 2017, 97(8),
2443-2452.
[http://dx.doi.org/10.1002/jsfa.8058] [PMID: 27696424]
Yuan, Y.; Wang, X.; Hatab, S.; Wang, Z.; Wang, Y.; Luo, Y.;[211]
Yue, T. Patulin reduction in apple juice by inactivated Alicycloba-
cillus spp. Lett. Appl. Microbiol., 2014, 59(6), 604-609.
[http://dx.doi.org/10.1111/lam.12315] [PMID: 25130934]
Halasz, A.; Lasztity, R.; Abonyi, T.; Bata, A. Decontamination of[212]
mycotoxin-containing food and feed by biodegradation. Food Rev.
Int., 2009, 25(4), 284-298.
[http://dx.doi.org/10.1080/87559120903155750]
Tanasupawat, S.; Thawai, C.; Yukphan, P.; Moonmangmee, D.;[213]
Itoh, T.; Adachi, O.; Yamada, Y. Gluconobacter thailandicus sp.
nov., an acetic acid bacterium in the α-Proteobacteria. J. Gen. Ap-
pl. Microbiol., 2004, 50(3), 159-167.
[http://dx.doi.org/10.2323/jgam.50.159] [PMID: 15486825]
Franco, T.S.; Garcia, S.; Hirooka, E.Y.; Ono, Y.S.; dos Santos,[214]
J.S. Lactic acid bacteria in the inhibition of Fusarium gramin-
earum and deoxynivalenol detoxification. J. Appl. Microbiol.,
2011, 111(3), 739-748.
[http://dx.doi.org/10.1111/j.1365-2672.2011.05074.x] [PMID:
21672097]
Meca, G.; Ritieni, A.; Mañes, J. Reduction in vitro of the minor[215]
Fusarium mycotoxin beauvericin employing different strains of
probiotic bacteria. Food Control, 2012, 28(2), 435-440.
[http://dx.doi.org/10.1016/j.foodcont.2012.04.002]
Var, I.; Erginkaya, Z.; Kabak, B. Inhibition of ochratoxin A pro-[216]
duction of Aspergillus carbonarius by yeast species. Czech J.
Food Sci., 2011, 29(3), 291-297.
[http://dx.doi.org/10.17221/179/2009-CJFS]
... Various treatments have been suggested by scienti c communities for the management of toxigenic fungi associated with rice grains used in food and feed. These include chemical detoxi cation, alkalization, sodium hydroxide treatment, sulfur dioxide, oxidation, irradiation, and pyrolysis etc. Abdi et al. 2021). But each one of them have their own demerits viz., physical treatments (pyrolysis and irradiation) are ineffective and less speci c against extensive toxin contamination while chemical treatments with sulfur and sodium hydroxide adversely affect the seed germination, seed vigor and increase the chances of prolonged fungal resistance against chemicals (Schatzmayr et against the toxic FFSC complex, it is therefore necessary to investigate nano-scale adsorbents that are inexpensively accessible commercially and enhance their antifungal e cacy by adding environmentally benign antifungal natural extracts. ...
Preprint
Full-text available
Infestation of fungi during storage of grains causes about 4.6–6% wastage of cereal grains. Among other fungi Fusarium fujikuroi species complex ( FFSC ) is one of the major causes of grain discoloration of rice. Due to the negative impacts of toxicgenic FFSC complex on seed health and quality metrics, it significantly reduces the market value of rice seeds. Apart from nutrient losses to rice seeds toxic fungal metabolites like fumonisin B1, fumonisin B2, fusaric acid, and gibberellic acid produced by FFSC complex interacts with physiological parameters influencing the integrity of seeds during storage. Due to the detrimental effects on seed germination and vitality, chemical seed treatments are usually avoided while the seeds are in storage. This paper presents the efficacy of novel sodium alumino silicate (SAS)-Margosa composite against toxigenic FFSC complex viz., F. fujikuroi, F. verticillioides , and F. proliferatum during storage of rice seeds with augemented antitoxin and seed invigorating potential. The novel SAS-Margosa composite was chemically and morphologically characterized using XRD, FTIR, TEM, and EDS analysis. The antifungal potential of margosa leaf extract adsorbed nanosilicate material (SAS) was examined using SEM, Flourescence microscopy and poison food technique. The toxin quantification of FFSC complex inoculated seeds treated with SAS-Margosa composite was carried out at monthly intervals upto a period of six months of storage along with control on two rice varieties viz., PR114 and Pusa Basmati 1121 respectively. The study revealed that SAS-Margosa composite coated seeds of PR114 and Pusa Basmati 1121 exhibited significanlty higher seed germination (87.63 & 83.57%) and minimum seed rot (5.70 & 5.24%) after 6 months of storage as compare to control. There was a significant decrease in levels of fusaric acid, fumonisins B1, B2 and gibberellic acid after six months as compare to control in both the varieties. The SAS-Margosa composite has been proved to be effective against toxigenic FFSC complex during rice seed storage without compromising the seed quality metrics.
... To further reduce the content of mycotoxins in cereals and avoid the shortcomings of the established methods, it is imperative to develop new methods. This paper presents a variety of practical approaches to biotechnological detoxification, including the microbial adsorption and binding of mycotoxins, the competitive inhibition of the growth of toxin-producing molds or mycotoxin production, or the specific screening of certain microorganisms or enzymes to detoxify or produce non-toxic degradation products by destroying or modifying mycotoxins with the participation of their secondary metabolites or secreted intracellular and extracellular enzymes [171,172]. Furthermore, some previous studies were used to systematically summarize the mechanisms of various biotechnological detoxification methods, providing a more intuitive reference for preventing and controlling cereal contamination in China. Combining these findings with the specific conditions of the Chinese environment is expected to reduce mycotoxin contamination in grains during processing, transportation, and storage. ...
Article
Full-text available
The presence of mycotoxins in cereals can pose a significant health risk to animals and humans. China is one of the countries that is facing cereal contamination by mycotoxins. Treating mycotoxin-contaminated cereals with established physical and chemical methods can lead to negative effects, such as the loss of nutrients, chemical residues, and high energy consumption. Therefore, microbial detoxification techniques are being considered for reducing and treating mycotoxins in cereals. This paper reviews the contamination of aflatoxins, zearalenone, deoxynivalenol, fumonisins, and ochratoxin A in major cereals (rice, wheat, and maize). Our discussion is based on 8700 samples from 30 provincial areas in China between 2005 and 2021. Previous research suggests that the temperature and humidity in the highly contaminated Chinese cereal-growing regions match the growth conditions of potential antagonists. Therefore, this review takes biological detoxification as the starting point and summarizes the methods of microbial detoxification, microbial active substance detoxification, and other microbial inhibition methods for treating contaminated cereals. Furthermore, their respective mechanisms are systematically analyzed, and a series of strategies for combining the above methods with the treatment of contaminated cereals in China are proposed. It is hoped that this review will provide a reference for subsequent solutions to cereal contamination problems and for the development of safer and more efficient methods of biological detoxification.
... Due to its thermostability, it cannot be eliminated by typical food processing methods, leading to its accumulation in food products and beverages. Therefore, eliminating OTA from dietary components is essential for lowering the health risks that the toxin poses (Abdi et al. 2021). The potential solution is required to reduce and eliminate OTA from food sources. ...
Article
Full-text available
The environmental pollution caused by chemical dyes is a growing concern nowadays. Limitations of traditional methods opened the route for nanotechnology; owing to the versatile properties of nanomaterials, gold nanoparticles (AuNPs) became a potential strategy for different applications. In the present study, biosynthesis of gold nanoparticles (BioAuNPs) was carried out by reacting chloroauric acid (HAuCl4) with cell-free filtrate of Penicillium rubens sp. nov. NCIM 1937. The AuNPs were then characterized by UV–visible spectroscopy, HR-TEM, FTIR, and DLS analysis to further examine their efficacious biosynthesis and morphological properties including size, shape, and stability. The biogenic AuNPs are polydisperse in nature, with a mean size of 14.92 ± 5 nm. These AuNPs exhibited promising antimicrobial activity against Escherichia coli NCIM-2065, Bacillus subtilis NCIM-2010, and Penicillium verrucosum MTCC 4935. In vitro quantitative HPLC results revealed that BioAuNPs significantly inhibited the biosynthesis of ochratoxin A (OTA). Microbial fuel cells (MFCs) are intriguing for power generation and wastewater treatment since they can directly transform chemical energy stored in organic matter to electricity by extracellular electron transfer (EET) via membrane proteins. AuNPs also showed excellent potential for dye degradation of organic pollutants, viz., methylene blue (MB), phenol red (PR), bromothymol blue (BTB), Congo red (CR), and 4-nitrophenol (4-NP). All dye removal efficiencies were estimated and fitted to pseudo-first-order processes using kinetic rate constants (Ka).The present study reveals a simple, original, and eco-friendly method for the synthesis of multifunctional biogenic AuNPs that could be effective in OTA detoxification in food products and organic pollutant removal during wastewater treatment for a sustainable environment. Graphical Abstract
... Adsorption and/or degradation are the two mechanisms responsible for dye decolorization by microorganisms (Ohadi et al., 2020;Ghazvini et al., 2016). ...
Article
Full-text available
Background: The major concern to meet environmental regulations is related with the decolorization and detoxification of industrial dyes contaminated wastewater. So, this study was undertaken to examine the use of bacteria isolated from wastewater of textile factories in the removal of the synthetic textile dyes (Sudan Black, Methyl Red, Malachite Green, Rhodamine B and Brilliant Cresyl Blue). Methods: Dye contaminated wastewater was collected from some synthetic textile factories in Gorgan and Gonbad, Iran, and evaluated for the screening and isolation of bacteria capable of decolorizing textile dyes. The effect of function of operational parameters includes temperature (25, 37 and 50 °C), pH (4, 6 and 8) and initial dye concentration (100, 200 and 300 mg/mL) on the efficiency and rate of discoloration was assessed. Results: Totally, out of the 19 bacterial isolates from textile wastewater: Five bacterial isolates showed dye discoloration ability and the most efficient bacterial isolates were Enterococcus faecium and Pantoea spp. that decolorized Methyl Red, Sudan Black and Malachite Green dyes at 25-37°C, concentration of 200-300 mg/mL and slightly acidic to neutral pH. Enterococcus bacterium was able to decolorize Sudan Black to the 19.79% in the concentration of 100 mg/ml and pH=8 and temperature of 50°C. The highest amount of decolorizing was observed by Pantoea on Malachite Green to the amount of 73%. Enterococcus had the highest decolorizing on Methyl Red to the 65.7%. The amount of decolorizing on Sudan Black by Enterococcus (49.9%) was also higher than Pantoea (39.7%). Conclusion: Isolated bacteria had a significant reduction in toxicity and cationic malachite green dye and azo dye-methyl red. Thus, bacteria can be used in full-scale industrial wastewater treatment with the bio-synergy and its application in discoloration.
... One such example of yeast is Rhodosporidium kratochvilovae strain LS11 that degraded PAT into desoxypatulinic acid, which is less toxic . Biological detoxification involves two steps: reducing PAT toxicity by detoxification and removing PAT from solution by adsorption (Ohadi et al., 2020;Sajid et al., 2019). Some yeasts act as biocontrol agents by competing with plant pathogens for nutrients, thus inhibiting their growth (Sperandio et al., 2015). ...
Article
The mycotoxin patulin is primarily produced as a secondary metabolite by numerous fungal species and predominantly by Aspergillus, Byssochlamys, and Penicillium species. It is generally associated with fungal infected food materials. Penicillium expansum is considered the only fungal species liable for patulin contamination in pome fruits, especially in apples and apple-based products. This toxin in food poses serious health concerns and economic threat, which has aroused the need to adopt effective detection and mitigation strategies. Understanding its origin sources and biosynthetic mechanism stands essential for efficiently designing a management strategy against this fungal contamination. This review aims to present an updated outline of the sources of patulin occurrence in different foods and their biosynthetic mechanisms. It further provides information regarding the detrimental effects of patulin on human and agriculture as well as its effective detection, management, and control strategies.
Article
Zearalenone (ZEN) is a potent oestrogenic mycotoxin that is mainly produced by Fusarium species and is a serious environmental pollutant in animal feeds. Apiotrichum mycotoxinivorans has been widely used as a feed additive to detoxify ZEN. However, the effects of ZEN on A. mycotoxinivorans and its detoxification mechanisms remain unclear. In this study, transcriptomic and bioinformatic analyses were used to investigate the molecular responses of A. mycotoxinivorans to ZEN exposure and the genetic basis of ZEN detoxification. We detected 1424 significantly differentially expressed genes (DEGs), of which 446 were upregulated and 978 were downregulated. Functional and enrichment analyses showed that ZEN-induced genes were significantly associated with xenobiotic metabolism, oxidative stress response, and active transport systems. However, ZEN-inhibited genes were mainly related to cell division, cell cycle, and fungal development. Subsequently, bioinformatic analysis identified candidate ZEN-detoxification enzymes. The Baeyer–Villiger monooxygenases and carboxylesterases, which are responsible for the formation and subsequent hydrolysis of a new ZEN lactone, respectively, were significantly upregulated. In addition, the expression levels of genes related to conjugation and transport involved in the xenobiotic detoxification pathway were significantly upregulated. Moreover, the expression levels of genes encoding enzymatic antioxidants and those related to growth and apoptosis were significantly upregulated and downregulated, respectively, which made it possible for A. mycotoxinivorans to survive in a highly toxic environment and efficiently detoxify ZEN. This is the first systematic report of ZEN tolerance and detoxification in A. mycotoxinivorans. We identified the metabolic enzymes that were potentially involved in detoxifying ZEN in the GMU1709 strain and found that ZEN-induced transcriptional regulation of genes is key to withstanding highly toxic environments. Hence, our results provide valuable information for developing enzymatic detoxification systems or engineering this detoxification pathway in other species.
Article
Full-text available
The present study was aimed to evaluate the bio-control efficacy of Pediococcus pentosaceus isolated from traditional fermented dairy products originated from India, against the growth and zearalenone (ZEA) production of Fusarium graminearum. The cell-free supernatants of P. pentosaceus (PPCS) were prepared and chemical profiling was carried out by GC-MS and MALDI-TOF analysis. Chemical profiling of PPCS evidenced that, the presence of phenolic antioxidants, which are responsible for the antifungal activity. Another hand, MALDI-TOF analysis also indicated the presence of antimicrobial peptides. To know the antioxidant potential of PPCS, DPPH free radical scavenging assay was carried out and IC 50 value was determined as 32 ± 1.89 µL/mL. The antifungal activity of P. pentosaceus was determined by dual culture overlay technique and zone of inhibition was recorded as 47 ± 2.81%, and antifungal activity of PPCS on F. graminearum was determined by micro-well dilution and scanning electron microscopic techniques. The minimum inhibitory concentration (MIC) of PPCS was determined as 66 ± 2.18 µL/mL in the present study. Also a clear variation in the micromorphology of mycelia treated with MIC value of PPCS compared to untreated control was documented. Further, the mechanism of growth inhibition was revealed by ergosterol analysis and determination of reactive oxygen species (ROS) in PPCS treated samples. The effects of PPCS on mycelial biomass and ZEA production were observed in a dose-dependent manner. The mechanism behind the suppression of ZEA production was studied by reverse transcriptase qPCR analysis of ZEA metabolic pathway genes (PKS4 and PKS13), and results showed that there is a dose dependent down-regulation of target gene expression in PPCS treated samples. The results of the present study were collectively proved that, the antifungal and ZEA inhibitory activity of PPCS against F. graminearum and it may find a potential application in agriculture and food industry as a natural bio-controlling agent.
Article
Full-text available
Fungi belonging to Fusarium genus can infect crops in the field and cause subsequent mycotoxin contamination, which leads to yield and quality losses of agricultural commodities. The mycotoxin zearalenone (ZEN) produced by several Fusarium species (such as F. graminearum and F. culmorum) is a commonly-detected contaminant in foodstuffs, posing a tremendous risk to food safety. Thus, different strategies have been studied to manage toxigenic pathogens and mycotoxin contamination. In recent years, biological control of toxigenic fungi is emerging as an environment-friendly strategy, while Trichoderma is a fungal genus with great antagonistic potentials for controlling mycotoxin producing pathogens. The primary objective of this study was to explore the potentials of selected Trichoderma isolates on ZEN-producing F. graminearum, and the second aim was to investigate the metabolic activity of different Trichoderma isolates on ZEN. Three tested Trichoderma isolates were proved to be potential candidates for control of ZEN producers. In addition, we reported the capacity of Trichoderma to convert ZEN into its reduced and sulfated forms for the first time, and provided evidences that the tested Trichoderma could not detoxify ZEN via glycosylation. This provides more insight in the interaction between ZEN-producing fungi and Trichoderma isolates.
Article
Full-text available
The aim of the study was to neutralize zearalenone by lactic acid bacteria (LAB) such as Lactococcus lactis and Bifidobacterium sp. and investigate the mechanism of zearalenone (ZEA) binding. Neutralization of ZEA by LAB was confirmed by identification of binding kinetics and spectroscopic studies such as Fourier transform infrared spectroscopy (FT-IR) and matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF-MS). The obtained results showed that the kinetic process of zearalenone binding to L. lactis is not homogeneous but is expressed with an initial rapid stage with about 90% of ZEA biosorption and with a much slower second step. In case of Bifidobacterium sp., the neutralization process is homogeneous; the main stage can be described with about 88% of ZEA biosorption. MALDI–TOF-MS measurements and FTIR analysis confirmed the uptake of zearalenone molecules by bacterial species. Moreover, the assessment of dead and live lactic acid bacteria cells after zearalenone treatment was performed using fluorescence microscopy. Graphical abstractMicrobiology neutralization of zearalenone using Lactococcus lactis and Bifidobacterium sp. was confirmed by identification of binding kinetics and spectroscopic studies such as FT-IR spectroscopy and MALDI-TOF-MS spectrometry. The mechanism of ZEA binding was also investigated.
Article
Full-text available
Fumonisin B 1 (FB 1) is among the most common mycotoxins found in maize kernels and maize products worldwide. The microbiological process of detoxification and transformation of toxic organic pollutants is a promising method for foodstuffs decontamination. Some basidiomycetes, such as the Pleurotus eryngii species complex, include several important commercial edible varieties that can detoxify polycyclic organic compounds and a range of wastes and pollutants. We investigated the potential role of P. eryngii, one of the most consumed mushrooms, in the decontamination of FB 1 in maize. In addition, selected antioxidant enzymes, (soluble peroxidase (POD), cata-lase (CAT) and ascorbate peroxidase), primarily involved in control of cell hydrogen peroxide levels, and lignin degradation, were analyzed, to evaluate their contributions to the molecular mechanisms of FB 1 by P. eryngii. FB 1 decontamination by P. eryngii and involvement of CAT and POD enzymes in the control of toxic decontamination levels of H 2 O 2 were demonstrated. A consistent reduction of FB 1 was observed at different incubation times. The average decrease levels of FB 1 , with respect to the control cultures, ranged from 45 to 61% (RSD < 15%). This study is a possible eco-friendly approach to reducing this mycotoxin in the feed supply chains.
Article
Full-text available
Lactic acid bacteria commonly named as probiotics have a broad range of utilization area in human diet and food industry. Besides these known properties of probiotics, in recent years many researchers have focused on their anti-genotoxic called DNA-bioprotective effects. Human gets exposure with genotoxins such as mycotoxins, polycyclic aromatic hydrocarbons, and n-nitroso-compounds through diet and environmental contaminations. Aflatoxin B1 is mycotoxin which was reported one of the most potent hepato-carcinogen and its exposure stems from human diet. In this study, we aimed to investigate DNA-bioprotective effect of Lactobacillus plantarum on human colon adenocarcinoma (Caco-2) cells against Aflatoxin B1 (AFB1) with comet assay without metabolic activation. The results showed that DNA-bioprotective effect of L. plantarum did reduce the AFB1’s genotoxic effect on colon adenocarcinoma (Caco-2) cells. Positive control (50 ìM H2O2 applied) and high dose (>10ppm) AFB1 applied cells have the same comet tail appearance. Only visual scoring is performed. Besides negative control cells (Only PBS) and Lactobacillus plantarum+AFB1 mixture showed the same manner with each other; no comet tail detected. These results clearly indicate that L. plantarum is capable of reduce AFB1 safely without producing any by-products.
Article
Full-text available
Background: Fumonisin B1 (FB1 ) is a mycotoxin produced by several Fusarium species and is a very common contaminant of maize-based food and feed throughout the world. The selection and use of FB1 -degrading microorganisms appears as a promising alternative to cope with the problem of toxicity towards humans and livestock. High moisture maize grain silage, which is based on natural maize fermentation, could be an interesting reservoir of such microorganisms. Results: Using an in vitro simulated silage model with FB1 naturally contaminated grains, we demonstrated a significant raw decrease in FB1 during ensiling process ascribed to biodegradation mechanisms. A panel of 98 bacteria and yeasts were isolated from this matrix and selected for their ability to use FB1 as the sole source of C and N. For nine of them, the ability to degrade FB1 in vitro was evidenced. Notably, two bacteria identified as Lactobacillus sp. were highlighted for their efficient FB1 -degrading capacity and production of hydrolysed FB1 as intermediate degradation metabolite. Conclusion: Fermentation of high moisture maize grain contaminated with FB1 leads to a significant reduction of the toxin and allows the isolation of FB1 -degrading microorganisms that could further be used as FB1 decontaminating agents. © 2016 Society of Chemical Industry.
Article
Full-text available
The worldwide contamination of feeds and foods with mycotoxins is a significant problem. Mycotoxins pose huge health threat to animals and humans. As well, mycotoxins bring enormous economic losses in food industry and animal husbandry annually. Thus, strategies to eliminate or inactivate mycotoxins in food and feed are urgently needed. Traditional physical and chemical methods have some limitations such as limited efficacy, safety issues, losses in the nutritional value and the palatability of feeds, as well as the expensive equipment required to implement these techniques. Biological degradation of mycotoxins has shown promise because it works under mild, environmentally friendly conditions. Aflatoxin (AF), zearalenone (ZEA) and deoxynivalenol (DON) are considered the most economically important mycotoxins in terms of their high prevalence and significant negative effects on animal performance. Therefore, this review will comprehensively describe the biological degradation of AF, ZEA and DON by microorganisms (including fungi and bacteria) and specific enzymes isolated from microbial systems that can convert mycotoxins with varied efficiency to non- or less toxic products. Finally, some strategies and advices on existing difficulties of biodegradation research are also briefly proposed in this paper.
Article
Full-text available
Food and feedstuff contamination with aflatoxins (AFTs) is a serious health problem for humans and animals, especially in developing countries. The present study evaluated antifungal activities of two lactic acid bacteria (LAB) against growth and aflatoxin production of toxigenic Aspergillus parasiticus. The mycelial growth inhibition rate of A. parasiticus PTCC 5286 was investigated in the presence of Bifidobacterium bifidum PTCC 1644 and Lactobacillus fermentum PTCC 1744 by the pour plate method. After seven days incubation in yeast extract sucrose broth at 30°C, the mycelial mass was weighed after drying. The inhibitory activity of LAB metabolites against aflatoxin production by A. parasiticus was evaluated using HPLC method. B. bifidum and L. fermentum significantly reduced aflatoxin production and growth rate of A. parasiticus in comparison with the controls (p≤0.05). LAB reduced total aflatoxins and B 1 , B 2 , G 1 and G 2 fractions by more than 99%. Moreover, LAB metabolites reduced the level of standard AFB 1 , B 2 , G 1 and G 2 from 88.8% to 99.8% (p≤0.05). Based on these findings, B. bifidum and L. fermentum are recommended as suitable biocontrol agents against the growth and aflatoxin production by aflatoxigenic Aspergillus species.
Article
Full-text available
Sorghum is a food and feed crop grown commonly in Asian and African countries. Aflatoxin contamination of sorghum is a common problem in all tropical countries. Since even low doses of aflatoxin intake, over a period of time, can prove to be carcinogenic, removal of aflatoxin from sorghum is necessary. Biodecontamination of food grains is now gaining popularity since it is considered to be a safer alternative to chemical methods. Probiotic organisms such as lactic acid bacteria (LAB), have been commonly used in fermentation of various food products. Studies have been done to prove the ability of the probiotic organisms to bind and remove aflatoxin from food. Here the ability of four strains of lactic acid bacteria (Lactobacillus plantarum, Lactobacillus rhamnosus, Lactobacillus brevis, and Lactobacillus delbrueckii subsp. lactis) to reduce aflatoxin levels in sorghum during fermentation has been studied. The initial and final aflatoxin content in sorghum was analyzed using HPLC and UV spectrophotometer. It was concluded that L. delbrueckii had the maximum antiaflatoxigenic ability. It showed 74.38% reduction of aflatoxin in Sorghum bicolor. The research outcome concluded that L. delbrueckii may be used as potential probiotic for reducing the aflatoxin content in sorghum.
Article
Aflatoxin B1 (AFB1) and zearalenone (ZEA) are the secondary toxic metabolites of fungi which contaminate a wide range of food and feedstuffs. Limiting exposure of humans and livestock to them is very essential. Among numerous methods of mycotoxin-degradation, biodegradation by microorganisms and enzymes is an effective and promising approach to eliminate their hazards. The present study aims to optimize the proportion of different species of beneficial microbes by means of response surface methodology (RSM) and its combination with mycotoxin-degradation enzymes. The results indicated that AFB1 and ZEA degradation rates were 38.38% and 42.18% by individual Bacillus subtilis (P < 0.05); however, AFB1 and ZEA degradation rates reached 45.49% and 44.90% (P < 0.05) when three probiotic species such as Bacillus subtilis, Lactobacillus casein and Candida utilis were at a ratio of 1:1:1, corresponding with the predictive value of the RSM model. The further experiment showed that AFB1 and ZEA degradation rates were 63.95% and 73.51% (P < 0.05) when the compound of three probiotic species was combined with mycotoxin-degradation enzymes from Aspergillus oryzae at a ratio of 3:2. This result indicated that the combination of probiotics with mycotoxin-degradation enzymes is a promising new approach for synchronous detoxification of AFB1 and ZEA.