Next Article in Journal
Chronodisruption: A Poorly Recognized Feature of CKD
Previous Article in Journal
Variation of Fusarium Free, Masked, and Emerging Mycotoxin Metabolites in Maize from Agriculture Regions of South Africa
Previous Article in Special Issue
Function of crzA in Fungal Development and Aflatoxin Production in Aspergillus flavus
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Aflatoxin Biosynthesis and Genetic Regulation: A Review

1
Toxalim (Research Center in Food Toxicology), Université de Toulouse, INRAE, ENVT, EI-Purpan, 31300 Toulouse, France
2
Centre d’analyse et de recherche, Unité de recherche technologies et valorisations agro-alimentaires, Faculté des Sciences, Université Saint-Joseph, P.O. Box 17-5208, Mar Mikhael Beirut 1104, Lebanon
3
Laboratory of Microbiology, Department of Life and Earth Sciences, Faculty of Sciences I, Lebanese University, Hadath Campus, P.O. Box 5, Beirut 1104, Lebanon
*
Author to whom correspondence should be addressed.
Authors equally contributed to this work.
Submission received: 19 December 2019 / Revised: 27 January 2020 / Accepted: 25 February 2020 / Published: 28 February 2020
(This article belongs to the Special Issue Production Mechanisms and Biosynthesis of Aflatoxin)

Abstract

:
The study of fungal species evolved radically with the development of molecular techniques and produced new evidence to understand specific fungal mechanisms such as the production of toxic secondary metabolites. Taking advantage of these technologies to improve food safety, the molecular study of toxinogenic species can help elucidate the mechanisms underlying toxin production and enable the development of new effective strategies to control fungal toxicity. Numerous studies have been made on genes involved in aflatoxin B1 (AFB1) production, one of the most hazardous carcinogenic toxins for humans and animals. The current review presents the roles of these different genes and their possible impact on AFB1 production. We focus on the toxinogenic strains Aspergillus flavus and A. parasiticus, primary contaminants and major producers of AFB1 in crops. However, genetic reports on A. nidulans are also included because of the capacity of this fungus to produce sterigmatocystin, the penultimate stable metabolite during AFB1 production. The aim of this review is to provide a general overview of the AFB1 enzymatic biosynthesis pathway and its link with the genes belonging to the AFB1 cluster. It also aims to illustrate the role of global environmental factors on aflatoxin production and the recent data that demonstrate an interconnection between genes regulated by these environmental signals and aflatoxin biosynthetic pathway.
Key Contribution: Based on bibliographic data; this review provides an overview of the factors interacting with the aflatoxin gene cluster and hence aflatoxin production in Aspergillus species.

1. Introduction

Aflatoxins are toxic secondary metabolites produced by several fungal species mostly belonging to Aspergillus section Flavi [1]. Toxigenic fungi in this section synthesize four main molecules, aflatoxin B1, B2, G1 and G2, all of which can yield other derivatives when metabolized by animals after their ingestion. Aflatoxin B1 (AFB1) is one of the most important compounds due to its demonstrated carcinogenic properties in human and its frequent presence in many foodstuffs worldwide [2]. Studies demonstrated that chronic exposure to AFB1 can lead to numerous diseases including immune suppression in humans and animals, malabsorption of nutrients, infertility, endocrine problems as well as teratogenic effects related with congenital malformations and hepatocellular carcinoma [3]. Moreover, it has been recently demonstrated that intermediates precursors compounds within the aflatoxin biosynthetic pathway, such as versicolorin A, could also represent a potential risk due to their cytotoxic effects [4].
It was long assumed that AFB1 contamination was a major public health issue in tropical and subtropical regions because the climate may favor the development of aflatoxigenic species in the field or during storage. However, with ongoing global climate changes, AFB1 is predicted to be an emerging threat in areas where it was not previously present [5,6]. Such is the case of several regions of Europe [7,8]. The genotoxic property of AFB1 justifies limiting consumer exposure to this toxin to the greatest extent possible as reflected in worldwide regulations allowing only a few µg of the toxin per kg of food [9]. To reach such ambitious objectives, many preventive strategies have been developed to reduce AFB1 occurrence in food commodities. These strategies range from good agricultural practices to the use of biocontrol agents or natural compounds able to block toxin production [10]. Even if the exact mechanism of action has not yet been completely elucidated, a few studies already demonstrated that some of these natural products can inhibit AFB1 production by a transcriptional down-regulation of the genes involved in AFB1 synthesis [11,12,13,14,15,16,17,18]. Aflatoxins are the product of a complex biosynthesis pathway involving at least 27 enzymatic reactions [19,20,21,22,23,24]. The genes coding for these enzymes are grouped in a cluster and their expression is coordinated by two cluster-specific regulators: aflR and aflS [25,26]. Nevertheless, as a secondary metabolite, AF synthesis also depends on other complex mechanisms triggered in response to environmental stimuli including pH, light, nutrient sources and oxidative stress response, which may activate different cell signaling pathways resulting in the modulation of the expression of genes involved in toxin production [27,28,29,30,31]. The recent development of molecular tools in the field of fungal physiology allowed the demonstration of the interaction between many genes involved in response to environmental stimuli and the AF cluster, even if the exact level of connection is often not completely elucidated. Understanding the connection between the AF cluster and environmental stimuli may help to define new strategies to limit mycotoxin production by specifically targeting genes involved upstream of the cluster of toxins.
As one of the most hazardous mycotoxins, there are many reviews concerning different aspects of aflatoxin biosynthesis [29,32,33,34,35,36,37]. Thus, the present review is timely and provides an overview of the different factors interacting with the AFB1 gene cluster and hence with toxin production. However, an effort to link these factors in order to synthesize the global network regulating aflatoxin production is proposed. Data on Aspergillus flavus and A. parasiticus, the two main producers of aflatoxins in crops, are presented. Studies of A. nidulans linked with sterigmatocystin production are included since this mycotoxin is the penultimate stable intermediate in the AFB1 synthesis cascade and its cluster shares 25 homologous genes with the AF cluster [24]. After reminding the AFB1 biosynthetic pathway and its internal gene regulation, we describe other genes found to interact with AFB1 synthesis according to their role in fungal metabolism. These genes are divided into several categories: cell signaling, reproductive process, growing conditions and oxidative stress as shown in Figure 1.

2. Aflatoxin Biosynthetic Pathway

2.1. The Aflatoxin Gene Cluster

In Aspergilli, DNA information is organized in 8 chromosomes where genes responsible for aflatoxin production are located in the 54th cluster, 80 kb from the telomere of chromosome 3 [29]. This cluster includes 30 genes and its activation is mainly regulated by aflR and aflS [25,26] (Figure 2).
The aflatoxin gene cluster has been widely studied in A. flavus and A. parasiticus. The homology of the clustered genes between the two fungal species is 90–99% [38]. One of the main differences between the two species is their ability to produce B and G type aflatoxins. A. flavus mainly produces B type AFs (AFB1 and AFB2) whereas A. parasiticus produces B and G type molecules (AF-B1, -B2, -G1 and -G2). Functional genes involved in G-type aflatoxin production correspond to aflU, aflF and nadA respectively coding for a cytochrome P-450 monooxygenase, an aryl alcohol dehydrogenase and an oxidase [39,40]. Understanding the AF cluster has also benefited from experiments on A. nidulans regarding its capacity to produce sterigmatocystin. Indeed, the homology between A. parasiticus and A. nidulans clusters is 55–75% [38].

2.2. Enzymatic Cascade Leading to Aflatoxin Synthesis

Aflatoxins are produced by a polyketide pathway that was first proposed by Birch in 1967 [41] and today, at least 27 enzymatic reactions have been shown to be involved in this process. It has long been considered that, among natural secondary metabolites, aflatoxin biosynthesis is one of the longest and most complex processes due to the quantity of oxidative rearrangements it includes [42]. Three critical oxygen elements were characterized in this pathway by Dutton in 1988 [43]:
(i)
Monooxygenases: responsible for incorporating one oxygen atom in another being reduced, with nicotinamide adenine dinucleotide phosphate (NADPH) acting as a cofactor.
(ii)
Dioxygenases: involved in ring-cleavage reactions.
(iii)
Baeyer-Villiger reactions: responsible of inserting oxygen atoms between two carbons.
Cytochromes P-450 also play a key role in aflatoxin synthesis. These enzymes are involved in attaching functional groups (i.e., methyl, acetyl) during biosynthesis [44]. Indeed, aflatoxin gene cluster contains the highest number of cytochromes P-450 among the mycotoxin biosynthetic pathways known to date [45].
In order to schematize the principal enzymatic reactions occurring during AFB1 production, Figure 3 shows the various intermediates in the biosynthesis of AFB1 as well as the genes present in the aflatoxin gene cluster responsible for encoding the enzymes for their synthesis.
To remind the specific characteristics and interventions of the genes participating in the AFB1 biosynthesis pathway, it will be presented in four paragraphs according to the stable metabolites that appear during AFB1 production: norsoloric acid, versicolorin A, sterigmatocystin and finally, aflatoxin B1. Genes are indicated with new and former (in brackets) nomenclatures.

2.3. Conversion of Acetate into Norsolorinic Acid

Aflatoxins are polyketide-derivates requiring the formation of hexanoate units (from acetyl-CoA and malonyl-CoA) to start the cascade reaction [23]. The first steps in the pathway lead to the transformation of the starter unit into the first stable metabolite norsolorinic acid (NOR). The enzymes involved in these reactions are mainly encoded by four genes:
- aflA (fas-2) and aflB (fas-1) formerly named “fas” genes as they code for fatty acid synthases. Their corresponding synthesized proteins are α and β sub-units primarily responsible for transforming the hexanoate units into a polyketide structure [23,56].
- aflC (pksA) is a gene coding for the synthesis of polyketide skeletons. In general, secondary metabolites that are acetate derivatives, as is the case of Aflatoxin B1, are subject to chain elongation induced by this polyketide synthase (e.g., from 2 MalonylCoA to 7 MalonylCoA). In addition, this enzyme is involved in further transformations of the polyketide structure into norsolorinic acid anthrone (NAA) [23].
- hypC (hypB1) is a gene located between aflC and aflD. It encodes for a 17-kDa enzyme that was demonstrated to be involved in the catalytic conversion of NAA into NOR [57].

2.4. Conversion of Norsolorinic Acid into Versicolorin A

Norsolorinic acid is further transformed into averantin (AVN), in a step governed by the aflD gene. For years, the involvement of two other genes (aflE and aflF) was associated with this step. However, further studies provided evidence for their implication in other steps in the AFB1 enzymatic cascade, and their role is described later.
- aflD (nor-1) encodes a norsolorinic acid ketoreductase needed for the conversion of the 1′-keto group of NOR to the 1′-hydroxyl group of AVN [58]. In the 1990s, Yabe et al. [59] identified, in A. parasiticus, two enzymatic reactions involved in the conversion of averantin into averufin (AVF). This conversion first involves a P-450 monooxygenase responsible for converting the AVN into 5′-hydroxyaverantin (HAVN) and second, the transformation of HAVN into averufin by an alcohol dehydrogenase. The transformation from HAVN into AVF was originally thought to occur in one step. However, Sakuno et al. [50] discovered that an intermediate compound identified as 5′-oxoaverantin (OAVN) is generated during the process. The genes responsible for the above enzymatic processes are:
- aflG (avnA), encoding a cytochrome P-450 monooxygenase that catalyzes the hydroxylation of the polyketide anthraquinone averantin into 5′ -hydroxyaverantin [56,60,61].
- aflH (adhA), coding for the alcohol dehydrogenase that is needed for the conversion of HAVN into 5′-oxoaverantin (OAVN) [62].
- aflK (vbs) was firstly associated with the conversion of versiconal into versicolorin B but was also reported to be responsible for the transformation of OAVN into AVF [47]. Indeed, when Sakuno et al. [47] elucidated the intermediate metabolite between HAVN and AVF, the 5′-oxoaverantin (OAVN), this transformation was also associated with the aflK gene. These authors were the first to demonstrate that the same enzyme can catalyze two different reactions in the AFB1 pathway. They hypothesized that this phenomenon is due to the evolution of the AFB1 gene cluster that may previously have comprised two copies of the aflK gene.
Further conversion from averufin into versicolorin B is governed by different genes:
- the aflI (avfA) gene is involved in the transformation of averufin (AVF) into versiconal hemiacetal acetate (VHA). Indeed, the deletion of this gene led to the accumulation of AVF [38] and it is generally assumed that the enzyme encoded by aflI catalyzes the ring-closure step during the formation of hydroxyversicolorone (HVN). Nevertheless, the exact role of AflI in the oxidation of AVF has never been elucidated [24].
- aflV (cypX) and aflW (moxY) encode a P-450 microsomal monooxygenase and a cytosolic monooxygenase, respectively [48]. AflV has been shown to catalyze the reaction from AVF to HVN and AflW to catalyze the transformation of HVN to VHA by a Baeyer–Villiger reaction.
The further transformation from VHA into versiconal (VAL) is governed by the gene aflJ, after which aflK is responsible for the transformation of VAL into versicolorin B (VERB).
- aflJ (estA) encodes an esterase catalyzing the conversion of VHA to versiconal (VAL). Interestingly, this enzyme has been reported to allow the reversible transformation of VHA into versiconol acetate (VOAc) and from versiconol (VOH) to VAL, aflJ being therefore responsible for both reactions during AFB1 biosynthesis [51,63].
- aflK (vbs) encodes a cyclase involved in the transformation of VAL into VERB [64] but also, as already mentioned, during the transformation of OAVN into AVF. The coded enzyme is responsible for bisfuran ring closure in aflatoxins, which, in turn, is responsible for DNA binding after metabolization and the mutagenic effect of aflatoxins [33].
Finally, in A. nidulans, the transformation of VERB into versicolorin A (VERA) is presumed to be done by the product of the aflL gene, coding for a cytochrome P-450 monooxygenase/desaturase [65]. The main characteristic of this step refers to the final metabolic transformation before the principal branch leading to the synthesis of B or G-type aflatoxins [46].

2.5. Conversion of Versicolorin A into Sterigmatocystin

Four genes are involved in the transformation of VERA into demethylsterigmatocystin (DMST):
- aflM (ver-1), predicted to encode a ketoreductase, is involved in the conversion of VERA into DMST [52] and its homologous gene in A. nidulans has been also identified [66]. Further studies demonstrated that the promoter of this gene contains cAMP-response element sites (CRE) that are also present in genes involved in fungal oxidative stress response [67]. However, the exact intervention of this enzyme has not been yet completely elucidated.
- aflN (verA) codes for a cytochrome P450-type monooxygenase which exact function also remains to be determined. Some studies suggest that this protein could be involved in the passage of VERA into a hypothetical intermediate [52]. Its homolog in A. nidulans corresponds to stcS [68].
- aflY (hypA) seems to mostly intervene between two hypothetical intermediate structures between the transformation of VERA into DMST by a Baeyer–Villiger reaction. Disrupted A. parasiticus strain ΔaflY accumulated VERA, which is the main reason to include this gene in the VERB-DMST group [54].
- aflX (ordB) encodes an oxidoreductase that catalyzes the oxidative decarboxylation and ring-closure of the Baeyer–Villiger intermediate resulting from the AflY-catalyzed oxidation [53].
- aflO (omtB), which codes for an O-methyltransferase, is involved in the conversion of DMST into sterigmatocystin (ST), the penultimate intermediate in the AFB1 pathway and the final metabolite in A. nidulans [38].

2.6. Conversion of Sterigmatocystin into Aflatoxin B1

- aflP (omtA) codes for one of the enzymes identified in A. parasiticus as an O-methyltransferase. It is one of the main genes responsible for transforming ST into O-methylsterigmatocystin (OMST) [69]. This gene has been shown to be expressed only in aflatoxin permissive conditions [70].
The final conversion of OMST into AFB1 is governed by the following genes: aflQ, hypB, aflE and hypE. During this final transformation, the intervention of aflQ and hypB has been precisely identified while only partial information is reported for the other genes.
- aflQ (ordA) an adjacent gene to aflP in the AFB1 cluster codes for a P-450 monooxygenase. This gene is involved in the conversion of OMST into AFB1 by the oxidation of the A-ring of OMST [24,71]. This reaction leads to an AFB1 precursor named 11-hydroxy-O-methylsterigmatocystin (HOMST) [55].
- hypB (hypB2) is a gene coding for an oxidase involved in the second transformation step from HOMST into a 370 Da 7-ring lactone and expressed under aflatoxin permissive conditions [57]. The transformation of this compound into another unknown intermediate is hypothesized to be catalyzed by hydrolytic enzymes encoded by genes that do not belong to the AF cluster [24].
- aflE (norA) is a homologous gene of aflD in the aflatoxin cluster encoding a short-chain aryl alcohol dehydrogenase [22]. The first characterization of aflE in A. parasiticus suggested that this gene was involved in the transformation of NOR to AVN [72]. However, further studies by Ehrlich et al. [73] in A. flavus, demonstrated that aflE was mainly involved in the two final steps of AFB1 formation (even if the exact position has not yet been identified) since the absence of AflE led to the accumulation of deoxyaflatoxin.
- hypE (aflLa) is a gene responsible for the final steps of AFB1 production since its disruption leads to an intermediate compound before deoxyAFB1 synthesis. HypE presents homologies with several bacterial enzymes due to its ethD domain and it was thus suggested that, together with a P-450 monooxygenase, it probably interacts in the aflatoxin enzymatic pathway [24].
As demonstrated, AFB1 production is a long and complex process corresponding to a coordinated enzymatic cascade in which at least 27 enzymes coded by genes belonging to the AFB1 cluster are involved. Nevertheless, to date, the roles of several other genes, also belonging to the cluster, remain unclear.

3. Genes Present in the AFB1 Cluster with an Unclear Role in the Aflatoxin Enzymatic Cascade

Within the AFB1 gene cluster, two genes have not been directly associated with the biosynthetic pathway:
- aflT is a gene coding for a fungal transporter that belongs to the major facilitator superfamily (MFS). This gene was widely characterized by Chang et al. [74] who demonstrated that even if it resides in the AF gene cluster, its role is not directly linked to aflatoxin biosynthesis. In fact, its deletion in A. parasiticus did not affect the final amounts of aflatoxins compared to the control. These results are in agreement with others reporting that aflT is neither regulated by the main activator of the AFB1 pathway aflR, nor by aflS but instead by an external factor, FadA, a sub-unit of the G-protein signaling pathway [74]. In the same study, experiments using S. cerevisiae yeast led the authors to suggest that aflT is not implicated in the transport of aflatoxin. However, Chanda et al. [75] reported that AflT resides in the aflatoxisomes, which are implicated in the exocytose of aflatoxins. So, the exact role of aflT in relation with aflatoxin production is still not clear.
- hypD (aflNa) coding for a 129 Da integral membrane-binding protein is another gene whose role is still not clear. Ehrlich [24] reported that deletion of hypD in an A. parasiticus strain resulted in increased sporulation of cultures along with reduced levels of AFB1 production, suggesting that HypD is involved in both fungal development and toxin production [15,18].

4. Aflatoxin Cluster-Specific Regulators

4.1. The aflR Transcription Factor

In A. flavus, A. parasiticus and A. nidulans, the aflatoxin and sterigmatocystin biosynthetic pathways are mainly regulated by the aflR gene [31,76]. In A. flavus, the AflR protein binds to at least 17 of the genes in the AF cluster that results in the activation of the enzymatic cascade and leads to the production of different AFs. AflR is classified as a zinc cluster Zn(II)2Cys6 transcription factor of the Gal4-type family [77]. This type of transcription factor has a specific structure that is only attributed to the fungi kingdom. Such transcription factors are able to bind to DNA using a DNA-binding domain, which is one of the most important elements in transcriptional and translational processes [78]. In fact, the regulation process occurs in the nucleus. Once there, the zinc finger parts of AflR bind to DNA and it is believed that this binding occurs in a homodimer manner [78,79]. In particular, AflR was demonstrated to preferentially recognize the palindromic pattern 5′-TCG(N5)CGA-3′, but it also binds to other sequences such as 5‘-TTAGGCCTAA-3′ and to the 5′-TCGCAGCCCGG-3′ pattern [36,80]. Interestingly, aflR conserves the palindromic pattern 5′-TTAGGCCTAA-3′ within its own promoter, suggesting that, apart from being the major modulator of the AF gene cluster, aflR could also be auto-regulated [26,81]. aflR acts as a positive regulator of the AF gene cluster. In A. flavus, over-expression of aflR up-regulates several AF genes thereby increasing the production of aflatoxin 50-fold [82]. In the same way, aflR deletion in an O-methylsterigmatocystin (OMST)-accumulating strain of A. parasiticus, had a negative effect on aflatoxin pathway genes and hence on the production of OMST [83]. Recently, it has been demonstrated that different inhibitors of aflatoxin production are capable to down-regulate aflR expression [12,13,17,84,85,86]. Finally, it has been demonstrated that even if aflR is the main activator of the aflatoxin gene cluster, it interacts with aflS that plays the role of enhancer in the regulatory biosynthesis process [25].

4.2. The aflS Transcription Enhancer

aflS (previously named aflJ) was characterized by Meyers and co-workers [87] and was shown to be required for aflatoxin synthesis. In A. parasiticus, this gene encodes a 438-amino acid protein that has no important homology with any other proteins found in databases [22]. The aflS gene is located adjacent to aflR in the AF biosynthetic cluster, and aflS was also shown to be regulated by aflR [24]. Both genes share a 737-bp intergenic region from their translational starting sites [22]. In A. flavus and A. parasiticus, it was also demonstrated that aflS interacts with aflR but not with the biosynthetic enzymes, revealing the co-active function of aflS [25,88]. Disruption of aflS in A. parasiticus resulted in mutants with a 5- to 20-fold reduction in the expression of some AFB1 genes, such as aflC, aflD, aflM and aflP, and a subsequent loss in the ability to synthesize AF intermediates [25]. Otherwise, its overexpression in A. flavus resulted in higher levels of AFB1 production with 4- to 5-fold increase in the level of aflC and aflD expression as well as in averantin synthesis [88]. Even if it is still not clear how aflS increases the transcription levels of the genes involved in the AF pathway, it has been demonstrated that a dimer can be formed between both corresponding proteins in order to activate the AF gene cluster. This mechanism was proposed by Du et al. [88] who reported that this dimer-complex could recognize specific sites in the promoter regions of aflC and aflD (involved in early steps in the AFB1 biosynthesis) increasing their transcription and hence AFB1 production. Concerning the protein-dimer formation, it was further hypothesized that when both genes (aflR and aflS) are expressed normally, a protein complex made of 4 AflS for 1 AflR is formed, allowing correct binding to the promoter regions of the aflatoxin cluster genes [89]. Previous studies demonstrated that under certain AFB1 inhibitory conditions, a marked decrease in cluster gene expression and AFB1 production was observed with only low-level changes in aflR expression [90]. We obtained a similar result in one of our studies in which AFB1 levels were significantly reduced whereas aflR expression was not, although a major down-regulation of aflS and aflatoxin production was observed [15]. We suggested that the differential expression of aflR and aflS may have modified the ratio of available AflR and AflS proteins, leading to the formation of an insufficient number of complexes. Consequently, all the AflR-binding sites may not have been reached and the subsequent cluster transcription was not complete. The main argument supporting this hypothesis is the observation that genes encoding enzymes involved in the final stages of the AFB1 enzymatic cascade were more impacted than those involved in the initial steps. Therefore, the limited number of AflR/AflS complexes available might have been rapidly used at the beginning of AFB1 synthesis and were no longer available to properly activate the last cluster genes [18]. Even if great progress has been made in the characterization of aflS, the exact mechanism of action by which this gene modulates transcription of the aflatoxin pathway is still under investigation.

5. Global Regulation of Aflatoxin Production

Global regulatory factors here refer to genes that do not belong to the aflatoxin gene cluster but that have been demonstrated to have a link with aflatoxin production. In fact, aflatoxin biosynthesis can be modulated by different genes, external to the cluster, and that are themselves activated or repressed by environmental stimuli. Some of these influencing environmental stimuli are nutrient sources, exposure to environmental changes as well as oxidative fungal response. Many recent works demonstrated an interaction between the genes involved in the response to these environmental factors and the genes belonging to the aflatoxin cluster, even if the mechanisms of such interactions are often not completely understood. This is mainly due to the complexity of the pathways involved and their multiple interconnections. In this part of the review, to describe the recent data demonstrating links between global regulatory factors and AF production, genes interfering with aflatoxin production are classified in several groups based on their demonstrated function or role in cellular metabolism. The data are presented in summary tables including the genes involved, their known role in fungal metabolism and the works reporting their link with AF synthesis. When available, data obtained in A. nidulans and related to ST production are also included.

5.1. Growing Conditions

Growing conditions are one of the factors that have the most influence on the production of fungal secondary metabolites. They are highly dependent on many environmental variables. For instance, temperature and water activities are key parameters for fungal development and toxinogenesis [91,92] but to date the specific genes involved in such response and their link with aflatoxin cluster has not been described. However, it has been reported that several genes belonging to the aflatoxin cluster including the principal regulators aflR and aflS can be modulated by changes in temperature and water activity demonstrating a close link between these external factors and aflatoxin production [91]. Other environmental parameters including nutrient sources, pH and light exposure were found to interfere directly with aflatoxin production [31,93] and the role of some genes has been highlighted in such phenomenon.

5.1.1. Nutrient Sources

Carbon Source

It has long been known that the availability and type of carbon source can modulate the number of secondary metabolites. Sugars are the favored carbohydrates for AFB1 production since they generate polyketide starter units (e.g., Acetyl-CoA) [94,95]. Simple sugars (e.g., sucrose, glucose, fructose, sorbitol) used as carbon sources have been correlated with higher levels of AF production in A. flavus, A. parasiticus and A. nidulans [29,96,97,98]. However, the use of D-glucal (a glucose-derivative not metabolized by fungi), used as the main source of sugar in the culture medium, inhibited AFB1 production [99].
Even if glucose is the best carbohydrate for the production of aflatoxins, they can be also produced using other carbon sources (e.g., ribose, xylose, or glycerol) [94]. In Aspergillus, the utilization of a given carbon source involves the sugar cluster. It contains four genes grouped in a 7.5 kB cluster located next to the AF gene cluster in A. parasiticus and A. flavus [96,100].
Carbon catabolic repression (CCR) is a strategic mechanism used by Aspergilli to preserve energy and adjust carbon catabolism in order to use the most favorable carbon source [93,101]. During this process the transcription factor (TF) CreA, along with the genes that interact with it, are presumed to be the main regulators (Table 1). The relation with AF production is first based on the fact that several genes belonging to the AF cluster have CreA-binding sites near their promoter regions [29]. In addition, it was demonstrated that this TF is needed for several fungal functions including AFB1 production in A. flavus [102].

Nitrogen Source

In Aspergillus species, nitrogen sources can affect aflatoxin and/or ST production in different ways and are regulated by the nitrogen metabolite repression mechanism [97]. AreA plays a crucial role in the process because it is responsible for modulating genes coding for the utilization of alternative sources of nitrogen. The nitrogen source is closely linked to AF production since some substrates (e.g., asparagine, ammonium salts, glutamate) support aflatoxin production while others do not (e.g., sodium nitrate, tryptophan) [22] (Table 2).

5.1.2. pH Conditions

pH is another extracellular condition to which fungal organisms react and that strongly modulates the production of secondary metabolites [115,116]. Indeed, the influence of pH has been linked to AF/ST production in Aspergillus species where pacC seems to be the key transcription factor involved in the response to pH [117] (Table 3). In addition, the PacC signaling pathway includes six other proteins with Pal domain, involved in pH sensing [118].

5.1.3. Light

Light stimulus is another important factor for fungi since it has a broad impact on fungal adaptation and survival. Light sensing can affect the circadian clock and growth, and has an impact on several morphological features and on the production of secondary metabolites [122] (Table 4). Light response in Aspergilli was shown to strongly involve the “velvet complex” where the global regulator veA governs a great number of genetic elements including photoreceptors [123,124,125].

5.2. Reproductive Processes

5.2.1. Sexual Development

For years, it was believed that the reproductive process of A. flavus only occurred in an asexual way but then sexual reproduction was demonstrated in A. flavus and A. parasiticus. Both species were characterized as heterothallic species containing one of two mating-type genes: MAT1-1 or MAT1-2 (Table 5). In A. flavus sexual reproduction occurs within sclerotia when they recombine with the opposite mating type [142,143,144].

5.2.2. Asexual Development

Secondary metabolite production is also coordinated with fungal development which, in turn, is intrinsically linked to conidiation (asexual reproduction) [31,93,97]. In fact, an interconnection between conidiation and AF/ST production has been demonstrated and several genes including fadA, fluG, flb- genes and brlA, were shown to be involved in this phenomenon [31,36] (Table 6).

5.3. Oxidative Stress

In fungi, changes in environmental conditions can alter the normal intracellular balance between the production of reactive oxygen species (ROS) and the production of scavengers. In response to this phenomenon, several transcription factors are expressed to activate the enzymatic defenses that protect cells from excessive ROS levels and possible subsequent damage to DNA, proteins, and lipids [28]. In addition, their oxidative status has also been linked to the production of secondary metabolites and demonstrated to be a prerequisite for AFB1 production [45,164] (Table 7). In fact, it is proposed that, in A. parasiticus and A. flavus, aflatoxin production is part of the fungal oxidative stress response [67,164]. Indeed, AFs and their metabolic precursors (e.g., OMST, versicolorins, norsolorinic acid) are highly oxygenated molecules and thus subject to redox regulation [165]. AFB1 biosynthesis is activated by high levels of oxidative stress-inducing factors (e.g., lipid hydroperoxides) whereas several antioxidants (e.g., polyphenols) can block toxin production [166]. The relation between oxidative stress and secondary metabolites production has been thoroughly investigated in fungi.

5.3.1. Superoxide Dismutases and Catalases

Superoxide dismutases (SOD), catalases (CAT) and glutathione peroxidase (GPX) are key enzymes in fungal defense against ROS, thus mediating cellular defense against oxidative stress. SOD act as a first line of defense by converting the free radicals into hydrogen peroxide (H2O2) and oxygen (O2), then peroxidases and catalases convert H2O2 into H2O and/or O2 and H2O respectively [176]. Both also interact with AF production (Table 8).

5.3.2. β-oxidation

β-oxidation of fatty acids is a fungal process that degrades fatty acids into acetate units [181]. Since AFB1 biosynthesis is triggered by acetate units, this pathway may be indirectly linked to AFB1 production [95,181] (Table 9). In addition, β-oxidation of fatty acids occurs in peroxisomes in A. flavus [182]. Vacuoles and vesicles are also participating in aflatoxin production since they are involved in final steps of aflatoxin synthesis and export [75,183].

5.4. Cell Signaling

Cell signaling in fungal cells helps overcome environmental stresses by activating rapid signal transduction systems throughout the cell that enable the organism to adapt to its surroundings [30]. G-protein coupled receptors (GPCR) and oxylipins are two of the actors involved in such processes. It has been shown that regulation of G-protein signaling plays an important role in secondary metabolite production [31,186] (Table 10) and that oxylipins influence secondary metabolism at a transcriptional level [187].
Many genes involved in different cell functions were found to be able to interact directly or indirectly with aflatoxin gene cluster and thus modulate the biosynthesis of these mycotoxins. In order to identify the possible connections between these genes and understand what could be the starting point of the modulation of aflatoxin production in response to diverse environmental stimuli, a schematic representation is proposed on Figure 4.

6. Conclusions

As demonstrated, AFB1, one of the most hazardous mycotoxins contaminating foods and feeds, is produced by a very complex process involving numerous enzymes encoded by genes grouped in a cluster. The internal regulation of this process is also complex and under the direct dependence of two cluster-specific regulators (aflR-aflS) that interact together where it seems to be that the proportion of their corresponding proteins may be a key point in the activation of the gene cluster. The identification of natural products able to down modulate the expression of one of the two, without affecting fungal physiology could be a promising strategy to control aflatoxin B1 contamination of foods and feeds. It may allow avoiding toxin synthesis without any impact on biodiversity and no risk of resistance. Understanding how cluster-specific regulators are impacted by global environmental signals and the identification of genes involved is an ambitious goal. Fungal secondary metabolism is very diverse and complex and we are still at the beginning of its elucidation. The development of molecular tools will surely strongly help deciphering the role and function of numerous genes involved in secondary metabolite production. The development of mutant strains demonstrated the direct interaction of some genes primary involved in the response of environmental changes or adaptation and the aflatoxin cluster genes. These experiments were done using a targeted approach that brought useful information. It seems now important to identify the starting point of the interaction. The use of untargeted approaches such as RNAseq or micro-array techniques could help to define if there is the common function/gene that is implicated upstream the cluster in aflatoxin inhibition that can be observed after incubation of the toxigenic fungi with different inhibitors and/or in certain culture conditions. It could allow the identification of new targets to control aflatoxin production and ensure food safety using eco-friendly and sustainable means. However, to date, the still limited functional annotation of the genome of aflatoxin-producing fungal species remains a strong limit to this type of approach.

Funding

The authors thank ANSES for financial support (Project AFLAFrance, N 2019/1/016).

Acknowledgments

The authors would like to thank the Consejo Nacional de Ciencia y Tecnología (CONACYT) México funding Isaura Caceres, as well as the Agence Nationale de la Recherche (ANR), the Research Council of Saint-Joseph University (Lebanon) and the National Veterinary School of Toulouse funding Rhoda El Khoury and Anthony Al Khoury. This work has benefited to the International Cooperation Program CAPES/COFECUB (Sv 947/19).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Frisvad, J.C.; Hubka, V.; Ezekiel, C.N.; Hong, S.B.; Nováková, A.; Chen, A.J.; Arzanlou, M.; Larsen, T.O.; Sklenář, F.; Mahakarnchanakul, W.; et al. Taxonomy of Aspergillus section Flavi and their production of aflatoxins, ochratoxins and other mycotoxins. Stud. Mycol. 2019, 93, 1–63. [Google Scholar] [CrossRef]
  2. Mahato, D.K.; Lee, K.E.; Kamle, M.; Devi, S.; Dewangan, K.N.; Kumar, P.; Kang, S.G. Aflatoxins in Food and Feed: An Overview on Prevalence, Detection and Control Strategies. Front. Microbiol. 2019, 10, 1–10. [Google Scholar] [CrossRef]
  3. Bbosa, G.S.; Kitya, D.; Lubega, A.; Ogwal-Okeng, J.; Anokbonggo, W.W.; Kyegombe, D.B. Review of the Biological and Health Effects of Aflatoxins on Body Organs and Body Systems. In Aflatoxin-Recent Advances and Future Prospects; Razzaghi, A., Ed.; InterchOpen: London, UK, 2013; Volume 12, pp. 239–265. [Google Scholar]
  4. Gauthier, T.; Duarte-Hospital, C.; Vignard, J.; Boutet-Robinet, E.; Sulyok, M.; Snini, S.P.; Alassane-Kpembi, I.; Lippi, Y.; Puel, S.; Oswald, I.P.; et al. Versicolorin A, a precursor in aflatoxins biosynthesis, is a food contaminant toxic for human intestinal cells. Environ. Int. 2020, 137, 105568. [Google Scholar] [CrossRef] [PubMed]
  5. Medina, A.; Rodriguez, A.; Magan, N. Effect of climate change on Aspergillus flavus and aflatoxin B1 production. Front. Microbiol. 2014, 5, 348. [Google Scholar] [CrossRef] [PubMed]
  6. Van der Fels-Klerx, H.J.; Liu, C.; Battilani, P. Modelling climate change impacts on mycotoxin contamination. World Mycotoxin J. 2016, 9, 717–726. [Google Scholar] [CrossRef] [Green Version]
  7. Bailly, S.; Mahgubi, A.E.; Carvajal, A.; Lorber, S.; Puel, O.; Oswald, I.P.; Bailly, J.D.; Orlando, B. Occurrence and Identification of Aspergillus Section Flavi in the Context of the Emergence of Aflatoxins in French Maize. Toxins 2018, 10, 525. [Google Scholar] [CrossRef] [Green Version]
  8. Battilani, P.; Toscano, P.; Van der Fels-Klerx, H.; Moretti, A.; Leggieri, M.C.; Rortais, B.A.; Goumperis, T.; Robinson, T. Aflatoxin B1 contamination in maize in Europe increases due to climate change. Sci. Rep. 2016, 6, 24328. [Google Scholar] [CrossRef] [Green Version]
  9. Brankov, T.P.; Jovanovic, M.; Grujic, B. Aflatoxin Standards and Maize Trade. Econ. Agric. 2013, 3, 595–607. [Google Scholar]
  10. Sakuda, S.; Yoshinari, T.; Furukawa, T.; Jermnak, U.; Takagi, K.; Iimura, K.; Yamamoto, T.; Suzuki, M.; Nagasawa, H. Search for aflatoxin and trichothecene production inhibitors and analysis of their modes of action. Biosci. Biotechnol. Biochem. 2016, 80, 43–54. [Google Scholar] [CrossRef] [Green Version]
  11. Zhou, W.; Hu, L.B.; Zhao, Y.; Wang, M.Y.; Zhang, H.; Mo, H.Z. Inhibition of fungal aflatoxin B1 biosynthesis by diverse botanically-derived polyphenols. Trop. J. Pharm. Res. 2015, 14, 605–609. [Google Scholar] [CrossRef] [Green Version]
  12. Jahanshiri, Z.; Shams-Ghahfarokhi, M.; Allameh, A.; Razzaghi-Abyaneh, M. Effect of curcumin on Aspergillus parasiticus growth and expression of major genes involved in the early and late stages of aflatoxin biosynthesis. Iran. J. Public Health 2012, 41, 72–79. [Google Scholar] [PubMed]
  13. Moon, Y.-S.; Choi, W.-S.; Park, E.-S.; Bae, I.; Choi, S.-D.; Paek, O.; Kim, S.-H.; Chun, H.; Lee, S.-E. Antifungal and Antiaflatoxigenic Methylenedioxy-Containing Compounds and Piperine-Like Synthetic Compounds. Toxins 2016, 8, 240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. El Khoury, R.; Caceres, I.; Puel, O.; Bailly, S.; Atoui, A.; Oswald, I.P.; El Khoury, A.; Bailly, J.D. Identification of the Anti-Aflatoxinogenic Activity of Micromeria graeca and Elucidation of Its Molecular Mechanism in Aspergillus flavus. Toxins 2017, 9, 87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Caceres, I.; El Khoury, R.; Bailly, S.; Oswald, I.P.; Puel, O.; Bailly, J.D. Piperine inhibits aflatoxin B1 production in Aspergillus flavus by modulating fungal oxidative stress response. Fungal Genet. Biol. 2017, 107, 77–85. [Google Scholar] [CrossRef]
  16. Caceres, I.; Snini, S.; Puel, O.; Mathieu, F. Streptomyces roseolus, A Promising Biocontrol Agent Against Aspergillus flavus, the Main Aflatoxin B1 Producer. Toxins 2018, 10, 442. [Google Scholar] [CrossRef] [Green Version]
  17. Liang, D.; Xing, F.; Selvaraj, J.N.; Liu, X.; Wang, L.; Hua, H.; Zhou, L.; Zhao, Y.; Wang, Y.; Liu, Y. Inhibitory Effect of Cinnamaldehyde, Citral, and Eugenol on Aflatoxin Biosynthetic Gene Expression and Aflatoxin B1 Biosynthesis in Aspergillus flavus. J. Food Sci. 2015, 80, M2917–M2924. [Google Scholar] [CrossRef]
  18. Caceres, I.; El Khoury, R.; Medina, Á.; Lippi, Y.; Naylies, C.; Atoui, A.; El Khoury, A.; Oswald, I.P.; Bailly, J.D.; Puel, O. Deciphering the Anti-Aflatoxinogenic Properties of Eugenol Using a Large-Scale q-PCR Approach. Toxins 2016, 8, 123. [Google Scholar] [CrossRef] [Green Version]
  19. Yu, J.; Bhatnagar, D.; Ehrlich, K.C. Aflatoxin biosynthesis. Rev. Iberoam. Micol. 2002, 19, 191–200. [Google Scholar]
  20. Yabe, K.; Nakajima, H. Enzyme reactions and genes in aflatoxin biosynthesis. Appl. Microbiol. Biotechnol. 2004, 64, 745–755. [Google Scholar] [CrossRef]
  21. Yu, J.; Ehrlich, K.C. Aflatoxin Biosynthetic Pathway and Pathway Genes. In Aflatoxins–Biochemistry and Molecular Biology; Guevara Gonzalez, R., Ed.; InterchOpen: London, UK, 2011; Volume 3, pp. 41–66. [Google Scholar]
  22. Yu, J. Current understanding on aflatoxin biosynthesis and future perspective in reducing aflatoxin contamination. Toxins 2012, 4, 1024–1057. [Google Scholar] [CrossRef] [Green Version]
  23. Roze, L.V.; Hong, S.-Y.; Linz, J.E. Aflatoxin Biosynthesis: Current Frontiers. Annu. Rev. Food Sci. Technol. 2013, 4, 293–311. [Google Scholar] [CrossRef] [PubMed]
  24. Ehrlich, K.C. Predicted roles of the uncharacterized clustered genes in aflatoxin biosynthesis. Toxins 2009, 1, 37–58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Chang, P. The Aspergillus parasiticus protein AFLJ interacts with the aflatoxin pathway-specific regulator AFLR. Mol. Genet. Genom. 2003, 268, 711–719. [Google Scholar] [CrossRef]
  26. Price, M.S.; Yu, J.; Nierman, W.C.; Kim, H.; Pritchard, B.; Jacobus, C.A.; Bhatnagar, D.; Cleveland, T.E.; Payne, G.A. The aflatoxin pathway regulator AflR induces gene transcription inside and outside of the aflatoxin biosynthetic cluster. FEMS Microbiol. Lett. 2006, 255, 275–279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Klich, M.A. Environmental and developmental factors influencing aflatoxin production by Aspergillus flavus and Aspergillus parasiticus. Mycoscience 2007, 48, 71–80. [Google Scholar] [CrossRef]
  28. Montibus, M.; Pinson-Gadais, L.; Richard-Forget, F.; Barreau, C.; Ponts, N. Coupling of transcriptional response to oxidative stress and secondary metabolism regulation in filamentous fungi. Crit. Rev. Microbiol. 2013, 41, 295–308. [Google Scholar] [CrossRef] [PubMed]
  29. Georgianna, D.R.; Payne, G.A. Genetic regulation of aflatoxin biosynthesis: From gene to genome. Fungal Genet. Biol. 2009, 46, 113–125. [Google Scholar] [CrossRef]
  30. Affeldt, K.J.; Carrig, J.; Amare, M.; Keller, N.P. Global Survey of Canonical Aspergillus flavus G Protein-Coupled Receptors. Am. Soc. Microbiol. 2014, 5, e01501–e01514. [Google Scholar] [CrossRef] [Green Version]
  31. Yu, J.-H.; Keller, N. Regulation of secondary metabolism in filamentous fungi. Annu. Rev. Phytopathol. 2005, 43, 437–458. [Google Scholar] [CrossRef]
  32. Woloshuk, C.P.; Prieto, R. Genetic organization and function of the aflatoxin B1 biosynthetic genes. FEMS Microbiol. Lett. 1998, 160, 169–176. [Google Scholar] [CrossRef]
  33. Yu, J.; Chang, P.; Ehrlich, K.C.; Cary, J.W.; Bhatnagar, D.; Cleveland, T.E.; Payne, G.A.; Linz, J.E.; Woloshuk, C.P.; Bennett, J.W. Clustered Pathway Genes in Aflatoxin Biosynthesis. Appl. Environ. Microbiol. 2004, 70, 1253–1262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Trail, F.; Mahanti, N.; Linz, J. Molecular biology of aflatoxin biosynthesis. Microbiology 1995, 141, 755–765. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Do, J.H.; Choi, D.-K. Aflatoxins: Detection, Toxicity, and Biosynthesis. Biotechnol. Bioprocess Eng. 2007, 12, 585–593. [Google Scholar] [CrossRef]
  36. Payne, G.A.; Brown, M.P. Genetics and physiology of aflatoxin biosynthesis. Annu. Rev. Phytopathol. 1998, 36, 329–362. [Google Scholar] [CrossRef]
  37. Amaike, S.; Keller, N.P. Aspergillus flavus. Annu. Rev. Phytopathol. 2011, 49, 107–133. [Google Scholar] [CrossRef]
  38. Yu, J.; Woloshuk, C.P.; Bhatnagar, D.; Cleveland, T.E. Cloning and characterization of avfA and omtB genes involved in aflatoxin biosynthesis in three Aspergillus species. Gene 2000, 248, 157–167. [Google Scholar] [CrossRef]
  39. Ehrlich, K.C.; Chang, P.-K.; Yu, J.; Cotty, P.J. Aflatoxin Biosynthesis Cluster Gene cypA Is Required for G Aflatoxin Formation. Appl. Environ. Microbiol. 2004, 70, 6518–6524. [Google Scholar] [CrossRef] [Green Version]
  40. Ehrlich, K.C.; Scharfenstein, L.L.; Montalbano, B.G.; Chang, P.K. Are the genes nadA and norB involved in formation of aflatoxin G1? Int. J. Mol. Sci. 2008, 9, 1717–1729. [Google Scholar] [CrossRef]
  41. Birch, A.L. Biosynthesis of polyketides and related compounds. Science 1967, 156, 202–206. [Google Scholar] [CrossRef]
  42. Minto, R.E.; Townsend, C.A. Enzymology and Molecular Biology of Aflatoxin Biosynthesis. Chem. Rev. 1997, 97, 2537–2556. [Google Scholar] [CrossRef]
  43. Dutton, M.F. Enzymes and aflatoxin biosynthesis. Microbiol. Rev. 1988, 52, 274–295. [Google Scholar] [CrossRef] [PubMed]
  44. Nelson, D.R. Progress in tracing the evolutionary paths of cytochrome P450. Biochim. Biophys. Acta 2011, 1814, 14–18. [Google Scholar] [CrossRef] [PubMed]
  45. Roze, L.V.; Laivenieks, M.; Hong, S.Y.; Wee, J.; Wong, S.S.; Vanos, B.; Awad, D.; Ehrlich, K.C.; Linz, J.E. Aflatoxin biosynthesis is a novel source of reactive oxygen species-A potential redox signal to initiate resistance to oxidative stress? Toxins 2015, 7, 1411–1430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Yabe, K.; Ando, Y.; Hamasaki, T. Desaturase Activity in the Branching Step Between Aflatoxins B1 And G1 and Aflatoxins B2 and G2. Agric. Biol. Chem. 1991, 55, 1907–1911. [Google Scholar] [CrossRef]
  47. Sakuno, E.; Wen, Y.; Hatabayashi, H.; Arai, H.; Aoki, C.; Yabe, K.; Nakajima, H. Aspergillus parasiticus Cyclase Catalyzes Two Dehydration Steps in Aflatoxin Biosynthesis. Appl. Environ. Microbiol. 2005, 71, 2999–3006. [Google Scholar] [CrossRef] [Green Version]
  48. Wen, Y.; Hatabayashi, H.; Arai, H.; Kitamoto, H.K.; Yabe, K. Function of the cypX and moxY Genes in Aflatoxin Biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 2005, 71, 3192–3198. [Google Scholar] [CrossRef] [Green Version]
  49. Yabe, K.; Chihaya, N.; Hamamatsu, S.; Sakuno, E.; Hamasaki, T.; Nakajima, H.; Bennett, J.W. Enzymatic conversion of averufin to hydroxyversicolorone and elucidation of a novel metabolic grid involved in aflatoxin biosynthesis. Appl. Environ. Microbiol. 2003, 69, 66–73. [Google Scholar] [CrossRef] [Green Version]
  50. Sakuno, E.; Yabe, K.; Nakajima, H. Involvement of Two Cytosolic Enzymes and a Novel Intermediate, 5′-Oxoaverantin, in the Pathway from 5′-Hydroxyaverantin to Averufin in Aflatoxin Biosynthesis. Appl. Environ. Microbiol. 2003, 69, 6418–6426. [Google Scholar] [CrossRef] [Green Version]
  51. Chang, P.K.; Yabe, K.; Yu, J. The Aspergillus parasiticus estA-encoded esterase converts versiconal hemiacetal acetate to versiconal and versiconol acetate to versiconol in aflatoxin biosynthesis. Appl. Environ. Microbiol. 2004, 70, 3593–3599. [Google Scholar] [CrossRef] [Green Version]
  52. Henry, K.M.; Townsend, C.A. Ordering the reductive and cytochrome P450 oxidative steps in demethylsterigmatocystin formation yields general insights into the biosynthesis of aflatoxin and related fungal metabolites. J. Am. Chem. Soc. 2005, 127, 3724–3733. [Google Scholar] [CrossRef]
  53. Cary, J.W.; Ehrlich, K.C.; Bland, J.M.; Montalbano, B.G. The aflatoxin biosynthesis cluster gene, aflX, encodes an oxidoreductase involved in conversion of versicolorin a to demethylsterigmatocystin. Appl. Environ. Microbiol. 2006, 72, 1096–1101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Ehrlich, K.C.; Montalbano, B.; Boué, S.M.; Bhatnagar, D. An aflatoxin biosynthesis cluster gene encodes a novel oxidase required for conversion of versicolorin A to sterigmatocystin. Appl. Environ. Microbiol. 2005, 71, 8963–8965. [Google Scholar] [CrossRef] [Green Version]
  55. Zeng, H.; Hatabayashi, H.; Nakagawa, H.; Cai, J.; Suzuki, R.; Sakuno, E.; Tanaka, T.; Ito, Y.; Ehrlich, K.C.; Nakajima, H.; et al. Conversion of 11-hydroxy-O methylsterigmatocystin to Aflatoxin G1 in Aspergillus parasiticus. Appl. Microbiol. Biotechnol. 2011, 90, 635–650. [Google Scholar] [CrossRef] [PubMed]
  56. Yu, J.; Bhatnagar, D.; Cleveland, T.E. Completed sequence of aflatoxin pathway gene cluster in Aspergillus parasiticus. FEBS Lett. 2004, 564, 126–130. [Google Scholar] [CrossRef] [Green Version]
  57. Ehrlich, K.C.; Li, P.; Scharfenstein, L.; Chang, P.K. HypC, the Anthrone Oxidase Involved in Aflatoxin Biosynthesis. Appl. Environ. Microbiol. 2010, 76, 3374–3377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Zhou, R.; Linz, J.E. Enzymatic function of the Nor-1 protein in aflatoxin biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 1999, 65, 5639–5641. [Google Scholar] [CrossRef] [Green Version]
  59. Yabe, K.; Nakamura, Y.; Nakajima, H.; Ando, Y. Enzymatic Conversion of Norsolorinic Acid to Averufin in Aflatoxin Biosynthesis. Appl. Environ. Microbiol. 1991, 57, 1340–1345. [Google Scholar] [CrossRef] [Green Version]
  60. Yu, J.; Chang, P.K.; Bhatnagar, D.; Cleveland, T.E. avnA, Encoding A Putative P-450 Monooxygenase, is Involved in the Conversion of Averantin to Averufin in Aflatoxin Biosynthesis in Aspergillus parasiticus. Appl. Environ. Microbiol. 1997, 63, 1349–1356. [Google Scholar] [CrossRef] [Green Version]
  61. Cary, J.W.; Montalbano, B.G.; Ehrlich, K.C. Promoter Elements Involved in The Expression of the Aspergillus parasiticus Aflatoxin Pathway Gene avnA. Biochim. Biophys. Acta 2000, 1491, 7–12. [Google Scholar] [CrossRef]
  62. Chang, P.K.; Yu, J.; Ehrlich, K.C.; Boue, S.M.; Montalbano, B.G.; Bhatnagar, D.; Cleveland, T.E. adhA in Aspergillus parasiticus is involved in conversion of 5′-hydroxyaverantin to averufin. Appl. Environ. Microbiol. 2000, 66, 4715–4719. [Google Scholar] [CrossRef] [Green Version]
  63. Yu, J.; Chang, P.K.; Bhatnagar, D.; Cleveland, T.E. Cloning and functional expression of an esterase gene in Aspergillus parasiticus. Mycopathologia 2002, 156, 227–234. [Google Scholar] [CrossRef] [PubMed]
  64. Lin, B.K.; Anderson, J.A. Purification and properties of versiconal cyclase from Aspergillus parasiticus. Arch. Biochem. Biophys. 1992, 193, 67–70. [Google Scholar] [CrossRef]
  65. Kelkar, H.S.; Skloss, T.W.; Haw, J.F.; Keller, N.P.; Adams, T.H. Aspergillus nidulans stcL encodes a putative cytochrome P-450 monooxygenase required for bisfuran desaturation during aflatoxin/sterigmatocystin biosynthesis. J. Biol. Chem. 1997, 272, 1589–1594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Keller, N.P.; Kantz, N.J.; Adams, T.H. Aspergillus nidulans verA is required for production of the mycotoxin sterigmatocystin. Appl. Environ. Microbiol. 1994, 60, 1444–1450. [Google Scholar] [CrossRef] [Green Version]
  67. Hong, S.Y.; Roze, L.V.; Wee, J.; Linz, J.E. Evidence that a transcription factor regulatory network coordinates oxidative stress response and secondary metabolism in Aspergilli. Microbiologyopen 2013, 2, 144–160. [Google Scholar] [CrossRef]
  68. Keller, N.P.; Segner, S.; Bhatnagar, D.; Adams, T.H. stcS, a putative P-450 monooxygenase, is required for the conversion of versicolorin A to sterigmatocystin in Aspergillus nidulans. Appl. Environ. Microbiol. 1995, 61, 3628–3632. [Google Scholar] [CrossRef] [Green Version]
  69. Yabe, K.; Ando, Y.; Hashimoto, J.; Hamasaki, T. Two distinct O-methyltransferases in Aflatoxin Biosynthesis. Appl. Environ. Microbiol. 1989, 55, 2172–2177. [Google Scholar] [CrossRef] [Green Version]
  70. Scherm, B.; Palomba, M.; Serra, D.; Marcello, A.; Migheli, Q. Detection of transcripts of the aflatoxin genes aflD, aflO, and aflP by reverse transcription-polymerase chain reaction allows differentiation of aflatoxin-producing and non-producing isolates of Aspergillus flavus and Aspergillus parasiticus. Int. J. Food Microbiol. 2005, 98, 201–210. [Google Scholar] [CrossRef]
  71. Yu, J.; Chang, P.K.; Ehrlich, K.C.; Cary, J.W.; Montalbano, B.; Dyer, J.M.; Bhatnagar, D.; Cleveland, T.E. Characterization of the Critical Amino Acids of an Aspergillus parasiticus Cytochrome P-450 Monooxygenase Encoded by ordA that is Involved in the Biosynthesis of Aflatoxins B1, G1, B2, and G2. Appl. Environ. Microbiol. 1998, 64, 4834–4841. [Google Scholar] [CrossRef] [Green Version]
  72. Cary, J.W.; Wright, M.; Bhatnagar, D.; Lee, R.; Chu, F.S. Molecular Characterization of an Aspergillus parasiticus dehydrogenase gene, norA, located on the Aflatoxin Biosynthesis Gene. Appl. Environ. Microbiol. 1996, 62, 360–366. [Google Scholar] [CrossRef] [Green Version]
  73. Ehrlich, K.C.; Chang, P.-K.; Scharfenstein, L.L.; Cary, J.W.; Crawford, J.M.; Townsend, C.A. Absence of the aflatoxin biosynthesis gene, norA, allows accumulation of deoxyaflatoxin B1 in Aspergillus flavus cultures. FEMS Microbiol. Lett. 2010, 305, 65–70. [Google Scholar] [CrossRef] [PubMed]
  74. Chang, P.K.; Yu, J.; Yu, J.H. aflT, a MFS transporter-encoding gene located in the aflatoxin gene cluster, does not have a significant role in aflatoxin secretion. Fungal Genet. Biol. 2004, 41, 911–920. [Google Scholar] [CrossRef] [PubMed]
  75. Chanda, A.; Roze, L.V.; Linz, J.E. A Possible Role for Exocytosis in Aflatoxin Export in Aspergillus parasiticus. Eukaryot. Cell 2010, 9, 1727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Yu, J.-H.; Butchko, R.E.; Fernandes, M.; Keller, N.; Leonard, T.; Adams, T. Conservation of structure and function of the aflatoxin regulatory gene aflR from Aspergillus nidulans and A. flavus. Curr. Genet. 1996, 29, 549–555. [Google Scholar] [CrossRef] [PubMed]
  77. Shimizu, K.; Hicks, J.K.; Huang, T.P.; Keller, N.P. Pka, Ras and RGS Protein Interactions Regulate Activity of AflR, a Zn(II)2Cys6 Transcription Factor in Aspergillus nidulans. Genetics 2003, 165, 1095–1104. [Google Scholar] [PubMed]
  78. MacPherson, S.; Larochelle, M.; Turcotte, B. A fungal family of transcriptional regulators: The zinc cluster proteins. Microbiol. Mol. Biol. Rev. 2006, 70, 583–604. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Woloshuk, C.P.; Foutz, K.R.; Brewer, J.F.; Bhatnagar, D.; Cleveland, T.E.; Payne, G.A. Molecular characterization of aflR, a regulatory locus for aflatoxin biosynthesis. Appl. Environ. Microbiol. 1994, 60, 2408–2414. [Google Scholar] [CrossRef] [Green Version]
  80. Ehrlich, K.C.; Montalbano, B.G.; Cary, J.W. Binding of the C6-zinc cluster protein, AFLR, to the promoters of aflatoxin pathway biosynthesis genes in Aspergillus parasiticus. Gene 1999, 230, 249–257. [Google Scholar] [CrossRef]
  81. Chang, P.-K.; Ehrlich, K.C.; Yu, J.; Bhatnagar, D.; Cleveland, T.E. Increased expression of Aspergillus parasiticus aflR, encoding a sequence- specific DNA-binding protein, relieves nitrate inhibition of aflatoxin biosynthesis. Appl. Environ. Microbiol. 1995, 61, 2372–2377. [Google Scholar] [CrossRef] [Green Version]
  82. Flaherty, J.E.; Payne, G.A. Overexpression of aflR Leads to Upregulation of Pathway Gene Transcription and Increased Aflatoxin Production in Aspergillus flavus. Appl. Environ. Microbiol. 1997, 63, 3995–4000. [Google Scholar] [CrossRef] [Green Version]
  83. Cary, J.W.; Ehrlich, K.C.; Wright, M.; Chang, P.K.; Bhatnagar, D. Generation of aflR disruption mutants of Aspergillus parasiticus. Appl. Microbiol. Biotechnol. 2000, 53, 680–684. [Google Scholar] [CrossRef] [PubMed]
  84. Dhanamjayulu, P.; Boga, R.B.; Mehta, A. Inhibition of aflatoxin B1 biosynthesis and down regulation of aflR and aflB genes in presence of benzimidazole derivatives without impairing the growth of Aspergillus flavus. Toxicon 2019, 170, 60–67. [Google Scholar] [CrossRef] [PubMed]
  85. Jahanshiri, Z.; Shams-Ghahfarokhi, M.; Allameh, A.; Razzaghi-Abyaneh, M. Inhibitory effect of eugenol on aflatoxin B1 production in Aspergillus parasiticus by downregulating the expression of major genes in the toxin biosynthetic pathway. World J. Microbiol. Biotechnol. 2015, 31, 1071–1078. [Google Scholar] [CrossRef]
  86. Jermnak, U.; Yoshinari, T.; Sugiyama, Y.; Tsuyuki, R.; Nagasawa, H.; Sakuda, S. Isolation of methyl syringate as a specific aflatoxin production inhibitor from the essential oil of Betula alba and aflatoxin production inhibitory activities of its related compounds. Int. J. Food Microbiol. 2012, 153, 339–344. [Google Scholar] [CrossRef] [PubMed]
  87. Meyers, D.M.; Obrian, G.; Du, W.L.; Bhatnagar, D.; Payne, G.A. Characterization of aflJ, a gene required for conversion of pathway intermediates to aflatoxin. Appl. Environ. Microbiol. 1998, 64, 3713–3717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Du, W.; Obrian, G.R.; Payne, G.A. Function and regulation of aflJ in the accumulation of aflatoxin early pathway intermediate in Aspergillus flavus. Food Addit. Contam. 2007, 24, 1043–1050. [Google Scholar] [CrossRef]
  89. Kong, Q.; Chi, C.; Yu, J.; Shan, S.; Li, Q.; Li, Q.; Guan, B.; Nierman, W.C.; Bennett, J.W. The inhibitory effect of Bacillus megaterium on aflatoxin and cyclopiazonic acid biosynthetic pathway gene expression in Aspergillus flavus. Appl. Environ. Microbiol. 2014, 98, 5161–5172. [Google Scholar] [CrossRef]
  90. Yoshinari, T.; Akiyama, T.; Nakamura, K.; Kondo, T.; Takahashi, Y.; Muraoka, Y.; Nonomura, Y.; Nagasawa, H.; Sakuda, S. Dioctatin A is a strong inhibitor of aflatoxin production by Aspergillus parasiticus. Microbiology 2007, 153, 2774–2780. [Google Scholar] [CrossRef] [Green Version]
  91. Gallo, A.; Solfrizzo, M.; Epifani, F.; Panzarini, G.; Perrone, G. Effect of temperature and water activity on gene expression and aflatoxin biosynthesis in Aspergillus flavus on almond medium. Int. J. Food Microbiol. 2016, 217, 162–169. [Google Scholar] [CrossRef]
  92. Gizachew, D.; Chang, C.H.; Szonyi, B.; De La Torre, S.; Ting, W.E. Aflatoxin B1 (AFB1) production by Aspergillus flavus and Aspergillus parasiticus on ground Nyjer seeds: The effect of water activity and temperature. Int. J. Food Microbiol. 2019, 296, 8–13. [Google Scholar] [CrossRef]
  93. Deepika, V.B.; Murali, T.S.; Satyamoorthy, K. Modulation of genetic clusters for synthesis of bioactive molecules in fungal endophytes: A review. Microbiol. Res. 2015, 182, 125–140. [Google Scholar] [CrossRef] [PubMed]
  94. Davis, N.D.; Diener, U.L. Growth and Aflatoxin Production by Aspergillus parasiticus from Various Carbon Sources. Appl. Microbiol. 1968, 16, 158–159. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Maggio-Hall, L.A.; Wilson, R.A.; Keller, N.P. Fundamental contribution of beta-oxidation to polyketide mycotoxin production in planta. Mol. Plant Microbe Interact. 2005, 18, 783–793. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Bhatnagar, D.; Cary, J.W.; Ehrlich, K.; Yu, J.; Cleveland, T.E. Understanding the genetics of regulation of aflatoxin production and Aspergillus flavus development. Mycopathologia 2006, 162, 155–166. [Google Scholar] [CrossRef] [PubMed]
  97. Calvo, A.M.; Wilson, R.A.; Bok, J.W.; Keller, N.P. Relationship between Secondary Metabolism and Fungal Development. Microbiol. Mol. Biol. Rev. 2002, 66, 447–459. [Google Scholar] [CrossRef] [Green Version]
  98. Liu, J.; Sun, L.; Zhang, N.; Zhang, J.; Guo, J.; Li, C.; Rajput, S.A.; Qi, D. Effects of Nutrients in Substrates of Different Grains on Aflatoxin B1 Production by Aspergillus flavus. BioMed Res. Int. 2016, 2016, 1–10. [Google Scholar]
  99. Zhang, J.-D.; Han, L.; Yan, S.; Liu, C.-M. The non-metabolizable glucose analog D-glucal inhibits aflatoxin biosynthesis and promotes kojic acid production in Aspergillus flavus. BMC Microbiol. 2014, 14, 1–9. [Google Scholar] [CrossRef] [Green Version]
  100. Yu, J.; Chang, P.K.; Bhatnagar, D.; Cleveland, T.E. Cloning of a Sugar Utilization Gene Cluster in Aspergillus parasiticus. Biochim. Biophys. Acta 2000, 1493, 211–214. [Google Scholar] [CrossRef]
  101. Ruijter, G.J.G.; Visser, J. Carbon repression in Aspergilli. FEMS Microbiol. Lett. 1997, 151, 103–114. [Google Scholar] [CrossRef]
  102. Fasoyin, O.E.; Wang, B.; Qiu, M.; Han, X.; Chung, K.R.; Wang, S. Carbon catabolite repression gene creA regulates morphology, aflatoxin biosynthesis and virulence in Aspergillus flavus. Fungal Genet. Biol. 2018, 115, 41–51. [Google Scholar] [CrossRef]
  103. Shroff, R.A.; Lockington, R.A.; Kelly, J.M. Analysis of mutations in the creA gene involved in carbon catabolite repression in Aspergillus nidulans. Can. J. Microbiol. 1996, 42, 950–959. [Google Scholar] [CrossRef] [PubMed]
  104. Ries, L.N.A.; Beattie, S.R.; Espeso, E.A.; Cramer, R.A.; Goldma, G.H. Diverse Regulation of the CreA Carbon Catabolite Repressor in Aspergillus nidulans. Genetics 2016, 203, 335–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Dowzer, C.E.A.; Kelly, J.M. Cloning of the Crea gene from Aspergillus nidulans–A gene involved in carbon catabolite repression. Curr. Genet. 1989, 15, 457–459. [Google Scholar] [CrossRef] [PubMed]
  106. Zhao, X.; Zhi, Q.-Q.; Li, J.-Y.; Keller, N.; He, Z.-M. The Antioxidant Gallic Acid Inhibits Aflatoxin Formation in Aspergillus flavus by Modulating Transcription Factors FarB and CreA. Toxins 2018, 10, 270. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Wilkinson, J.R.; Yu, J.; Abbas, H.K.; Scheffler, B.E.; Kim, H.S.; Nierman, W.C.; Bhatnagar, D.; Cleveland, T.E. Aflatoxin formation and gene expression in response to carbon source media shift in Aspergillus parasiticus. Food Addit. Contam. 2007, 24, 1051–1060. [Google Scholar] [CrossRef] [PubMed]
  108. Kudla, B.; Caddick, M.X.; Langdon, T.; Martinez-Rossi, N.M.; Bennett, C.F.; Sibley, S.; Davies, R.W.; Arst, H.N. The regulatory gene areA mediating nitrogen metabolite repression in Aspergillus nidulans. Mutations affecting specificity of gene activation alter a loop residue of a putative zinc finger. EMBO J. 1990, 9, 1355–1364. [Google Scholar] [CrossRef]
  109. Amaike, S.; Affeldt, K.J.; Yin, W.B.; Franke, S.; Choithani, A.; Keller, N.P. The bZIP Protein MeaB Mediates Virulence Attributes in Aspergillus flavus. PLoS ONE 2013, 8, e74030. [Google Scholar] [CrossRef] [Green Version]
  110. Chang, P.K.; Yu, J.H.; Bhatnagar, D.; Cleveland, T.E. Characterization of the Aspergillus parasiticus Major Nitrogen Regulatory Gene, areA. Biochim. Biophys. Acta 2000, 1491, 263–266. [Google Scholar] [CrossRef]
  111. Fasoyin, O.E.; Yang, K.; Qiu, M.; Wang, B.; Wang, S.; Wang, S. Regulation of Morphology, Aflatoxin Production, and Virulence of Aspergillus flavus by the Major Nitrogen Regulatory Gene areA. Toxins 2019, 11, 718. [Google Scholar] [CrossRef] [Green Version]
  112. Ehrlich, K.C.; Cotty, P.J. Variability in Nitrogen Regulation of Aflatoxin Production by Aspergillus flavus Strains. Microbiol. Biotechnol. 2002, 60, 174–178. [Google Scholar]
  113. Ehrlich, K.C.; Cary, J.W.; Montalbano, B.G. Characterization of the promoter for the gene encoding the aflatoxin biosynthetic pathway regulatory protein AFLR. Biochim. Biophys. Acta Gene Struct. Expr. 1999, 1444, 412–417. [Google Scholar] [CrossRef]
  114. Feng, G.H.; Leonard, T.J. Culture Conditions Control Expression Of The Genes For Aflatoxin and Sterigmatocystin Biosynthesis in Aspergillus parasiticus and A. nidulans. Appl. Environ. Microbiol. 1998, 64, 2275–2277. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Luo, Z.; Ren, H.; Mousa, J.J.; Rangel, D.E.; Zhang, Y.; Bruner, S.D.; Keyhani1ing, N.O. The PacC transcription factor regulates secondary metabolite production, stress response, but has only minor effects on virulence in the insect pathogenic fungus Beauveria bassiana. Environ. Microbiol. 2017, 19, 788–802. [Google Scholar] [CrossRef]
  116. Selvig, K.; Alspaugh, J.A. pH Response Pathways in Fungi: Adapting to Host-derived and Environmental Signals. Mycobiology 2011, 39, 249–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Tilburn, J.; Sarkar, S.; Widdick, D.A.; Espeso, E.A.; Orejas, M.; Mungroo, J.; Peñalva, M.A.; Arst, H.N. The Aspergillus PacC zinc finger transcription factor mediates regulation of both acid- and alkaline-expressed genes by ambient pH. EMBO J. 1995, 14, 779–790. [Google Scholar] [CrossRef]
  118. Cornet, M.; Gaillardin, C. pH signaling in human fungal pathogens: A new target for antifungal strategies. Eukaryot. Cell 2014, 13, 342–352. [Google Scholar] [CrossRef] [Green Version]
  119. Cotty, P.J. Aflatoxin and Sclerotial Production by Aspergillus flavus: Influence of pH. Physiol. Biochem. 1988, 78, 1250–1253. [Google Scholar]
  120. Buchanan, R.L.; Ayres, J.C. Effects of initial pH on aflatoxin production. Appl. Microbiol. 1975, 30, 1050–1051. [Google Scholar] [CrossRef] [Green Version]
  121. Keller, N.P.; Nesbitt, C.; Sarr, B.; Phillips, T.D.; Burow, G.B. pH Regulation of Sterigmatocystin and Aflatoxin Biosynthesis in Aspergillus spp. Phytopathology 1997, 87, 643–648. [Google Scholar] [CrossRef] [Green Version]
  122. Rangel, D.E.N.; Alder-Rangel, A.; Dadachova, E.; Finlay, R.D.; Kupiec, M.; Dijksterhuis, J.; Braga, G.U.L.; Corrochano, L.M.; Hallsworth, J.E. Fungal stress biology: A preface to the Fungal Stress Responses special edition. Curr. Genet. 2015, 61, 231–238. [Google Scholar] [CrossRef]
  123. Calvo, A.M. The VeA regulatory system and its role in morphological and chemical development in fungi. Fungal Genet. Biol. 2008, 45, 1053–1061. [Google Scholar] [CrossRef] [PubMed]
  124. Purschwitz, J.; Müller, S.; Kastner, C.; Schöser, M.; Haas, H.; Espeso, E.A.; Atoui, A.; Calvo, A.M.; Fischer, R. Functional and Physical Interaction of Blue- and Red-Light Sensors in Aspergillus nidulans. Curr. Biol. 2008, 18, 255–259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Bayram, O.; Krappmann, S.; Ni, M.; Bok, J.W.; Helmstaedt, K.; Valerius, O.; Braus-Stromeyer, S.; Kwon, N.-J.; Keller, N.P.; Yu, J.-H.; et al. VelB/VeA/LaeA complex coordinates light signal with fungal development and secondary metabolism. Science 2008, 320, 1504–1506. [Google Scholar] [CrossRef] [PubMed]
  126. Sarikaya-Bayram, O.; Palmer, J.M.; Keller, N.; Braus, G.H.; Bayram, O. One Juliet and four Romeos: VeA and its methyltransferases. Front. Microbiol. 2015, 6, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Bayram, Ö.; Braus, G.H. Coordination of secondary metabolism and development in fungi: The velvet family of regulatory proteins. FEMS Microbiol. Rev. 2012, 36, 1–24. [Google Scholar] [CrossRef] [Green Version]
  128. Ahmed, Y.L.; Gerke, J.; Park, H.S.; Bayram, O.; Neumann, P.; Ni, M.; Dickmanns, A.; Kim, S.C.; Yu, J.H.; Braus, G.H.; et al. The velvet family of fungal regulators contains a DNA-binding domain structurally similar to NF-kB. PLoS Biol. 2013, 11, e1001750. [Google Scholar] [CrossRef] [Green Version]
  129. Palmer, J.M.; Theisen, J.M.; Duran, R.M.; Grayburn, W.S.; Calvo, A.M.; Keller, N.P. Secondary Metabolism and Development Is Mediated by LlmF Control of VeA Subcellular Localization in Aspergillus nidulans. PLoS Genet. 2013, 9, e1003193. [Google Scholar] [CrossRef] [Green Version]
  130. Purschwitz, J.; Muller, S.; Fischer, R. Mapping the interaction sites of Aspergillus nidulans phytochrome FphA with the global regulator VeA and the White Collar protein LreB. Mol. Genet. Genom. 2009, 281, 35–42. [Google Scholar] [CrossRef]
  131. Sarikaya-Bayram, O.; Bayram, O.; Feussner, K.; Kim, J.H.; Kim, H.S.; Kaever, A.; Feussner, I.; Chae, K.S.; Han, D.M.; Han, K.H.; et al. Membrane-bound methyltransferase complex VapA-VipC-VapB guides epigenetic control of fungal development. Dev. Cell 2014, 29, 406–420. [Google Scholar] [CrossRef] [Green Version]
  132. Duran, R.M.; Cary, J.W.; Calvo, A.M. Production of cyclopiazonic acid, aflatrem, and aflatoxin by Aspergillus flavus is regulated by veA, a gene necessary for sclerotial formation. Appl. Microbiol. Biotechnol. 2007, 73, 1158–1168. [Google Scholar] [CrossRef]
  133. Duran, R.M.; Cary, J.W.; Calvo, A.M. The Role of veA in Aspergillus flavus Infection of Peanut, Corn and Cotton. Open Mycol. J. 2009, 3, 27–36. [Google Scholar] [CrossRef] [Green Version]
  134. Amaike, S.; Keller, N.P. Distinct roles for VeA and LaeA in development and pathogenesis of Aspergillus flavus. Eukaryot. Cell 2009, 8, 1051–1060. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Chang, P.K.; Scharfenstein, L.L.; Ehrlich, K.C.; Wei, Q.; Bhatnagar, D.; Ingber, B.F. Effects of laeA deletion on Aspergillus flavus conidial development and hydrophobicity may contribute to loss of aflatoxin production. Fungal Biol. 2012, 116, 298–307. [Google Scholar] [CrossRef]
  136. Cary, J.W.; Han, Z.; Yin, Y.; Lohmar, J.M.; Shantappa, S.; Harris-Coward, P.Y.; Mack, B.; Ehrlich, K.C.; Wei, Q.; Arroyo-Manzanares, N.; et al. Transcriptome Analysis of Aspergillus flavus Reveals veA -Dependent Regulation of Secondary Metabolite Gene Clusters, Including the Novel Aflavarin Cluster. Eukaryot. Cell 2015, 14, 983–997. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Calvo, A.M.; Bok, J.W.; Brooks, W.; Keller, N.P. Vea Is Required for Toxin And Sclerotial Production In Aspergillus parasiticus. Appl. Environ. Microbiol. 2004, 70, 4733–4739. [Google Scholar] [CrossRef] [Green Version]
  138. Fischer, R. Developmental biology. Sex and poison in the dark. Science 2008, 320, 1430–1431. [Google Scholar] [CrossRef]
  139. Kato, N.; Brooks, W.; Calvo, A.M. The Expression of Sterigmatocystin and Penicillin Genes in Aspergillus nidulans is controlled by veA, a gene required for Sexual Development. Eukaryot. Cell 2003, 2, 1178–1186. [Google Scholar] [CrossRef] [Green Version]
  140. Bok, J.W.; Keller, N.P. LaeA, a Regulator of Secondary Metabolism in Aspergillus spp. Eukaryot. Cell 2004, 3, 527–535. [Google Scholar] [CrossRef] [Green Version]
  141. Atoui, A.; Kastner, C.; Larey, C.M.; Thokala, R.; Etxebeste, O.; Espeso, E.A.; Fischer, R.; Calvo, A.M. Cross-talk between light and glucose regulation controls toxin production and morphogenesis in Aspergillus nidulans. Fungal Genet. Biol. 2010, 47, 962–972. [Google Scholar] [CrossRef] [Green Version]
  142. Perrone, G.; Gallo, A.; Logrieco, A.F. Biodiversity of Aspergillus section Flavi in Europe in relation to the management of aflatoxin risk. Front. Microbiol. 2014, 5, 377. [Google Scholar] [CrossRef] [Green Version]
  143. Horn, B.W.; Moore, G.G.; Carbone, I. Sexual reproduction in Aspergillus flavus. Mycologia 2009, 101, 423–429. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Horn, B.W.; Sorensen, R.B.; Lamb, M.C.; Sobolev, V.S.; Olarte, R.A.; Worthington, C.J.; Carbone, I. Sexual Reproduction in Aspergillus flavus Sclerotia Naturally produced in Corn. Genet. Resist. 2014, 104, 75–85. [Google Scholar]
  145. Ramirez-Prado, J.H.; Moore, G.G.; Horn, B.W.; Carbone, I. Characterization and Population Analysis of the Mating-Type Genes in Aspergillus flavus and Aspergillus parasiticus. Fungal Genet. Biol. 2008, 45, 1292–1299. [Google Scholar] [CrossRef] [PubMed]
  146. Olarte, R.A.; Horn, B.W.; Dorner, J.W.; Monacell, J.T.; Singh, R.; Stone, E.A.; Carbone, I. Effect of sexual recombination on population diversity in aflatoxin production by Aspergillus flavus and evidence for cryptic heterokaryosis. Mol. Ecol. 2012, 21, 1453–1476. [Google Scholar] [CrossRef]
  147. Seo, J.A.; Guan, Y.; Yu, J.H. Suppressor mutations bypass the requirement of fluG for asexual sporulation and sterigmatocystin production in Aspergillus nidulans. Genetics 2003, 165, 1083–1093. [Google Scholar]
  148. Lee, B.N.; Adams, T.H. FluG and flbA function interdependently to initiate conidiophore development in Aspergillus nidulans through brlA beta activation. EMBO J. 1996, 15, 299–309. [Google Scholar] [CrossRef]
  149. Wieser, J.; Lee, B.N.; Fondon, J.W.; Adams, T.H. Genetic requirements for initiating asexual development in Aspergillus nidulans. Curr. Genet. 1994, 27, 62–69. [Google Scholar] [CrossRef]
  150. Andrianopoulos, A.; Timberlake, W.E. The Aspergillus nidulans abaA gene encodes a transcriptional activator that acts as a genetic switch to control development. Mol. Cell. Biol. 1994, 14, 2503–2515. [Google Scholar] [CrossRef] [Green Version]
  151. Han, S.; Adams, T.H. Complex control of the developmental regulatory locus brlA in Aspergillus nidulans. Mol. Genet. Genom. 2001, 266, 260–270. [Google Scholar] [CrossRef]
  152. Ni, M.; Yu, J.H. A novel regulator couples sporogenesis and trehalose biogenesis in Aspergillus nidulans. PLoS ONE 2007, 2, e970. [Google Scholar] [CrossRef] [Green Version]
  153. Cary, J.W.; Harris-Coward, P.Y.; Ehrlich, K.C.; Mack, B.M.; Kale, S.P.; Larey, C.; Calvo, A.M. NsdC and NsdD affect Aspergillus flavus morphogenesis and aflatoxin production. Eukaryot. Cell 2012, 11, 1104–1111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Gilbert, M.K.; Mack, B.M.; Wei, Q.; Bland, J.M.; Bhatnagar, D.; Cary, J.W. RNA sequencing of an nsdC mutant reveals global regulation of secondary metabolic gene clusters in Aspergillus flavus. Microbiol. Res. 2016, 182, 150–161. [Google Scholar] [CrossRef] [PubMed]
  155. Lohmar, J.M.; Harris-Coward, P.Y.; Cary, J.W.; Dhingra, S.; Calvo, A.M. rtfA, a putative RNA-Pol II transcription elongation factor gene, is necessary for normal morphological and chemical development in Aspergillus flavus. Appl. Microbiol. Biotechnol. 2016, 100, 5029–5041. [Google Scholar] [CrossRef] [PubMed]
  156. Lohmar, J.M.; Puel, O.; Cary, J.W.; Calvo, A.M. The Aspergillus flavus rtfA gene regulates plant and animal pathogenesis and secondary metabolism. Appl. Environ. Microbiol. 2019, 85, e02446-18. [Google Scholar] [CrossRef] [Green Version]
  157. Cary, J.W.; Harris-Coward, P.; Scharfenstein, L.; Mack, B.M.; Chang, P.K.; Wei, Q.; Lebar, M.; Carter-Wientjes, C.; Majumdar, R.; Mitra, C.; et al. The Aspergillus flavus homeobox gene, hbx1, is required for development and aflatoxin production. Toxins 2017, 9, 315. [Google Scholar] [CrossRef] [Green Version]
  158. Cary, J.W.; Entwistle, S.; Satterlee, T.; Mack, B.M.; Gilbert, M.K.; Chang, P.K.; Scharfenstein, L.; Yin, Y.; Calvo, A.M. The transcriptional regulator Hbx1 affects the expression of thousands of genes in the aflatoxin-producing fungus Aspergillus flavus. G3 Genes Genomes Genet. 2019, 9, 167–178. [Google Scholar] [CrossRef] [Green Version]
  159. Kale, S.P.; Cary, J.W.; Baker, C.; Walker, D.; Bhatnagar, D.; Bennett, J.W. Genetic Analysis Of Morphological Variants of Aspergillus parasiticus Deficient in Secondary Metabolite Production. Mycol. Res. 2003, 107, 831–840. [Google Scholar] [CrossRef] [Green Version]
  160. Chang, P.K.; Scharfenstein, L.L.; Mack, B.; Ehrlich, K.C. Deletion of the Aspergillus flavus orthologue of A. nidulans fluG reduces conidiation and promotes production of sclerotia but does not abolish aflatoxin biosynthesis. Appl. Environ. Microbiol. 2012, 78, 7557–7563. [Google Scholar] [CrossRef] [Green Version]
  161. Hicks, J.K.; Yu, J.H.; Keller, N.P.; Adams, T.H. Aspergillus sporulation and mycotoxin production both require inactivation of the FadA Gα protein-dependent signaling pathway. EMBO J. 1997, 16, 4916–4923. [Google Scholar] [CrossRef]
  162. Shimizu, K.; Keller, N. Genetic Involvement of a cAMP-dependent Protein Kinase in a G Protein Signaling Pathway Regulating Morphological and Chemical Transitions in Aspergillus nidulans. Genetics 2001, 157, 591–600. [Google Scholar]
  163. Ramamoorthy, V.; Shantappa, S.; Dhingra, S.; Calvo, A. veA- dependent RNA-pol II transcription elongation factor-like protein, RtfA, is associated with secondary metabolism and morphological development in Aspergillus nidulans. Mol. Microbiol. 2012, 85, 795–814. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Linz, J.E.; Hong, S.Y.; Roze, L.V. Oxidative stress-related transcription factors in the regulation of secondary metabolism. Toxins 2013, 5, 683–702. [Google Scholar]
  165. Narasaiah, K.V.; Sashidhar, R.B.; Subramanyam, C. Biochemical analysis of oxidative stress in the production of aflatoxin and its precursor intermediates. Mycopathologia 2006, 162, 179–189. [Google Scholar] [CrossRef] [PubMed]
  166. Grintzalis, K.; Vernardis, S.I.; Klapa, M.I.; Georgiou, C.D. The role of oxidative stress in sclerotial differentiation and aflatoxin B1 biosynthesis in Aspergillus flavus. Appl. Environ. Microbiol. 2014, 80, 5561–5571. [Google Scholar] [CrossRef] [Green Version]
  167. Gomez del Arco, P.; Martínez-Martínez, S.; Calvo, V.; Armesilla, A.L.; Redondo, J.M. Antioxidants and AP-1 activation: A brief overview. Immunobiology 1997, 198, 273–278. [Google Scholar] [CrossRef]
  168. Miskei, M.; Karányi, Z.; Pócsi, I. Annotation of stress-response proteins in the Aspergilli. Fungal Genet. Biol. 2009, 46, S105–S120. [Google Scholar] [CrossRef] [Green Version]
  169. Amare, M.G.; Keller, N.P. Molecular mechanisms of Aspergillus flavus secondary metabolism and development. Fungal Genet. Biol. 2014, 66, 11–18. [Google Scholar] [CrossRef]
  170. Yang, K.; Qin, Q.; Liu, Y.; Zhang, L.; Liang, L.; Lan, H.; Chen, C.; You, Y.; Zhang, F.; Wang, S. Adenylate Cyclase AcyA Regulates Development, Aflatoxin Biosynthesis and Fungal Virulence in Aspergillus flavus. Front. Cell. Infect. Microbiol. 2016, 6, 190. [Google Scholar] [CrossRef] [Green Version]
  171. Yuan, J.; Chen, Z.; Guo, Z.; Li, D.; Zhang, F.; Shen, J.; Zhang, Y.; Wang, S.; Zhuang, Z. PbsB regulates morphogenesis, aflatoxin B1 biosynthesis, and pathogenicity of Aspergillus flavus. Front. Cell. Infect. Microbiol. 2018, 8, 1–10. [Google Scholar] [CrossRef] [Green Version]
  172. Chang, P.K.; Scharfenstein, L.L.; Luo, M.; Mahoney, N.; Molyneux, R.J.; Yu, J.; Brown, R.L.; Campbell, B.C. Loss of msnA, a putative stress regulatory gene, in Aspergillus parasiticus and Aspergillus flavus increased production of conidia, aflatoxins and kojic acid. Toxins 2011, 3, 82–104. [Google Scholar] [CrossRef] [Green Version]
  173. Guan, X.; Zhao, Y.; Liu, X.; Shang, B.; Xing, F.; Zhou, L.; Wang, Y.; Zhang, C.; Bhatnagar, D.; Liu, Y. The bZIP transcription factor Afap1 mediates the oxidative stress response and aflatoxin biosynthesis in Aspergillus flavus. Rev. Argent. Microbiol. 2019, 51, 292–301. [Google Scholar] [CrossRef] [PubMed]
  174. Roze, L.V.; Chanda, A.; Wee, J.; Awad, D.; Linz, J.E. Stress-related transcription factor AtfB integrates secondary metabolism with oxidative stress response in Aspergilli. J. Biol. Chem. 2011, 286, 35137–35148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Reverberi, M.; Zjalic, S.; Ricelli, A.; Punelli, F.; Camera, E.; Fabbri, C.; Picardo, M.; Fanelli, C.; Fabbri, A.A. Modulation of antioxidant defense in Aspergillus parasiticus is involved in aflatoxin biosynthesis: A role for the ApyapA gene. Eukaryot. Cell 2008, 7, 988–1000. [Google Scholar] [CrossRef] [Green Version]
  176. Weydert, J.C.; Cullen, J. Measurement of Superoxide Dismutase, Catalase and Glutathione Peroxidase in cultured cells and tissue. Nat. Protoc. 2011, 5, 51–66. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. He, Z.-M.; Price, M.S.; Obrian, G.R.; Georgianna, D.R.; Payne, G.A. Improved protocols for functional analysis in the pathogenic fungus Aspergillus flavus. BMC Microbiol. 2007, 7, 104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Sun, Q.; Shang, B.; Wang, L.; Lu, Z.; Liu, Y. Cinnamaldehyde inhibits fungal growth and aflatoxin B1 biosynthesis by modulating the oxidative stress response of Aspergillus flavus. Appl. Microbiol. Biotechnol. 2015, 100, 1355–1364. [Google Scholar] [CrossRef]
  179. Reverberi, M.; Fabbri, A.A.; Zjalic, S.; Ricelli, A.; Punelli, F.; Fanelli, C. Antioxidant enzymes stimulation in Aspergillus parasiticus by Lentinula edodes inhibits aflatoxin production. Appl. Microbiol. Biotechnol. 2005, 69, 207–215. [Google Scholar] [CrossRef]
  180. Leiter, É.; Park, H.-S.; Kwon, N.-J.; Han, K.-H.; Emri, T.; Oláh, V.; Mészáros, I.; Dienes, B.; Vincze, J.; Csernoch, L.; et al. Characterization of the aodA, dnmA, mnSOD and pimA genes in Aspergillus nidulans. Sci. Rep. 2016, 6, 20523. [Google Scholar] [CrossRef] [Green Version]
  181. Maggio-Hall, L.A.; Keller, N.P. Mitochondrial β-oxidation in Aspergillus nidulans. Mol. Microbiol. 2004, 54, 1173–1185. [Google Scholar] [CrossRef]
  182. Reverberi, M.; Punelli, M.; Smith, C.A.; Zjalic, S.; Scarpari, M.; Scala, V.; Cardinali, G.; Aspite, N.; Pinzari, F.; Payne, G.A.; et al. How Peroxisomes Affect Aflatoxin Biosynthesis in Aspergillus flavus. PLoS ONE 2012, 7, e48097. [Google Scholar] [CrossRef]
  183. Chanda, A.; Roze, L.V.; Kang, S.; Artymovich, K.A.; Hicks, G.R.; Raikhel, N.V.; Calvo, A.M.; Linz, J.E. A key role for vesicles in fungal secondary metabolism. Proc. Natl. Acad. Sci. USA 2009, 106, 19533–19538. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Hynes, M.J.; Murray, S.L.; Khew, G.S.; Davis, M.A. Genetic analysis of the role of peroxisomes in the utilization of acetate and fatty acids in Aspergillus nidulans. Genetics 2008, 178, 1355–1369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Hong, S.Y.; Linz, J.E. Functional Expression and Sub-Cellular Localization of the Early Aflatoxin Pathway Enzyme Nor-1 in Aspergillus parasiticus. Mycol. Res. 2009, 113, 591–601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Martin, J.F.; Van Den Berg, M.A.; Ver Loren Van, T.; Liras, P. Sensing and transduction of nutritional and chemical signals in filamentous fungi: Impact on cell development and secondary metabolites biosynthesis. Biotechnol. Adv. 2019, 37, 107392. [Google Scholar] [CrossRef] [PubMed]
  187. Tsitsigiannis, D.I.; Keller, N.P. Oxylipins act as determinants of natural product biosynthesis and seed colonization in Aspergillus nidulans. Mol. Microbiol. 2006, 59, 882–892. [Google Scholar] [CrossRef]
  188. Dos Reis, T.F.; Mellado, L.; Lohmar, J.M.; Lilian, P.S.; Zhou, J.-J.; Calvo, A.M.; Goldman, G.H.; Brown, N.A. GPCR-mediated glucose sensing system regulates light-dependent fungal development and mycotoxin production. PLoS Genet. 2019, 15, e1008419. [Google Scholar] [CrossRef]
Figure 1. Schematization of the different factors linked with aflatoxin production.
Figure 1. Schematization of the different factors linked with aflatoxin production.
Toxins 12 00150 g001
Figure 2. Organization of the aflatoxin gene cluster including the old and new cluster gene nomenclatures. This figure was adapted from the works of [24,33,37]. Red dotted lines represent the binding sites of AflR in the above pathway.
Figure 2. Organization of the aflatoxin gene cluster including the old and new cluster gene nomenclatures. This figure was adapted from the works of [24,33,37]. Red dotted lines represent the binding sites of AflR in the above pathway.
Toxins 12 00150 g002
Figure 3. Schematization of the main intermediates produced during the AFB1 biosynthetic pathway and the confirmed or putative (indicated by?) level of intervention of the genes belonging to the AFB1 cluster. This figure is adapted from the works of [23,24,33,46,47,48,49,50,51,52,53,54,55].
Figure 3. Schematization of the main intermediates produced during the AFB1 biosynthetic pathway and the confirmed or putative (indicated by?) level of intervention of the genes belonging to the AFB1 cluster. This figure is adapted from the works of [23,24,33,46,47,48,49,50,51,52,53,54,55].
Toxins 12 00150 g003
Figure 4. Schematic representation of a putative map of genes involved in aflatoxin regulation in diverse Aspergillus species.
Figure 4. Schematic representation of a putative map of genes involved in aflatoxin regulation in diverse Aspergillus species.
Toxins 12 00150 g004
Table 1. Genes involved in carbon catabolic repression.
Table 1. Genes involved in carbon catabolic repression.
Carbon Catabolic Repression
GenesCoding Function
creACys2His2 Zinc finger transcription repressor
creBCys2His2 Zinc finger ubiquitin processing protease
creCCys2His2 Zinc finger ubiquitin processing protease
alcRRegulon specific transcription factor
aldAAldehyde dehydrogenase
alcAAlcohol dehydrogenase
Relation between genes or corresponding proteins in Aspergilli spp.:
creA represses the expression of aclR while the latter is a positive regulatory factor for the genes aldA and alcA [101,103].
CreB-CreC form a protein complex that is essential for CreA function and stability [104].
Demonstration of the connection with AF/ST synthesis
Gene expression in the AF/ST gene cluster is regulated either positively or negatively by CreA depending on the fungal species [31,105].
In A. flavus:
ΔcreA strains produced insignificant amounts of aflatoxin in complete medium and their ability to colonize hosts was also impaired [102].
The repression of creA by gallic acid lead to the inhibition of AFB1 production [106].
In A. parasiticus:
Several genes in the AFB1 cluster were demonstrated to have CreA-sites near their promoter regions [107].
Table 2. Genes involved in nitrogen utilization.
Table 2. Genes involved in nitrogen utilization.
Nitrogen Source
GenesCoding Function
areAZinc finger mediating nitrogen metabolite repression
nmrARepressive nitrogen
meaBRegulatory protein
niiANitrite reductase
niaDNitrate reductase
Relation between genes or corresponding proteins in Aspergilli spp.:
areA and meaB are regulatory genes that mediate nitrogen metabolite repression [108,109]
AreA regulates nitrate transporters binding at the intergenic regions of niiA and niaD [110]
MeaB is a regulatory factor that activates NmrA, which is a repressor of AreA [109].
Demonstration of the connection with AF/ST synthesis
InA. flavus:
areA was recently identified in A. flavus and demonstrated to be partially involved in aflatoxin production. The lack and overexpression of areA resulted in increased/decreased amounts of aflatoxin depending on the nitrogen source media [111].
The direct implication of AreA in aflS expression has been demonstrated in the presence of different sources of nitrate and ammonium [112].
Overexpressing meaB strains lost their capacity to produce AFs on seeds while ΔmeaB produced only a few aflatoxins [109].
niiA is one of the genes that does not belong to the AFB1 gene cluster but that is presumed to be regulated by the major internal AF regulator aflR [26].
In A. parasiticus:
AreA binds to the GATA sites of the promoters of aflR/aflS [110,113].
aflC and aflD genes are expressed on ammonium and peptone media but not on nitrate sources [114].
In A. nidulans:
Nitrate medium increases the production of sterigmatocystin while ammonium medium does not [114].
Table 3. Genes involved in pH modulation.
Table 3. Genes involved in pH modulation.
pH Impact
GeneCoding Function
pacCCys2His2 (C2-H2)- Zinc finger pH regulator
Relation between genes or corresponding proteins in Aspergilli spp.:
The transcription factor PacC is strongly expressed under alkaline conditions [117].
Demonstration of the connection with AF/ST synthesis
InA. flavus:
An increase in pH in nitrate-based medium results in lower levels of AF, while lower pH (4.0) resulted in a 10-fold increase in production of AF [119].
InA. parasiticus:
pH levels below 6.0 promote B-type aflatoxin production while higher levels induce G-type production [120].
A putative PacC binding site was identified in the aflR promoter region, but interestingly, alkaline conditions in which pacC is activated do not support AFB1 production [113].
aflM was higher expressed in acidic media than in neutral and alkaline media. Fungal growth reduced the pH of the medium and increased AF production with time [121].
In A. nidulans:
Acidic conditions are more favorable for ST biosynthesis than neutral or alkaline ones [121].
pacC-mutant strains produced 10-fold less ST than the control. The expression level of stcU (aflM homologous) was lower in increased pH media [121].
Table 4. Genes involved in the light response.
Table 4. Genes involved in the light response.
Light
GenesCoding function
veAGlobal regulator
vapBMethyltransferase
vipCMethyltransferase
fphAPhytochrome-like red light receptor
kapAα transport carrier
velBVelvet-like protein B
laeAPutative methyltransferase
velCVelvet-like protein C
lreABlue-light sensing protein
lreBBlue-light sensing protein
llmFLaeA-like methyltransferase
Relation between genes or corresponding proteins in Aspergilli spp.:
Vea-LaeA-VelB form a trimeric complex called the velvet complex. VeA interacts with LaeA in the nucleus and with VelB in the cytoplasm and the nucleus. This trimeric complex, together with other light-receptor proteins, perceives light signals and is an essential coordinator of secondary metabolism and fungal development [124,125,126,127,128].
In dark conditions, the nuclear localization of VeA increases its interaction with LaeA to enhance production of secondary metabolites but also with VelB to induce sexual development [129].
FphA interacts with VeA, with LreB and LreA in the nucleus [123,130] and this protein complex is involved in red and blue light sensing [124].
KapA and VeA physically interact in dark conditions. KapA supports the entry of the VeA-VelB complex into the nucleus [126].
Both VelC and VelB form a protein dimer with VosA (involved in spore viability) [126].
VipC and VapB reduce the nuclear accumulation of VeA, thereby reducing secondary metabolism [131].
LlmF interacts with VeA controlling its subcellular location [129].
Demonstration of the connection with AF/ST synthesis
InA. flavus:
veA is essential for AFB1 production [132,133].
In null mutants of veA or laeA, no aflR expression was observed [134].
ΔlaeA reduced aflR, aflD and aflS mRNA levels with no AF production [135].
Overexpression of laeA results in higher levels of AFB1 whereas ΔlaeA strains do not affect aflatoxin production [109].
VeA governs 28 out of the 56 secondary metabolite gene clusters including the AF cluster [136].
In A. parasiticus:
Deletion of veA resulted in the loss of the aflatoxin intermediate Versicolorin A. VeA is required for aflR/aflS expression [137].
In A. nidulans:
ΔveA or ΔveA/vipC mutants were unable to produce sterigmatocystin but ΔvipC did not affect the mycotoxin production [131].
Both sexual development and sterigmatocystin production were repressed in light conditions [138].
ΔveA strains resulted in no aflR expression and neither ST production [139].
Deletion of the laeA gene inhibits aflR and stcU expression [140].
LreA, LreB and FphA modulate sterigmatocystin biosynthesis depending on light and on the presence of glucose [141].
FphA represses sexual development and ST production while LreA and LreB stimulate both processes. Blue light represses ST production while red light has the opposite effect [124].
LlmF is a negative factor for ST production [129].
Table 5. Genes involved in sexual development.
Table 5. Genes involved in sexual development.
Sexual Development
GenesCoding Function
MAT1-1Mating type (alpha)
MAT1-2Mating type (HMG)
Relation between genes or corresponding proteins in Aspergilli spp.:
Either MAT1-1 or MAT1-2 is expressed in A. flavus and A. parasiticus strains and involved in sexual development [145].
Demonstration of the connection with AF synthesis
InA. flavus:
Strains of sexually developed A. flavus demonstrated that production of AF is highly heritable. In asexual development, non-aflatoxigenic populations are maintained while aflatoxigenicity increases in sexual development [146].
Table 6. Genes involved in asexual development.
Table 6. Genes involved in asexual development.
Asexual Development
GenesCoding Function
fadAα-subunit of heterotrimeric G-protein
fluGDevelopmental regulator
brlAC2H2 zinc finger protein transcriptional activator of conidiophore
abaATranscription factor for conidia formation
wetADevelopmental regulatory protein
nsdCZinc-finger transcription factor
nsdDZinc-finger transcription factor
pkaACatalytic subunit of protein kinase A
flbARGS protein/developmental regulator
flbBbZIP-type transcription factor
flbCPutative C2H2 conidiation transcription factor
flbDMYB family conidiophore development
flbEDevelopmental regulator
vosASpore viability/Developmental regulator/Trehalose production
rtfARNA-pol II transcription elongation factor-like protein
hbx1Homebox transcription factor
Relation between genes or corresponding proteins in Aspergilli spp.:
fluG activates flbA which in turn represses fadA signaling [147].
fadA up-regulates pkaA [97].
flbA is a regulator of flbB, which regulates flbC, flbD, flbE (flb genes) [148].
flb genes are required with fluG for the correct expression of brlA [149].
brlA is a negative regulator of abaA and abaA is a repressor of wetA [126,150,151].
brlA is regulated by veA [139].
vosA is a repressor of brlA [152]. VosA also forms a protein-complex with VelB and VelC (velvet proteins) [126,127].
nsdC and nsdD are repressors of brlA [153].
Demonstration of the connection with AF/ST synthesis
InA. flavus:
Expression of aflD, aflM, and aflP is strongly reduced in nsdC deleted mutants. Loss of NsdC or NsdD resulted in developmental alterations that impact the ability of AflR to activate the expression of AF biosynthesis genes [153]. In fact, A. flavus nsdC mutants are unable to produce AF or any other secondary metabolites [154].
FadA governs both AF and ST biosynthesis [74,97].
Deleted strains of rtfA greatly reduced AFB1 biosynthesis but interestingly, aflR, aflM and aflP were up-regulated within the 3 first days of incubation and then down-regulated on the 4th and 5th day. Moreover, rtfA controls veA and laeA expression [155] and production of other secondary metabolites [156].
Disrupted hbx1 strains produced null B-type aflatoxin, conidia and sclerotia [157].
Several genes belonging to the aflatoxin cluster were not expressed in hbx1 mutant strains [158].
In A. parasiticus:
Mutants defective in conidiation processes also had reduced levels of AF production [159].
FadA is presumed to regulate aflT, the MFS belonging to the AFB1 gene cluster [74].
In A. nidulans:
FadA up-regulates pkaA, which down-regulates conidiation. PkaA also inhibits AflR activity by phosphorylation [77].
fluG-deleted strains lost their ability to produce sterigmatocystin [160].
Mutations in flbA and fluG blocked both sterigmatocystin production and sporulation [161].
ΔfadA and ΔpkaA mutants failed in normal processes of conidiation and sterigmatocystin biosynthesis [162].
Co-regulation of brlA and aflR by the fadA signaling pathway genes was reported [139].
Sterigmatocystin production was positively regulated by rtfA [163].
Table 7. Genes involved in oxidative stress response.
Table 7. Genes involved in oxidative stress response.
Oxidative Stress Complex
GenesCoding Function
bZIP transcription factors
ap-1bZIP transcription factor
atfAbZIP transcription factor
atfBbZIP transcription factor
Stress Response Signaling Pathway
srrATranscription factor
msnATranscription factor
acyAAdenylate Cyclase
pbsBMAP kinase kinase
Relation between genes or corresponding proteins in Aspergilli spp.:
In cell systems, the ap-1 gene is activated under both, pro-oxidant and antioxidant conditions [167].
Together, AtfB, SrrA, Ap-1, PbsB and MsnA form a regulatory network involved in oxidative stress response and secondary metabolite production [67,168].
AtfA may interact with AtfB in response to oxidative stress [67,169].
acyA was shown to regulate AF biosynthesis as well as to intervene in hyperosmotic and oxidative stress [170].
pbsB was reported to be involved in stress responses and AFB1 biosynthesis [171].
Demonstration of the connection with AF/ST synthesis
InA. flavus:
The deleted ΔacyA strains were unable to produce aflatoxin in contrast to the control and some genes in the AF gene cluster (aflR and aflO) were also down-regulated [170].
PbsB positively regulates AFB1 production through aflR, the major regulator of the AFB1 gene cluster, as well as other genes such as aflC, aflD, aflK and aflQ [171].
msnA deletion results in 50% more aflatoxins as well as higher levels of reactive oxygen species (ROS) [172].
ap1 deletion reduces aflatoxin production while the expression of aflM and aflP was down-regulated even though aflR was up-regulated [173].
In A. parasiticus:
AtfB binds to promoters of seven genes belonging to the AF gene cluster [174].
ap-1 deletion increases AF production while the Ap-1 protein binds to the promoter region of the aflR gene [175].
Table 8. Genes coding for fungal superoxide dismutases and catalases.
Table 8. Genes coding for fungal superoxide dismutases and catalases.
Superoxide Dismutases and Catalases
GenesCoding Function
mnSODManganese superoxide dismutase
sod1Cu, Zn superoxide dismutase
catAConidia-specific catalase
catBMycelial catalase
hyr1Glutathione peroxidase
Demonstration of the connection with AF/ST synthesis
InA. flavus:
mnSOD and the genes aflA, aflM, and aflP belonging to the AF gene cluster are co-regulated [67].
Deletion of sod reduced AF production [177].
Increased expression of the genes of catA, cat2 and sod1 as well as an increased CAT enzymatic activity were observed in the presence of the AF inhibitor piperine [15]. The same increase in CAT activity was also observed with another AF inhibitor, cinnamaldehyde [178].
In A. parasiticus:
Inhibition of AF production induced by Lentinula edodes increases SOD enzymatic activity within its mechanism of action [179].
In A. nidulans:
Deletion of mnSOD increased both glutathione reductase and catalase activities while its overexpression reduced the activity of catalase but increased SOD activity [180].
Table 9. Genes involved in β-oxidation.
Table 9. Genes involved in β-oxidation.
β-oxidation
GenesCoding Function
pexKExistence of peroxisome
pex11Peroxisome proliferation
foxARegulation of fatty acid metabolism by β -oxidation
rab7Vesicle marker
vps16Vacuole marker
Relation between genes or corresponding proteins in Aspergilli spp.:
In filamentous fungi, peroxisomes are crucial for primary metabolism and play a role in the formation of some secondary metabolites [182].
Aflatoxisomes (aflatoxin vesicles) are partially regulated by VeA [183].
pex mutants are able to grow on acetate medium but their growth is affected by fatty acids, indicating β-oxidation enzymes require a peroxisomal location [184].
Demonstration of the connection with AF/ST synthesis
InA. flavus:
An increase in the number of peroxisomes enhances AFB1 production [182].
InA. parasiticus:
An increase in the number of vesicles is positively correlated with AF accumulation/export [183].
Nor-1 (protein involved in the norsolorinic acid biosynthesis) mainly occurs in the cytoplasm and vacuoles [185].
Table 10. Genes involved in cell signaling.
Table 10. Genes involved in cell signaling.
Cell Signaling
GenesCoding Function
ppoA(oxylipin) Dioxygenase
ppoB(oxylipin) Dioxygenase
ppoC(oxylipin) Dioxygenase
ppoD(oxylipin) Dioxygenase
lox(oxylipin) Lipoxygenase
gprKGPCR
gprAGPCR
gprPGPCR
rasAGTP-binding protein
Relation between genes or corresponding proteins in Aspergilli spp.:
GPCRs are involved in oxylipin response [30].
Demonstration of the connection with AF/ST synthesis
InA. flavus:
The GPCRs were shown to interact with AFB1 synthesis and its precursor, sterigmatocystin (ST). Deletion of gprK and grpA increased AF production compared to the control strain [30].
When all four ppo genes and the lox gene were disrupted simultaneously, the mutant strains showed reduced conidiation and increased AF production on maize and peanut seeds [169].
In A. nidulans:
ΔppoA; ΔppoB and ΔppoC mutants are unable to produce ST [187].
RasA has been demonstrated to control aflR activity [77].
GprH and GprM are negative regulators of ST biosynthesis [188].

Share and Cite

MDPI and ACS Style

Caceres, I.; Al Khoury, A.; El Khoury, R.; Lorber, S.; P. Oswald, I.; El Khoury, A.; Atoui, A.; Puel, O.; Bailly, J.-D. Aflatoxin Biosynthesis and Genetic Regulation: A Review. Toxins 2020, 12, 150. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins12030150

AMA Style

Caceres I, Al Khoury A, El Khoury R, Lorber S, P. Oswald I, El Khoury A, Atoui A, Puel O, Bailly J-D. Aflatoxin Biosynthesis and Genetic Regulation: A Review. Toxins. 2020; 12(3):150. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins12030150

Chicago/Turabian Style

Caceres, Isaura, Anthony Al Khoury, Rhoda El Khoury, Sophie Lorber, Isabelle P. Oswald, André El Khoury, Ali Atoui, Olivier Puel, and Jean-Denis Bailly. 2020. "Aflatoxin Biosynthesis and Genetic Regulation: A Review" Toxins 12, no. 3: 150. https://0-doi-org.brum.beds.ac.uk/10.3390/toxins12030150

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

Article Metrics

Back to TopTop