Next Article in Journal
Effectiveness of Copper Nanoparticles in Wound Healing Process Using In Vivo and In Vitro Studies: A Systematic Review
Previous Article in Journal
Assuring the Biofunctionalization of Silicone Covalently Bonded to Rhamnolipids: Antibiofilm Activity and Biocompatibility
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

How to Completely Squeeze a Fungus—Advanced Genome Mining Tools for Novel Bioactive Substances

Institute of Microbial Genetics, Department of Applied Genetics and Cell Biology, University of Natural Resources and Life Sciences Vienna, A-3430 Tulln/Donau, Austria
*
Author to whom correspondence should be addressed.
Pharmaceutics 2022, 14(9), 1837; https://doi.org/10.3390/pharmaceutics14091837
Submission received: 29 July 2022 / Revised: 23 August 2022 / Accepted: 29 August 2022 / Published: 31 August 2022
(This article belongs to the Special Issue The Role of Fungi in Pharmaceutical Biotechnology)

Abstract

:
Fungal species have the capability of producing an overwhelming diversity of bioactive substances that can have beneficial but also detrimental effects on human health. These so-called secondary metabolites naturally serve as antimicrobial “weapon systems”, signaling molecules or developmental effectors for fungi and hence are produced only under very specific environmental conditions or stages in their life cycle. However, as these complex conditions are difficult or even impossible to mimic in laboratory settings, only a small fraction of the true chemical diversity of fungi is known so far. This also implies that a large space for potentially new pharmaceuticals remains unexplored. We here present an overview on current developments in advanced methods that can be used to explore this chemical space. We focus on genetic and genomic methods, how to detect genes that harbor the blueprints for the production of these compounds (i.e., biosynthetic gene clusters, BGCs), and ways to activate these silent chromosomal regions. We provide an in-depth view of the chromatin-level regulation of BGCs and of the potential to use the CRISPR/Cas technology as an activation tool.

Graphical Abstract

1. Introduction

1.1. Fungal Secondary Metabolites as Biopharmaceuticals of Tomorrow

Bacteria, plants, and filamentous fungi are capable of producing many chemically diverse bioactive substances, referred to as natural products or secondary metabolites (SMs) [1]. Probably the most prominent of these fungal bioactive compounds is penicillin from Penicillium rubens, discovered by Alexander Fleming in 1929 [2,3] which initiated the “antibiotic revolution” that lead to a major paradigm shift in the treatment of patients [4]. Many fungi and their SMs made it into pharmaceutical application, such as cyclosporin from Tolypocladium inflatum which is used as an immunosuppressant [5,6], or lovastatin as a treatment for coronary heart diseases produced by Aspergillus terreus [7]. Other fields of application for microbial-derived natural products are the food industry (e.g., soy sauce, sake, and miso by Aspergillus oryzae or citric acid by Aspergillus niger [8]), the agricultural sector (biocontrol agents [9,10,11], growth hormones [12]) and many others [13,14,15,16]. The ever-rising demand due to the emergence of multi-resistant pathogens and resistant cancer types underscores the importance of mining new bioactive substances from fungal sources [17].
SMs are per definition not essential for growth but may provide selective advantages for the producing organism upon specific environmental challenges: e.g., pH stress (acids and alkaline substances [18]), temperature stress (regulators of membrane fluidity [19]), oxidative stress (antioxidants [20]), UV protection (pigments [21]), microbial warfare (antimicrobials [22]), communication (e.g., quorum sensing via homoserine lactones [23]), pathogenicity and host-infection (effectors or virulence factors [24,25]), support of ion uptake (sideophore production for iron acquisition [26]). As versatile as these functions are, as diverse are the chemical properties of such natural products. Not surprisingly, most of them are only produced under defined environmental or developmental conditions implying a complex regulation at the molecular level [27].
The blueprints for such bioactive substances are stored in biosynthetic genes that are, in most cases, physically linked in fungal genomes and together with genes for regulators and transporters sometimes constitute very large BGCs [28,29]. These BGCs are silenced by default through heterochromatic structures at the chromosomal level, but once activated by the specific signal, they work similar to an orchestrated factory to produce the cluster-specific compound. Bioinformatic tools have been developed that predict fungal BGCs based on whole-genome sequence data and these predictions show that so far sequenced fungi harbor between ~15 and 50 BGCs, but up to more than 80 BGCs have also been predicted in some genomes [28,30]. However, as outlined above, most of them are transcriptionally silent under standard laboratory conditions and hence the majority of associated natural products are not known [27].

1.2. Genomic Organization and Evolution of BGCs

Depending on the so-called backbone or key enzyme(s) involved in the biosynthesis, SMs are categorized according to the chemical families. They belong to e.g., polyketides, non-ribosomal peptides, terpenes and alkaloids or prenylated tryptophan derivates. Accordingly, they are synthesized by polyketide synthases (PKSs; further divided into highly reducing (HR-PKS) and non-reducing (NR-PKS)), non-ribosomal peptide synthetases (NRPSs), terpene cyclases (TCs) and dimethylallyltryptophan synthases (DMATSs) or hybrids of them [28,29].
Next to the backbone enzyme-encoding gene(s) BGCs often harbor genes involved in self-resistance against a potentially toxic SM produced by the BGC, transport of precursors, intermediates, or the resulting compound [31,32], transcriptional activation of the biosynthetic genes, and often so-called tailoring or decorating enzymes that modify the structure of the chemical scaffold compound released by the backbone enzyme (e.g., P450 monooxygenases, methyltransferases, hydroxylases, epimerases, etc. [29]).
A few exceptions exist that deviate from this standard configuration. For example, in Fusarium graminearum, polyketide synthase PKS8 was shown to be a stand-alone gene that is responsible for the synthesis of gibepyrone A [33], while the kojic acid BGC from Aspergillus oryzae does not have a backbone enzyme-encoding gene at all [34]. In other cases, the formation of superclusters was observed which encode for several independently synthesized biosynthetic products [35]. The opposite was also observed, where several BGCs or genes from distinct genomic locations communicate with each other to synthesize one product [36].
Fungi exhibit a very diversified and specialized chemical repertoire. The diversification of BGCs increases with the phylogenetic distance [37,38] but in some cases, it can even be observed within the same or closely related species, which was shown for Aspergillus fumigatus [39,40], Aspergillus niger [41] as well as between A. oryzae and its close relative Aspergillus flavus [42]. Constant interactions with other microorganisms and environmental conditions promote the intra and interspecies diversification process by exerting positive or negative selective pressure on BGCs in that specific ecosystem. If a BGC is non-essential or not beneficial in a certain microenvironment it is likely to acquire deleterious mutations or even gene loss over time [43]. A comparison of isolates from the same species but separated microenvironments revealed the emergence of different ecotypes as observed during population genetics studies on Streptomyces [44] and Umbilicaria pustulata [45].
BGCs are differently affected by evolution compared to the rest of the genome. They are located outside of so-called syntenic blocks, i.e., their position within the genome, and the BGC itself is less conserved than the rest of the genome [46,47]. In support of this, many BGCs were found to be enriched in sub-telomeric regions which are known to be hotspots for recombination [48]. This was shown in Sclerotinia sclerotium [49], Magnaporthe oryzae [50,51], A. fumigatus [52], and Zymoseptoria tritici [53].
From an evolutionary perspective, the ability to quickly adapt to new environmental challenges is fundamental for the survival of an organism. In the case of fungi and other microbes, this means that they can adapt and change their metabolomic arsenal fast enough to properly compete against other microbes or environmental changes. This implies that the relevant genomic regions and hence the genes situated in these regions (i.e., BGCs) experience regular mutations (single nucleotide polymorphisms (SNPs) such as missense, nonsense, frameshift, or gain/loss of function) and genomic rearrangements as for instance gene duplications, gene or BGC loss, BGC rearrangements, pseudogenization events and also the acquisition of genetic material via horizontal gene transfer [43]. Organisms that acquire beneficial mutations will surpass others that experience disadvantageous ones whereby the environmental context decides if a certain mutation is beneficial or not. Microbial competition is not only about having the “best weaponry” but also about defense, resource efficiency, and denying beneficial effects to competitors. This not only implies gain of function or the de novo assembly of a BGC but also reductive genomic evolution which led to the “Black Queen Hypothesis” (BQH) [54]. The BQH considers that selective pressure can lead to the loss of a gene or function if that gene is dispensable (which makes it more energy efficient not to have it), or if the gene product would give a competitor more benefits than the producing organism itself.
For the artificial activation of silent BGCs in a laboratory context, this means that some of the BGCs will carry deleterious mutations or even miss whole genes that would be necessary for the biosynthesis of the BGC product. In these cases, solely transcriptional activation will not suffice. Close examination of the BGC of interest can already unveil if genes carry a premature stop codon or if homologous BGCs indicate missing genes or frameshift mutations. A possible workaround in these scenarios is the exchange of single genes by their homologs or the genetic repair of a nonsense mutation as performed by Burkhardt et al. with the STC5 gene in Fusarium fujikuroi [55].
Multiple genomic, inter-species, and environmental factors contribute to this elevated diversification and evolution of BGCs. The occurrence of mutations and genomic rearrangements depend on many exogenous (e.g., radiation, heat, chemicals, environmental conditions, …) and endogenous factors (e.g., errors in DNA damage repair, or active transposable elements). As stated above, BGCs and pathogenicity-related genes are often in polymorphic regions that are prone to such events (e.g., sub-telomeric regions). Transposable elements (TEs) play an important role during evolution and also in the de novo assembly of BGCs. In Verticillium dahliae it was found that BGC and pathogenicity-related genes are often located close to transposable elements [56]. In a comparative study within Aspergillus fumigatus isolates, horizontally transferred BGCs, BGC idiomorphs (i.e., different BGCs in the same genomic location), and BGCs that are present in different genomic locations within the same or closely related species) are often flanked by TEs [39]. Similar findings were also found between A. oryzae and A flavus where two strain isolates had different BGCs in the same genomic locus [42] or in the case of several Botrytis species where the BOT BGC, involved in botrydial biosynthesis, was present at four different genomic locations across seven examined species [57,58]. When analyzing BGC genes and TE co-occurrence in Sclerotinia, they found that the overall co-occurrence of BGC genes and TEs is not significantly enhanced. Upon closer examination, however, it became evident that a subset of BGC genes, which is enriched in key-enzyme encoding genes, is co-occurrent with TEs and that genes in this subset often derive from recent gene duplication events [49]. BGC genes near TEs, therefore, were found less conserved than BGC genes more distal to the TE. All these findings point towards a TE-dependent heterogeneous evolution of BGCs (i.e., genes near TEs have an increased “evolutionary speed”). In the case of BGC prediction, TEs are, to our knowledge, not yet a considered factor, although it seems that these motives could very likely be good indicators for the presence of a BGC. Furthermore, TEs are probably also important factors for the de novo assembly of BGCs.
De novo assembly is hypothesized to be a two-step process. The first step starts with the framework establishment of the new metabolic pathway [59]. Often, this is accomplished by horizontal gene transfer, as shown for the sterigmatocystin cluster from A. nidulans to Podospora [60], or the transfer of the bikaverin BGC from Fusarium to Botrytis [61] and many others [59,62,63,64,65,66]). Other mechanisms involve the recruitment of host genes or gene duplication events (e.g., fatty acid synthase as ancestor of PKS enzymes [67]). In a second step, the BGC is clustered by genomic rearrangements. Sometimes, however, it also happens that two separate BGCs, which are not physically linked, are responsible for the biosynthesis of one compound (i.e., cluster crosstalk) [36] or that superclusters (e.g., fumagillin/pseurotin cluster from Aspergillus fumigatus [35]) are formed. Cluster crosstalk could be observed with the trichothecene supercluster of F. graminearum which consists of a 12 gene BGC and a separated 2 gene BGC plus one single gene at three different genomic loci [68].
All in all, the genomic comparative studies from A. fumigatus [39], A. flavus and A. oryzae [42], S. sclerotiorum [49], V. dahlia [56], and also the reviews from Lind et al. [39] and Rokas et al. [43] give very good insights into the evolution of BGCs and especially the role of TEs as one of their driving forces.
The fact that this evolutionarily driven arms-race is not only apparent between closely related species but also between different ecotypes of the same species illustrates that BGCs are evolving at a fast pace. This highlights that the research pool of fungal genera and species must be expanded to increase the chance of finding novel natural products in the future. However, most of the current research is focused on mining BGCs in only a few but very well-described genera, the so-called model organisms.
In this review, we highlight the concepts and technical possibilities on how to decrypt the secondary metabolome of a given (filamentous) fungus by efficiently deploying targeted and untargeted BGC-activation methods in synergy with genome mining. Special attention will be paid to more recent strategies involving the chromatin regulation of BGCs and the new possibilities currently emerging by the combination of CRISPR/Cas technology and transcriptional activators, global regulators, or chromatin-modifying enzymes.

2. Deciphering the Genomic Potential of Fungal Genomes for Secondary Metabolites

In silico-prediction of BGCs within the genome of the target fungus gives important information about the (theoretical) chemical potential of a fungus. This is not only important for deciding if a species is a promising candidate for novel bioactive compounds but is also a prerequisite for linking those compounds to a specific cluster and also for targeted BGC activation. Bioinformatic analyses of high-quality genome sequences and transcriptome datasets are becoming cornerstones to explore the theoretical chemical diversity of SMs in a given species. Those resources are routinely used today for gene annotation [69], BGC prediction [70], cluster phylogeny [28,71,72], prediction of the metabolite structure and bioactivity [73], drug target predictions [74,75] and also dereplication [76]. As outlined in Section 1.2, there are many different data types and sources that can be utilized for the prediction of BGCs. In the following paragraph, we will describe the basic principle and the developmental cycle of BGC prediction tools.
Figure 1 illustrates the workflow of genome mining-assisted BGC prediction and revelation. It comprises the crucial steps that are necessary for the prediction and verification of a BGC and its corresponding metabolite. It also shows the pitfalls which can easily remain unnoticed but significantly impact the reliability of the prediction.

BGC Prediction Methods

The majority of BGC prediction tools can be classified into either high-confidence/low-novelty (rule-based) or low-confidence/high-novelty (rule-independent) [70,77].
High-confidence/low novelty methods reliably detect BGCs that are closely related to already verified cluster types (i.e., PKS; NRPS, etc.) by scanning the genome for their specific signatures (genes, gene types, genomic arrangement, etc.). These signatures are retrieved from experimentally verified data that links the compound to the respective BGC, which are then implemented into the algorithm. The prediction tool antiSMASH [78] is a very well-established and regularly updated platform that also integrates many other published algorithms. It can be used for fungal, plant, and bacterial genomes and contains methods for finding canonical and also novel BGCs. During the prediction of known BGC types, the procedure often starts with an initial screen for backbone enzyme-encoding genes which serve as anchor-point for further analysis of the region. In a second step, the cluster borders are determined by scanning the area around this anchor point for the presence of other BGC-related genes [78,79]. This means, however, that high-confidence/low-novelty (rule-based) methods can only detect BGCs that share similarities with already discovered and described cluster types
In contrast to this, low-confidence/high-novelty (rule-independent methods) pursue the aim of detecting BGCs of a novel yet unknown type that does not follow currently known patterns (gene types, arrangement, etc.). There are multiple strategies for finding new BGC types that use various experimental and -omics data for this purpose.
MIPS-CG, for example, utilizes an algorithm that does not rely on the use of backbone genes as BGC-anchor points as some BGCs were shown to lack these (e.g., kojic acid BGC). Instead, to identify these non-canonical BGCs, MIPS-CG performs pairwise genome comparisons between different fungal species and removes syntenic blocks, because BGCs are often not located in such regions. Following the assumption that BGCs follow a basic common structure (similar gene functions) MIPS-CG scans for clustered genes that follow this pattern [47].
The EvoMining algorithm uses a phylogenomic approach to detect BGCs. It is based on three evolutionary concepts. First, those new enzymatic functions evolve by retaining the reaction mechanism of the enzyme but changing substrate specificity. Second, the evolution of metabolic pathways can lead to the deployment of existing enzyme families that perform new functions (e.g., fatty acid synthases as the ancestor of PKSs) and third, that BGCs have their origin in central metabolism [71,80].
AntiSMASH version 4 introduced the Clusterfinder algorithm which converts a genome into a string of Pfam domains (i.e., protein family domains). Subsequently, BGCs are annotated where stretches with a significant amount of BGC-related Pfam domains occur [81,82].
The ARTS algorithm (Antibiotic Resistant Target Seeker), although currently supporting only bacterial genomes, is a tool for specifically finding BGCs that encode for an antibiotic. When bacteria produce an antibiotic substance, they can either detoxify and shuttle the compound out of the cell or express a resistant homolog of the gene that is targeted by the antibiotic. The resistance pathway is exploited by ARTS. The algorithm compares the genome to a reference genome and looks for duplicate housekeeping genes in the genome. These resistant homologs often are inside the same BGC. Furthermore, the cluster is often acquired by horizontal gene transfer (HGT). With the combination of BGC prediction via antiSMASH, gene duplicate screening and the comparison of the phylogeny of core gene trees of the organism and the cluster genes, ARTS can predict clusters that are highly likely to be responsible for the biosynthesis of an antibiotic [83,84]. The FRIGG pipeline pursues a similar approach but can be used for fungal genomes [85].
Middas-M follows the hypothesis that BGCs can be found where an array of genes is coregulated. They combine transcriptomic data of several culture conditions with genomic data to predict BGCs [86]. This method will, however, only catch BGCs that actually can be activated under laboratory conditions.
Algorithms use various curated but also non-curated data sources for their calculation of BGC genes. These data repositories can basically hold any kind of information about BGCs (e.g., gene information, genomic arrangement, related species, protein information, educts and products, transcriptional information, phylogenetics, environmental information, etc.). Curated data sources contain fewer data entries, but the reliability and quality of these entries are often pristine, non-redundant, and experimentally backed which leads to more reliable predictions. Non-curated data sources basically hold all the information that is available (e.g., Genbank). This means that the algorithm has to process and filter a lot of data of various quality and actuality (e.g., low-quality sequencing data and annotation, experimentally not verified predictions, etc.). Depending on the popularity of the examined genus this also can lead to many redundancies that might contain flawed information. All in all, non-curated sources contain more, but less reliable data which can reduce the quality and reliability of the prediction.
It should be mentioned that, as the experimental information about a novel BGC type increases, it will subsequently pass over from the rule-independent (high novelty/low confidence) to the rule-based (low novelty/ high confidence) prediction methods. This developmental process of prediction tools will undoubtedly lead to the discovery of yet unknown BGCs in already known and well-described species; however, it is strongly dependent on high-quality and accurate experimental data, provided by laboratories. The elucidation of novel BGC types and their genomic signatures, plus their proper deposit in data repositories (e.g., MiBiG, NP-atlas 2.0, GNPS, etc.) will significantly improve BGC predictions in the future. For more detailed information about genome mining and predictive tools, please consult the review of Blin and coworkers from 2019 [70].
Table 1 lists many of the actual databases and BGC prediction tools. They are ordered according to their last update (Status: March 2022), and also summarizes their characteristics.

3. Activation of Microbial BGCs via Environmental Triggers

In the past decades, many strategies have been developed to shed light on this “dark matter” in fungal genomes and thereby unlock the chemical wealth residing within the fungal kingdom. In the pre-genomics era approaches predominantly sought to induce SM biosynthesis by mimicking environmental triggers. Genome sequencing and bioinformatic analyses for BGC predictions were developed later on and did not only reveal that filamentous fungi possess far more BGCs than anticipated, but also opened opportunities for a more targeted activation of single BGCs.
We are now at a point where the combination of technological possibilities, abundance of available methods, and the already substantial amount of collected data makes it possible to expand the research to novel yet not well described fungal species. There is no “one for all”-approach that is superior to other strategies, but available methods should rather be used in a complementary and context-dependent manner.

3.1. Mimicking Natural Habitats and Stress Conditions

Fungal SMs serve different purposes during the life cycle of a fungal colony and accordingly, their production profiles somehow mirror these environmental or developmental conditions. Therefore, one of the approaches to trigger the production of new SMs is to impose challenging, sometimes stressful growth conditions on the fungal cells. Indeed, testing different cultivation conditions and imposing environmental stress has proven to be a very effective way of triggering the production of otherwise absent secondary metabolites. This approach was termed “One Strain-Many Compounds” (OSMAC) [115,116] and is very useful for assessing which BGCs can be induced within a laboratory environment and which need more laborsome methods such as genetic manipulation for their activation. Especially in the case of less-studied fungi that are in the process of being established for laboratory work, different cultivation conditions also give insights into the fungal physiology and suitable media and parameters for their cultivation.
In some instances, it was shown that the fungus reacted as expected to the applied conditions. For example, UV light exposure can trigger the production of UV protectants [117], while oxidative stress can lead to the production of antioxidants [118]. It was also shown that A. terreus produces the pathogenesis-related SM terrain when fungal culture conditions mimic the environment of the plant or rhizosphere, which is associated with nitrogen starvation, iron shortage, and the presence of methionine [119].
Environmental stimuli are generally processed via global regulators. The purpose of global regulators is to coordinate the expression of an array of genes that ensure a resource-efficient response that is suited to the environmental circumstances [120,121]. Therefore, instead of mimicking such stimuli via media and cultivation conditions, the genetic engineering of global regulators (i.e., deletion, overexpression, knockdown, mutation) is also a successful way of triggering the activation of BGCs. This, however, can also lead to phenotypical abnormalities such as aberrant sexual and asexual cycle and growth defects [122,123] which could be linked to the intertwined character of many global regulatory networks [29].
The most thoroughly researched environmental signals that are connected to the production of fungal metabolites are nutrient availability and light regulation.
The type and concentration of the available nitrogen source also greatly impacts secondary metabolism, as shown for some well-established model organisms from the genus Fusarium [124,125,126,127] and Aspergillus [128]. The major GATA transcription factor AreA, AreB, and Nmr are global regulators of nitrogen-dependent genes and pathways in filamentous fungi. In F. fujikuroi, deletion of AREA and AREB led to differential regulation of many known but also yet cryptic BGCs [127]. In F. graminearum, AreA is required for the production of deoxynivalenol (DON), zearalenone, and fusarielin H [129]. Similar results were obtained in Acremonium chrysogenum where the deletion of areA led to the reduction of cephalosporin [130] and in Fusarium oxysporum where AreA is necessary for the synthesis of beauvericin and ferricrocin [131].
Light is also an important natural regulator of growth, development/conidiation and biosynthesis of natural products in fungi [132,133,134,135,136]. Genetic manipulation of components of the fungal-specific light regulator velvet complex showed its involvement in many developmental and also biosynthetic processes [137,138,139,140,141]. It consists of velvet-like proteins like, VelA, VelB and most importantly, the methyltransferase LaeA. Deletion of laeA in A. nidulans, resulted in the abolishment of the production of sterigmatocystin, penicillin, and lovastatin while overexpression of laeA led to increased lovastatin and penicillin production [142]. Similarly, LaeA was also found to be a positive regulator of secondary metabolism in Valsa mali [143], Aspergillus ochraceus [136], Penicillium expansum [144], Chaetomium globosum [145], Fusarium verticillioides [146] and F. fujikuroi [147]. Therefore, LaeA seems to have a conserved role of a positive regulator of BGCs in many filamentous fungi which makes it a promising target for overexpression studies.
Other variables influencing SM levels are the pH value of the growth medium [148,149,150,151,152,153], an effect mainly mediated by the transcription factor (TF) PacC [154,155,156], temperature stress [118,157,158], osmotic stress, and salinity [159,160,161,162], or metals and trace elements [163,164,165,166]. The concentration and quality of the available carbon source can also be a critical parameter [135,167,168] for SM production which is mediated by the carbon catabolite repressor CreA [122]. For oxidative stress [118,123,169,170,171] Skn7 [172] and ApyapA/AoyapA (Aflatoxin production) are known response regulators [173,174]. There are general repressors such as McrA, but the conditions for its activity remain unknown [175] or Nsd1(CsM1/Ltf1) which is involved in sexual development and conidiation in A. nidulans and other fungi [176].
Apart from media composition, the type of cultivation condition also can play a pivotal role in the physiology of the fungus and thus potentially influence the production of secondary metabolites. Differences in the metabolomic profiles could be observed when comparing submerse and solid cultivation [177,178], as well as between agitated or static fermentation [179]. The addition of a polymeric growth support to a submersed culture also had profound influences on growth and metabolite biosynthesis [180].
Table 2 provides an overview on methods and examples where they have been used to successfully induce BGCs by mimicking natural triggers or by genetic manipulation of global regulators.

3.2. Co-Cultivation

During co-cultivation [116,211,212], microorganisms are brought into an interactive environment with the aim to mimic natural competition that can also trigger the production of SMs. There are several possible setups for co-cultivation. One approach uses (random) pairwise co-cultivation of strains [213] (Table 3; Setup 1). A second approach tests a strain collection against certain microorganisms of interest, such as (antibiotic-resistant) pathogens [214,215] (Table 3; Setup 2). A third approach cultivates strains together with the (sterile-filtrated or extracted) culture supernatants of other microorganisms. The latter approach can be necessary if cultivation or media requirements of the individual species are not compatible for simultaneous cultivation or if the mechanisms of the induction of SMs (e.g., physical interaction or presence of a compound) is being investigated [216] (Table 3; Setup 3). This latter strategy was just recently chosen in our laboratory resulting in the identification of the signaling molecule, that triggers the fungal response in the interaction between the bacterium Streptomyces rapamycinicus and the filamentous ascomycete A. nidulans [184]. Another fruitful concept is the co-cultivation of a fungus with a bacteriophage (i.e., infection) which led to the production of a novel SM [217]. Co-cultivation can be very challenging due to the sheer endless combinatorial numbers of specimens and cultivation parameters that have to fit both organisms. However, if the co-cultivation approach is already well established, it can be a very powerful and rewarding tool. Table 3 lists publications that successfully applied co-cultivation approaches for the activation of BGCs.

3.3. Conclusions of Initial Assessment

OSMAC approaches, paired with metabolome analysis, transcriptome analysis, and bioinformatic methods can already create an extensive map of the secondary metabolome that is produced in a laboratory environment. To discriminate already known compounds from putatively new compounds, exact masses of the new compounds are compared to masses from already described compounds. This can be performed in the lab by screening against a library of SM standards [234] or by applying bioinformatic tools such as PRISM 4 [235] or DEREPLICATOR+ [76].
After this initial assessment, the remaining cryptic BGCs can then be targeted by molecular biology methods for releasing the full bioactive potential of the target organism.

4. BGC Activation Using Molecular Biology Tools: The Way to Fully Exploit the Genome for Novel Bioactive Substances

Applying molecular biology tools for genetic engineering enables the alteration of regulatory mechanisms at the genomic or epigenetic layer. This way, cryptic BGCs not transcribed under any tested laboratory condition can be activated in the native host by the manipulation of global regulators, pathway-specific regulators, chromatin modifiers, or by disruption of a repressive signaling pathway. Furthermore, if the fungus is not amenable to any approach that requires genetic transformation, the whole BGC may be transferred to and expressed in another organism (i.e., heterologous expression). There are untargeted methods that result in a global metabolomic and transcriptional response (e.g., deletion of global repressor or overexpression of global activator) and there are targeted methods, that aim to activate only a specific BGC by manipulating regulatory elements of the BGC (e.g., overexpression of pathway-specific transcription factor (TF)). The advantage of targeted methods is that changes in the metabolome can be much more easily linked to the BGC under study. Figure 2 illustrates the molecular genetic methods for the activation of silent BGCs.

4.1. Overexpression of a Pathway Specific Transcription Factor or Whole BGC

By overexpression of a pathway-specific TF, the whole cluster may be upregulated. This strategy was shown to be successful in increasing the production of aflatoxin in A. flavus [236] and also led to the discovery of aspyridone [237], and asperfuranone in A. nidulans [238] or trichosetin in F. fujikuroi [239]. Further examples can be retrieved from Table 4.
Bioinformatic analyses revealed that in the case of the genus Aspergillus 50% to up to 70% of the BGCs contain a pathway-specific TF [240,241]. Co-expression studies under different conditions in Aspergillus niger, however, provided evidence that eight of 45 BGCs that contain a TF, show no co-expression between the BGC-synthetic genes and the TF, which indicates that they are not affiliated with the BGC and hence, overexpression would not lead to its activation [242]. This hypothesis is supported by a publication where overexpression of the pathway-specific TF failed in 15 of 18 cases in A. nidulans [243]. This approach also fails if the TF needs to get post-translationally activated, as it is the case with AflR, the pathway-specific TF of the sterigmatocystin cluster of A. nidulans, which is inactivated by the presence of glucose [244] and activated by the RimO kinase in response to glucose starvation, as recently discovered in our laboratory [245]. Another possibility could be a necessary interaction between the BGC-specific TF and another factor or natural inducer to unfold the full activation potential. In this scenario, the overexpression of just the TF might not have any effect on the clustered genes [246]. A possible way to circumvent this was shown in the (+)-asperlin cluster in A. nidulans. The modular composition of many TFs (activator domain, DNA-binding domain) made it possible to fuse the DNA-binding domain of the asperlin BGC-specific TF (AlnR) with the trans-activator domain of another TF (afoA). This way, the overexpression of this hybrid TF led to the discovery of the antibiotic (+)-asperlin [246].
In the case of clusters that do not have a TF, serial promoter replacements of all BGC-specific genes can be conducted. This approach has been successful in the activation of several uncharacterized BGCs in A. nidulans, finally shown to be responsible for the expression of the Ivo gene cluster producing a new pigment [247], or for the discovery of the new proteasome inhibitor fellutamide B [248], and chicorine [243]. The idea was also tested in other species and Table 4 summarizes some of the successful attempts in the published literature. However, it is clear that exchanging many promoters in a multi-gene BGC can be quite cumbersome and may not always lead to successful transformation events, especially when the BGC under investigation is located in heterochromatic regions.

4.2. Heterologous Expression

In the case that none of the aforementioned methods worked out or the studied species is not amenable to transformation or does not yield stable transformants, heterologous expression in a model host can be a powerful tool for the activation of unknown BGCs. For this method, single protein domains or genes or a whole BGC are introduced into a well-known production host. For the expression of single domains or genes, bacterial hosts such as Escherichia coli are often sufficient [272,273]. For larger genes or small clusters, many (engineered) yeast overexpression hosts such as Saccharomyces cerevisiae, Hansenula polymorpha or Pichia pastoris exist [274,275].
If the expression scaffold of the donor organism (promoters, terminators, codon usage, splicing, etc.) is however not suited for the host there are two other possibilities. The regulatory elements of a BGC can be exchanged by host-specific ones, or the BGC can be heterologously expressed in a host that is closely related to the donor organism [274]. There are already engineered fungal hosts (low metabolite background, removal of proteases) that are readily deployed such as A. oryzae [276], A. nidulans [277] A. niger [278], but also F. grameniarum [279,280] and Penicillium strains [281,282] are emerging as expression platforms. Heterologous expression is often used when high yields of a certain SM are important (i.e., industrial application) but specifically for the activation of silent fungal BGCs, systems such as FAC-MS [283,284], HeX [285] or CoIN [286] have been developed. Table 5 lists publications that used heterologous expression for the activation of silent BGCs.

4.2.1. Fungal Artificial Chromosomes (FAC-MS) [283,284]

During the FAC-MS approach, the donors’ DNA is sheared and random chromosomal parts (up to 300 kb) which eventually contain a whole BGC are cloned into fungal artificial chromosomes (FACs). These FACs are then transformed into the host organism (e.g., A nidulans). This already led to the identification of 15 new metabolites from a library of 56 FACs each containing one unknown BGC from A. aculeatus, A. terreus or A. wentii in the heterologous expression host A. nidulans. It also resulted in the elucidation of the biosynthetic pathway of acu-dioxomorpholine from A. aculeatus [283,284,287]. As long as the expression host can utilize the AMA1-region and the donors’ genomic information, the FAC-MS approach is a viable option for BGC mining.

4.2.2. CoIN [286]

This approach uses the A. nidulans regulatory matrix (TFs and their downstream promoters) of the sterigmatocystin cluster in combination with the niiA/niAD promoter. In this system, the TF-encoding genes aflR and its cofactor aflS are put under the control of the bi-directional promoter of the niiA/niaD genes coding for nitrate- and nitrite reductase, respectively, in A. nidulans. The genes of the BGC of interest are put under the control of the eight sterigmatocystin promoters that are regulated by AflR/AflS. This way, the orchestrated expression of up to eight genes of an unknown BGC without reusing a single promoter sequence can be accomplished by the growth of nitrate as the sole N-source leading to a high induction of the TFs.

4.2.3. Single Promoter BGC Expression [296]

Hoefgen and coworkers utilized the 2A peptide of the picornavirus to enable the transcription and translation of a whole BGC with only one promoter and one terminator. The 2A peptide auto-cleaves itself after translation without leading to the disassembly of ribosomes. This way, several genes can be co-expressed with the use of only a single promoter. The cloning of the construct however can be difficult and laborious and the transcript levels of the genes will stay equal and cannot be individually regulated.

4.2.4. Hex [285]

With the Hex approach, it is possible to express whole fungal BGCs in an engineered strain of Saccharomyces cerevisiae. New promoter sequences that are coregulated with pADH2 (called Hex-promoters) have been utilized to enable cloning via a high throughput DNA assembly strategy by homologous recombination into 2 µ vectors. Genome mining assisted the selection of novel BCGs and subsequent cloning into yeast plasmids with an upper limit of seven genes per plasmid (maximum of two plasmids per clone). Expression is followed by untargeted metabolomics and LC-MS and NMR for structural elucidation of putatively novel compounds. For proof of concept, 41 investigated BGCs (from ascomycete and basidiomycete species) resulted in the detection of 22 new compounds.

5. Chromatin: An Important Layer of (BGC) Gene Regulation

While previously mentioned methods have already led to the discovery of many bioactive substances of fungal origin, many BGCs may still resist discovery with these approaches. Therefore, another strategy would be the manipulation of epigenetic and chromatin-based regulation of BGCs. Until recently, this could only be accomplished by the manipulation of global chromatin remodelers, however, as the general knowledge of this level of regulation increases, engineering of programmable chromatin regulators becomes feasible. The emergence of CRISPR/Cas-based methods, moreover, opens up new opportunities for activating BGCs by locus-specific modification of histones. These new approaches would not only make it possible to activate BGCs where other methods failed but have the potential of giving many new insights into the histone code of gene regulation. The next sections concentrate on the regulation of BGC expression via chromatin and histone modifications and describe also the current state of locus-specific chromatin modifications for the activation of secondary metabolites. After that, we will review the role of a CRISPR/Cas as a carrier for targeted BGC activation via locus-specific chromatin modification.
In eukaryotes, genomic DNA wraps around histone protein octamers (~146 bp DNA/nucleosome) to form nucleosome chains (i.e., chromatin) [298,299]. Each nucleosome consists of an octamer constituted of two of the four core histone proteins H2A, H2B, H3 and H4 with sometimes iso-forms (e.g., H2AZ, CENH3, H4.2, H4v) of the histones being incorporated into the nucleosome for the specific cellular processes (e.g., DNA repair, etc.) [300]. Histones, especially H3 and H4, can carry post-translational modifications (or histone marks) on specific amino acid residues, that are read and translated by so-called reader proteins. Depending on the histone marks, chromatin structure is either loosely packed and open for transcription or very compact which hinders access to the DNA and therefore transcription. Furthermore, chromatin reader proteins can also recruit other proteins such as TFs or components of the transcriptional machinery. This way, genes can be either packaged into a very tight polymer of nucleosomes that is transcriptionally silent, called heterochromatin, or into a more relaxed form of packaging, called euchromatin, that allows proteins to interact with the underlying DNA (i.e., TFs, transcriptional machinery, etc.) [301].

5.1. Heterochromatin

The DNA within heterochromatic regions is not freely accessible for transcriptional activation. Pioneering factors however are capable of binding to condensed chromatin and subsequently recruit TFs or chromatin remodelers. This way, heterochromatin can be reversed to a more open and accessible state which promotes the binding and interaction of TFs or other reader/writer/eraser proteins and subsequently can result in gene activation [302,303]. In general, there are two pathways to form heterochromatic regions within the genome which are conserved in most eukaryotes i.e., constitutive and facultative heterochromatin. Both chromatin states are associated with distinct histone methylation marks.

5.1.1. Constitutive Heterochromatin via H3K9 Methylation and Clr4/HP1

Methylation of H3K9 is a hallmark of constitutive heterochromatin. H3K9me3 is established by the histone methyl-transferase KMT1 (ClrD in A. nidulans, Schizosaccharomyces pombe Clr4 homolog) [304] and the heterochromatin protein 1 (HP1, A. nidulans homolog HepA) which is a reader protein for this mark [305]. In some fungi, it was shown that the DNA methyltransferase (e.g., Dim-2 in Neurospora crassa) is subsequently recruited for further silencing of the region [306]. In the case of fungi, the term “constitutive” may be misleading as it was shown that A. nidulans can activate its sterigmatocystin cluster that contains hallmarks of constitutive heterochromatin (i.e., H3K9me3, HepA occupancy). Our lab showed that under inducing conditions in the wild type (nutrient starvation), H3K9me3 peaks are depleted from the ST BGC, and dissociation of HepA and increase of acetylation levels in the BGC occur, and these events are correlated with transcriptional activation of the cluster [305]. A similar observation was made in Epichloe festucae where BGCs that are silent in axenic culture, had H3K9me3 marks, but during plant infection, the H3K9me3 marks were reduced and the BGCs were activated [307]. In contrast, however, the phytopathogenic fungus F. fujikuroi does not have any detectable H3K9me3-enriched areas inside BGCs. In F. fujikuroi, these marks are generally in pericentric and centromeric regions and sometimes in transposon-rich regions with few or no annotated genes [125]. In Fusarium mangiferae, deletion of the clrD homolog FmKMT1 led to increased production of beauvericin biosynthesis while fusapyrone biosynthesis was nearly abolished [308]. These examples suggest that heterochromatin via H3K9 methylation is differently involved in BGC silencing depending on the fungal species and the BGC in question. This pathway may be especially important in fungal species such as aspergilli, which lacks the molecular machinery for the establishment of H3K27me3 (PRC2 complex) associated with facultative heterochromatin in other fungi [309,310,311]. Thus, it is currently unknown if and how facultative heterochromatin is established in these genera that do not possess PRC2 homologs.

5.1.2. Facultative Heterochromatin via H3K27me3 and Polycomb Repressive Complex II (PRC2)

The second pathway for heterochromatin formation is mediated by the polycomb repressive complex II and marked by the histone modification H3K27me3 [307,312,313]. This form of heterochromatin is generally regarded as facultative as it can be conditionally restructured into euchromatin, hence promoting the transcription of underlying genes (e.g., regulation of developmental genes in drosophila) [314]. Some genera such as Aspergillus and Penicillium [309], Ustilago [310], and some yeasts [311] however are devoid of these chromatin marks. In fungi such as F. graminearum [312] and F. fujikuroi [315] many BGCs are located in H3K27me3 regions and knockdown or knockout of the methyltransferase-encoding gene KMT6 of the PRC2 led to de-repression of several BGCs. More interestingly, formerly H3K27me3 regions are acetylated in KMT6 knock-down strains as shown for the beauvericin BGC in F. fujikuroi [315]. Similar effects were observed for Verticillium dahliae where lack of Set7, the PRC2 methyl transferase, led to the induction of 839 genes of which 75% are within H3K27me3 domains in the wild type. Furthermore, many genes that are upregulated in planta are also located within these domains [316]. In the case of N. crassa, H3K27me3 marks are distributed throughout the genome but especially in the sub-telomeric regions where it was shown that a short DNA telomere repeat (i.e., (TTAGGG)17) can trigger the establishment of large domains of H3K27me2/3 [317]. In N. crassa facultative heterochromatin covers around 7% of the genome containing 1000 silent genes [313]. In Magnaporthe oryzae, 50% of TEs and many plant infection-related genes, which are suppressed in axenic culture, are covered by H3K27me3 [318].
These examples show the versatile nature of gene regulation by H3K27me3 and H3K9me3 in different fungal genera which still needs more research that elucidates the mechanism and chronological order of the (de)-repression process.
Accumulating evidence in mammals [319] and also in fungi (e.g., N. crassa [317]) suggest that there exist DNA motifs that trigger specific histone modifications. In humans and mouse, over 350 of such motifs have been found and linked to different histone modifications such as H3K4me1, H3K4me2, H3K4me3, H3K27ac, H3K27me3, H3K9me3, H3K9ac, and H3K36me3. Genetic manipulations of these motifs were successful in reducing H3K27 acetylation at a chosen locus [319]. If this is also applicable to filamentous fungi, especially H3K27me3 and H3K9me3 it would be a promising approach for relieving heterochromatic structures around silent BGCs.
Table 6 lists various approaches that lead to the production of SMs by interfering with chromatin active enzymes (via chemicals or via molecular biology methods such as overexpression, deletion, or knockdown).

5.2. Euchromatin

While heterochromatic regions suppress the transcription of underlying genes, euchromatin promotes gene expression. It is characterized by highly accessible DNA (i.e., low nucleosome density) and is often marked by histone acetylation and H3K4 trimethylation. Histone acetylation is regarded as an active mark and is established by histone acetyl transferases (HATs) [347]. Usually, they are part of protein complexes such as the Saga/Ada (HAT: GcnE) or the NuA4 complex (HAT: EsaA) [348]. Removal of acetylation marks is performed by histone deacetylases (HDACs) which often mediate transcriptional repression [324]. Some of these HDACs were shown to act on the genome-wide level and are involved in developmental and metabolic processes (e.g., HosA in A. flavus [349], HosB and HstA in A. nidulans [350]) while others are active in specific regions (i.e., Sir2 in Candida albicans in subtelomeric regions [351], HdaA regulates two telomer proximal but not distal clusters in A. nidulans [350]). These findings suggest that some HDACs have a location or even cluster-specific activity. Demethylation by HDACs can be inhibited by HDAC inhibitors. Because these chemicals often inactivate multiple HDACs simultaneously, their application results in a genome-wide increase in acetylation and de-regulation of thousands of genes [352]. For references concerning the inhibition or deletion of HDACs for activation of silent BGCs please consult Table 6.

5.2.1. Methylation of Histone 3 Lysin 4

H3K4me1/2/3 is probably one of the most discussed and controversial histone modifications in terms of gene transcription and activation. It is established by the COMPASS complex with Set1 as the catalytic subunit [353]. H3K4me3 is reversed by lysine demethylases such as KdmB in A. nidulans [354,355] or Kdm5 in F. graminearum [355]. This modification was initially thought to have a transcriptionally promoting character. Recent studies in mice, however, suggest that H3K4me3 is not necessary for transcription but instead prevents silencing of the region by PRC2 or DNA methylation [356]. This would explain several contradictory observations such as the underrepresentation of H3K4me3 in active clusters in several fungal species such as A. nidulans [354], F. fujikuroi [125] and F. graminearum [312] or the partial upregulation of BGCs after abolished H3K4me3 due to deletion of subunits of the COMPASS complex (e.g., cclA, sppA, setA). This was observed in many species such as A. nidulans (monodictyphenone, desacetylaustin, dehydroaustinol, citreorosein, emodic acid, 2-OH, 2-aminoemodin, emodin, and the polyketides F9775A and F9775B) [357,358], A. fumigatus (gliotoxin) [359], F. fujikuroi (BIK, FUS) [360], F. graminearum (ZON, fusarins) [360]. But also downregulation of BGCs in A. nidulans (Penicillin G, Sterigmatocystin) [355], F. fujikuroi (gibberellic acid) [360] and F. graminearum (deoxynivalenol) [360] was observed in these mutants. Further puzzling results came from knockout experiments of the H3K4 demethylase (i.e., kdmB) in A. nidulans by our laboratory. It led to a 20% increase of global H3K4me3 but 90% of all annotated genes that are connected to secondary metabolism were downregulated. Sterigmatocystin and orsellinic acid biosynthesis were reduced while products of the monodictyphenone pathway were increased (i.e., 2,ω-dihydroxyemodin, ω-hydroxyemodin, 2-hydroxyemodin, emodin) [354]. Following this observation, the role of KdmB as inducer of SM clusters goes beyond its function as H3K4m3 demethylase and several indirect effects on general SM regulators may account for the observed phenotype of kdmB deletion or overexpression. For example, one could assume that, due to the downregulation of the majority of secondary metabolism genes, the precursor pool may have been increased which may have led to a better turnover by the same amounts of enzymes from the mdp pathway, resulting in higher quantities of mpd-related metabolites [354]. Similar results were found in F. fujukuroi [337] and F. graminearum [355] but most interestingly, in the case of Kdm5 in F. graminearum, we found that many SMs are regulated by Kdm5 in a demethylase-independent role (regulation by a catalytically inactive variant of Kdm5) [355].
Taken together, results that try to link H3K4me3 and active transcription often result in complex and sometimes contradictory results. The role of H3K4me3 as protection against silencing would explain why this mark often but not always is correlated with transcriptional activation. Further insight into the local role of H3K4 trimethylation would be important to evaluate its role in the induction and transcription of silent BGCs.

5.2.2. H4K12 Acetylation

In A. nidulans, the H4K12 acetyltransferase EsaA is essential for viability and also involved in the up-regulation of three of four examined BGCs. EsaA overexpression led to depletion of the H4 nucleosome at investigated loci and increased expression levels of BGCs. In the case of the cryptic ORS cluster, increased acetylation and nucleosome depletion occurred, but the BGC was not activated. It was suggested by the authors that EsaA and H4K12 acetylation could be responsible for histone depletion at promoter regions so that other effectors such as TFs can access the locus and recruit the transcriptional machinery [361].
Similar findings were published by Roze et al. in A. parasiticus where they detected an increase of overall H4 acetylation in the aflatoxin BGC promoters followed by aflR binding and transcription [362].

5.2.3. H3K9 Acetylation/H3K14 Acetylation

H3K9 acetylation, which is the mark counteracting silencing by H3K9me3, is enriched in transcriptionally active promoters, sometimes enriched especially in BGCs. In F. fujikuroi it was seen that the overall correlation between genome-wide H3K9ac and transcription is weak but that in the case of BGCs the correlation is very strong, suggesting that H3K9ac has important roles in the production of SMs [125]. In the case of the fusarin C BGC, H3K9 acetylation was significantly enriched across the whole BGC when grown under inducing conditions [363]. Another study showed that deletion of the histone deacetylase FfHda1 in F. fujikuroi also results in increased levels of H3K9ac in gene promoters and increased transcriptional activity. This even resulted in the activation of the bikaverin cluster under repressive conditions [364].
The most intensively studied HAT GcnE, which is part of the SAGA/Ada complex, was shown to be responsible for the acetylation of H3K9 and H3K14 in A. nidulans, and additionally H3K18 and H3K27 in F. graminearum [365] and H3K4, H3K9, H3K18 and H3K27 in F. fujikuroi [366]. In the case of several BGCs in A. nidulans (orsellinic acid, sterigmatocystin, penicillin, terrequinone) GcnE is necessary for activation. While H3K14 acetylation was not enriched within clusters, H3K9 acetylation was only applied in BGCs (only sterigmatocystin and ORS clusters were examined for H3K9ac). While genes outside the cluster showed no signs of differential expression, all BGCs that were expressed in the wild type showed abolished transcription in the gcnE deletion mutant [367,368]. The acetylation pattern around active genes showed a strong correlation between the increase of H3K9ac around the TSS and transcription while in the case of H3K14ac, the correlation was rather weak. Upon acetylation, the H3 density around the TSS was reduced which suggests that at least one of these marks could be involved in nucleosome depletion around the TSS [367].
Recently the fungal-specific HAT Rtt109, which is responsible for H3K9 and/ or H3K56 acetylation (DNA damage repair), was described in A. flavus, Beauveria bassiana and F. graminearum [369]. Deletion in A. flavus led to the removal of both H3K9 and H3K56 acetylation which resulted in impaired growth and virulence, vulnerability to DNA damage stress and also to the abolished expression of genes of the aflatoxin cluster, further promoting H3K9 acetylation as an important mark for gene transcription [370].

5.2.4. H3K27 Acetylation

H3K27 acetylation is the counteracting mark of the silencing H3K27me3 modification. During chromatin accessibility studies in Neurospora crassa, it was found that H3K27 acetylation is one of the most predictive histone marks of open chromatin [371]. As already mentioned it was shown in F. fujikuroi [335] and M. oryzae [318] that H3K27ac spreads to regions previously marked by H3K27me3 if components of the PRC2 complex are deleted. This often resulted in transcriptional activation. Loss of H3K27ac on the other hand did not result in transcriptional inactivation. This suggests that H3K27ac is not necessary for active transcription but plays a role in gene activation/accessibility of previously silenced genes. It seems that H3K27 acetylation by the SAGA complex (i.e., GcnE) is one of the mechanisms that can revert silencing by the PRC2 machinery. The exchange mechanism itself and the involvement of other histone acetylations (by GcnE) in the transcriptional activation is however not fully understood yet.

5.3. The Histone Code of Actively Transcribed Genes

Actively transcribed genes have a distinct chromatin signature around the promoter region, gene body, and terminator. We assessed the histone mark occupation around the start codon in all active genes of Chromosome IV of A. nidulans. H3K9/K14 acetylation is enriched in the promoter with the highest level at the first nucleosome after the translation start site. H3K4me3 is enriched at the first three nucleosomes downstream of the ATG and H3K36me3 spreads across the active gene body. While H3K4me3 and H3K9/K14Ac decreased towards the 3′ of the gene, H3K36me3 increased [354]. Similar findings were published in V. dahliae [316] and Candida albicans [351]. Furthermore, it was found that transcriptional active genes correlate positively with increased acetylation within the promoter region in Aspergillus parasiticus [362]. Noteworthy, in human cells, artificial application of H3K27 acetylation in the promoter region led to subsequent H3K4me3 and transcription while the deployment of H3K4me3 did not result in increased acetylation or transcription [372].

5.4. Finding New Bioactive Compounds by Breaking the Chromatin Silencing Machinery

The production of SMs is a metabolically costly process requiring high energy input to produce the necessary reduction equivalents (NADPH) and ATP. Therefore, their biosynthesis is strictly economized. To ensure that energy is not wasted, fungi by default silence their BGCs at the level of heterochromatin formation, making these genomic regions inaccessible to the transcriptional machinery [305,315,348]. As soon as environmental or developmental conditions demand the production of a certain beneficial SM, the chromatin-based silencing is lifted and specific activation factors start the expression of BGC genes. Experimental setups that aim for induction of fungal BGCs via chromatin modifications either delete, inhibit or overexpress chromatin active enzymes. Most common approaches in this regard are the genetic of chemical inhibition of HDACs [324,331,335,344] or inactivation of components of the silencing machinery via PRC2 [307,312,334,338] or components related to heterochromatin formation via H3K9me (e.g., Dim5/Kmt1/ClrD/HP1, …) [305,307,308,334,339].
One particular successful and easy-to-apply strategy for artificial lifting heterochromatin based silencing was the application of chemical inhibitors for HDACs that are known to be essential for most heterochromatin formation complexes. HDAC inhibitors (HDACi) lead to higher acetylation of histones, a driver for gene expression. The most rewarding results following this strategy used SAHA, butyric acid, valproic acid, nicotinamide, and trichostatin as additives to the growth medium [188,342]. A similar approach that often is applied together with HDACi aims for the inhibition of DNA-methyltransferases (DNMTs). DNA methylation in fungi is mostly connected to the silencing of transposable and repetitive elements [373]. Inhibition or deletion of HDACs or DNMTs resulted in the discovery of nygerone A in A. niger [119], or the production of diorcinol 3-O-α-D-ribofuranoside in a marine-derived isolate from Aspergillus sp. [130] or in the induction of many silent BGCs during a screen of 162 mangrove endophytic fungi of which 72 produced bioactive substances [137]. Table 6 lists publications that used the genetic or chemical interference of chromatin remodelers for the induction of silent BGCs.
There are however downsides to strategies that interfere with globally active regulators. While approaches that mimic environmental conditions result in a native regulatory response of BGCs, the disruption of global regulatory mechanisms can lead to the synthesis of cryptic compounds that emerge due to uncoordinated regulation of genes or by bio-transformation of the inhibitors [332,344,374]. Additionally, chemical or genetic manipulation of globally active chromatin remodelers can be lethal [315] or lead to phenotypically aberrant transformants (i.e., growth, morphology, conidiation, …) [308,339,375,376,377]. All these factors can result in a substantial change of the whole metabolome [321,346] and transcriptome [326,351] which makes interpretation challenging, even with software solutions that assist with the detection of degradation products or bio-transformed moieties such as BioTransformer [378].
In recent years, however, CRISPR/Cas technology advanced to a point where new options for targeted local chromatin modifications are available and some have already been established in fungal species.

6. CRISPR/Cas: New Opportunities in Molecular Biology: Targeted Gene Deletion, Activation and Epigenetic Editing in Filamentous Fungi

CRISPR/Cas is a fast-advancing technology in molecular biology that is nowadays deployed in many prokaryotes and eukaryotes, illustrated by around 25,000 publications connected to “CRISPR” over the last 5 years. Its fast development in terms of genome editing but also in epigenome editing and as a tool for transcriptional activation makes it a very promising option for future application. Although CRISPR/Cas technology is widely used for gene manipulation, approaches for gene and BGC activation are just recently gaining attention in filamentous fungi. In the next section, we shortly introduce the principle of CRISPR/CAS technology and different classes and types that are available before moving to approaches that are applied in filamentous fungi and future perspectives of gene activation (i.e., CRISPRa) applications.

6.1. Principle of CRISPR-Cas Technology

The best-known variant of CRIPSR/Cas systems is the RNA-guided nuclease Cas9 from Streptococcus pyrogenes, SpCas9. CRISPR/Cas is the major component of the acquired bacterial or archaeal immune system [379]. It is a nuclease that inactivates foreign DNA by causing double-strand breaks at specific DNA-loci via previously acquired guide sequences.
The native lifecycle of the CRISPR/Cas adaptive immunity consists of three stages. Adaption, expression, and interference [380]. Upon initial infection, invading (e.g., phage) DNA is digested by Cas proteins that recognize foreign DNA by a nucleotide sequence that does not occur in the host genome. The so-called protospacer adjacent motif or PAM. Each CRISPR/Cas system has a distinct PAM sequence. A short stretch of the DNA (i.e., protospacer) is then cleaved out and inserted into a DNA locus called CRISPR-array. In the expression phase, the whole CRISPR array with all protospacers is transcribed (i.e., pre-CRISPR RNA or pre-crRNA) and processed into mature CRISPR RNAs (i.e., crRNA). Different CRISPR/Cas variants have different procedures during this phase, either demanding a complex interplay of many Cas proteins or requiring only one multi-domain protein [380]. During the interference stage, the mature crRNA or guide RNA is used as a reference pattern for the Cas nucleases to inactivate invading non-host DNAs by inducing double-strand breaks at loci that match the crRNA sequence and possess a PAM motif [380].
After the discovery of CRISPR/Cas, its potential instantly attracted the interest of the scientific community. Many research groups refined this strategy over the past years to one of the most versatile tools used in modern molecular biology that has now been established in many eukaryotic organisms, including filamentous fungi [381,382]. New variants of the CRISPR/Cas from different bacterial sources or engineered forms evolve(d). Some are only capable of nicking the DNA or even are incapable of modifying the DNA at all (nuclease- dead Cas9, dCas9). By fusing these variants with proteins of versatile functionality, researchers created a multitude of effector proteins that can be targeted precisely throughout the genome. This opened many new possibilities for addressing research questions that go beyond genome editing, such as CRISPR/Cas mediated chromatin modification [252], gene activation (CRISPRa) [251,253,271,383] or transcriptional interference (CRISPRi) [384,385] and locus-specific protein proximity labelling (CasID or BioID) [386].

6.2. Differences between CRISPR/Cas Variants

Following Cas9, different variants of CRISPR-Cas systems have been found in nature as reviewed in detail by [380]. They are divided into two classes which differ in their complexity. While class I CRISPR/Cas systems utilize many proteins for each phase of the natural defense (adaption, expression, interference), class II CRISPR/Cas systems only use 2–3 proteins during adaption and only 1 (Type V (i.e., Cas12a) and VI) or two (Type II (i.e., Cas9)) proteins during expression and interference. The simplicity of Class II systems makes them the prevalent choice for CRISPR/Cas applications in molecular biology. Depending on the type of Cas variant, they can either target RNA molecules (Type VI) or DNA (Type II and V) molecules [381]. Within Class II especially SpCas9 from Streptococcus pyrogenes or Cas12a (i.e., Cpf1) from Francisella novicida, Acidaminococcus sp. or Lachnospiraceae bacterium (i.e., FnCas12a, AsCas12a, LbCas12a) are being deployed for molecular biology in fungi. They have important differences that make both of them viable options, depending on the application [381].
Easier handling and cloning with Cas12a due to different gRNA processing. Cas9 needs the matured form of the tracrRNA and the crRNA which, in nature, anneal to produce a functional guide RNA that can then build a complex with the Cas9 protein. Genetic engineering-led, however, to the development of both crRNA and tracrRNA in one single guide RNA (sgRNA) which increased the handling comfort and efficiency [387]. Nevertheless, for the proper maturation of the sgRNA, it has to be ensured that only the sgRNA without any additional nucleotides is released from the transcript which is currently accomplished by either ribozymes [388] or tRNAs [389] that flank the sgRNA. The sgRNA maturation via ribozymes uses an RNA polymerase II promoter and terminator together with the Hammerhead and HDV ribozymes to release the mature sgRNA [388]. Nødvig and co-workers took this idea even further and developed a system capable of multiplexing sgRNAs utilizing the endogenous tRNA processing machinery. It consists of a cassette that uses an RNA polymerase III promoter to express sgRNAs which are flanked by tRNAs. This then results in the release of mature sgRNAs by tRNA-specific RNAse P and RNAse Z enzymes [389] (Figure 3a). This made it possible to express multiple sgRNAs at the same time thereby increasing editing efficiency by application of several gRNAs at the same locus which in turn allows for multiple gene editing. However, the genetic scaffold that is required for proper processing and release of multiple mature sgRNAs makes the cloning and transformation process of Cas9 multiplexing systems a challenging and labor-intensive endeavour in filamentous fungi. In the case of the ribozyme approach, the repetitive expression scaffold for one gRNA sequence is around 185 nt-long [388] and in the case of the tRNA-assisted approach, it spans even around 222 nts (containing 2x identical glycine tRNA) [389] (Figure 3a).
In stark contrast to Cas9, Cas12a can directly process the U6 promoter-driven CRISPR array into the mature guide RNAs which only consists of a 19 nt-long direct repeat (DR) and a 23–25 nt-long protospacer sequence (PS) (i.e., U6-DR-PS-DR-PS-DR-PS-, etc.) (Figure 3b). This greatly facilitates the cloning and application of single and multiple sgRNAs with Cas12a [390,391,392,393].
The DNA cleavage pattern is different between Cas9 and Cas12a. While SpCas9 produces a blunt end cut 3 bps upstream of the NGG PAM [387], Cas12a variant FnCas12a cleaves after the 23rd base on the targeted strand at the 18th base of the non-targeted strand which results in staggering cleavage sites distant from the PAM [394]. In the case of Cas9, the cleavage is mediated by two nuclease domains (HNH and RuvC) [395], while Cas12a only has one RuvC domain for the cleavage of both strands [394].
The PAM sequence of Cas9 is the short 3 nucleotide long G-rich motif NGG while Cas12a has a T-rich PAM (AsCas12a and LbCas12a: 5′-TTTN-3′ [381] or 5′-TTTV-3′ [396]; FnCas12a: 5′-TTN-3′ [394]. Other PAMs are possible, depending on the origin of Cas12a [397]. There are also engineered variants that accept a broader range of PAMs which will be described in Section 6.5. The PAM greatly limits the loci that can be targeted because, without a present PAM, neither annealing nor cutting occurs properly.
Genetic engineering led to the development of catalytically inactive (i.e., dead or dCas) and nicking CRISPR/Cas tools. Cas9 proteins that can only nick the DNA (i.e., nCas9) [398] resulted in improved single base editing efficacy [399]. There are also mutant versions of Cas12a proteins although they are not yet capable of perfect nicks as they still retain some of the catalytic activity for the second strand. Merely the ratio of single and double-strand breaks is shifted in favor of the nicks [400,401]. In contrast to Cas9, Cas12a only has one catalytic domain that cleaves both strands which makes the engineering of nCas12a variants more complicated.
The development of catalytically inactive Cas9 and later Cas12a variants (i.e., dead or dCas9 and dCas12a) made CRISPR/Cas to a very powerful tool in molecular biology [400,402]. dCas proteins can be used as guidance vessels for attached proteins or protein domains. This way, CRISPR/Cas can be deployed for a multitude of applications, involving basically any protein that can be functionally fused to the N and/or C-terminal or even as a mediator of transcriptional interference (CRISPRi) by the sterical hindrance of the transcriptional machinery within a gene body or at the transcriptional start site [403].
Although SpCas9 is currently the prevalent choice for gene editing and other applications, the Cas12a system has many benefits for working with filamentous fungi. The biggest advantage with regard to BGC activation is the easy deployment of multiple gRNAs as compared to SpCas9. While Cas12a can process the raw pre-crRNA transcript directly, Cas9 needs more complex cloning strategies involving features that allow for a proper release of blunt end sgRNAs (e.g., ribozymes or tRNA scaffolds). Cas12a is, therefore, much easier to handle which not only makes experiments more time efficient but also more robust as the system requires fewer components.
The second advantage for work with filamentous fungi specifically is the temperature maximum. Most filamentous fungi grow well around room temperature. SpCas9 has its catalytic maximum at 32 °C which can be challenging for some fungi. Cas12a, on the other hand, can be deployed at lower temperatures (i.e., 28 °C) which increases the efficiency in most fungal strains [404].
One of the drawbacks of the CRISPR/Cas system is the requirement of the PAM motif at the target locus. The difference in PAM sequences in different variants increases the versatility and flexibility in this matter which can be very important for experiments that require specific loci, with limited PAM options, to be targeted.

6.3. Gene Editing via CRISPR/Cas

Various approaches that involve CRISPR/Cas technology have been already applied to different fungal species [395]. Especially in fungal species where site-directed mutagenesis and gene editing rates by conventional methods are very low, CRISPR/Cas is a promising option [405,406].
To increase the editing efficiency by utilizing homologous DNA repair pathways, linear or circular double-stranded DNA with homologous regions or single-stranded DNA oligos with small homologous regions can be transformed together with the Cas construct. These then can get inserted at the Cas9 restriction site after cleavage which not only can increase the mutation rate but makes it also possible not only to mutate, but exchange the DNA sequence via homology-directed repair mechanisms (Figure 3c) [389,407].
The transformation of the CRISPR/Cas machinery and the guide RNA(s) can be performed on a single [388] or multiple plasmids [251], episomal plasmids by utilization of the AMA1 region of A. nidulans [388,389], by random integration into the genome [408] or within a specific genomic locus [251]. We recently also developed an inducible CRISPR/Cas variant in filamentous fungi that uses the human estrogen receptor [251,409]. Other variants utilize the Tet-ON system for induction [410].
If the fungus is not amenable to stable or transient genomic transformations, mutation by CRISPR/Cas systems can also be achieved by transformation of the in vitro pre-assembled (i.e., sgRNA and Cas-protein) ribonucleotide complex (RNP) [407,411,412]. For more information about CRISPR/CAS applications in fungi, please consult the recent review by Schuster and Kahmann [413].

6.4. Gene and BGC Activation via CRISPR/dCAS Variants

Catalytically inactive Cas variants (i.e., dCas9, dCas12a) have great potential to be used for the activation of silent BGCs. When fused to a trans-activator domain it can be deployed as a programmable TF for the activation of one or multiple genes (via sgRNA multiplexing; Figure 3a,b) [251,253]. Another potent approach is the fusion of a chromatin remodeler (domain) which would provide the possibility to change histone modifications in a targeted and locus-specific manner [252]. This way, BGCs that are buried within heterochromatin could be made accessible by replacing the silencing marks (i.e., H3K9me3, H3K27me3) with active marks (i.e., acetylation). Figure 4 illustrates both dCas-mediated approaches for the activation of silent BGCs.

6.4.1. Transcriptional Regulation

The fusion of a trans-activator or trans-activating domain and a dCas protein transforms it into a programmable transcription factor that can be sent to any spot within the genome (as long as an appropriate PAM is present). This CRISPRa method was established simultaneously in the filamentous fungi Aspergillus nidulans for dCas9 [251] and dCas12a [253] and shortly after also in P. rubens [271]. In every case, the trans-activator VPR [414] was chosen as it was shown to be more effective than other known trans-activator domains. It is a tripartite activator with protein domains that were shown to interact with parts of the transcriptional machinery and also the histone acetylase complex SAGA [415]. The hybrid trans-activator consists of four copies of the VP16 domain of the Herpes simplex virus, the p65 domain of the human activator NF-κB and the Rta domain of the Epstein-Barr virus TF. These publications tackle multiple challenges when it comes to BGC activation. Roux et al. successfully deployed dCas12a-VPR in an episomal and stable approach. They showed that the activation of a BGC can be achieved by upregulation of the pathway-specific TF but also by using guide RNA multiplexing to upregulate all genes of a BGC [253]. Simultaneously, we published a nucleosome positioning map of the genome of A. nidulans which was used to target VPR-dCas9 into nucleosome-free regions within the promoter [251]. This facilitates the binding of VPR-dCas9 as it was shown that steric hindrances such as nucleosomes or other proteins can prevent the annealing of (d)Cas proteins to their target [416]. Furthermore, we deployed an inducible version of VPR-dCas9 to precisely time the activation process because some metabolites could be toxic to the host and therefore interfere with the morphology and (chemical) phenotype of the fungus if the BGC product is synthesized constitutively [251].

6.4.2. Targeted Chromatin Remodelling Using CRISPR/Cas

Targeted locus-specific deposition of histone marks via fusion proteins of dCas and chromatin active domains is an emerging field in molecular biology. In human cancer cells, targeted modification of H3K9me3 and H3K27ac can be accomplished by deploying a fusion protein composed of dCas9 and the H3K9 methyl-transferase KRAB (i.e., Krüppel-associated box repressor) or the histone acetyl transferase p300, respectively. In both cases, the targeting of the fusion proteins to DNA accessible sites (e.g., nucleosome-free regions, NFRs) led to the respective modification at the locus (i.e., H3K9me3 and H3K27ac). In the case of H3K9me3, it was even shown that the DNA is inaccessible after the histone mark H3K9me3 is deployed [417,418].
In fungi, gene activation and repression by targeting chromatin active domains via CRISPR/Cas to target loci were accomplished in A. niger, although the actual effect on chromatin status (i.e., histone modifications) has not been assessed [252]. Li et al. hereby targeted the human acetyl transferase p300 to the promoter region of the HR-PKS-encoding gene FUM1 and managed to activate the fumonisin cluster in A. niger. Furthermore, they fused endogenous proteins such as GcnE, the acetyltransferase of the SAGA complex, histone deacetylases HosA and RpdA, and the H3K9 demethylase Lde to dCas9 separately and tested its effect on the SM gene breF. While the fusion of the catalytic core of human p300 fused to dCas9 led to an activation of the target gene as expected, fusion proteins of dCas9 and GcnE, RpdA and HosA resulted in repression of the gene while no difference was seen with the dCas9-Lde fusion protein. While most of the results are as expected, the repression effect by dCas9-GcnE is interesting. Although the authors state that the repression of breF by dCas9-GcnE is expected as GcnE was shown to be a negative regulator of 12 PKSs in A. niger, the result is surprising to us. Upregulation of BGCs due to the deletion of a broad domain regulator-encoding gene such as gcnE could be due to many indirect effects, for instance if GcnE was the activator of a repressor of these BGCs. Targeting the acetyltransferase to a specific genomic location, however, should lead to higher levels of acetylation only at the specific location and, therefore, to upregulation of only the targeted gene.
In the case of the dCas9-HosA, activation of the pigmentation gene fwnA was observed. Again, this is theoretically consistent with literature where fwnA expression decreased in hosA deletion strains. Such as with GcnE however, comparing a local effect (i.e., deacetylation of the promoter region) with a genome-wide effect (i.e., deletion of broad domain regulator such as HosA) does not take indirect regulatory effects into account. Unfortunately, histone modifications of the targeted loci were not analyzed, so it is not clear if really a change in histone modifications is responsible for the observations. Furthermore, there could also be the possibility of host-mediated recruitment of the endogenous fusion protein (i.e., HosA, RpdA, GncE, Lde) which could interfere with the experimental setup. Using only the catalytic domain (of homologs) would be a way of circumventing this.
Histone phosphorylation is another mechanism for gene induction that is prevalent in stress response gene promoters. A fusion of dCas9 and a hyperactive variant of the human histone kinase MSK1 (i.e., dMSK1) could be targeted to promoters for gene activation by phosphorylation of H3S28 and H3S10 [419]. Here, several (also non-stress response) genes were activated by targeting the promoter. When gene activation was successful, an increase in H3S10ph, H3S28ph, and H4K27ac was found. In cases where H3S10ph but not H3S28ph occurred, neither H4K27ac nor gene activation was observed. This was, however, not yet tested in fungal organisms.
In summary, the targeted deployment of histone modifications as a regulator of transcription is an emerging field in fungal research but experimental data is currently scarce. More research is needed to fully assess this strategy for the activation of silent BGCs. If successful, CRISPR/Cas-based targeted deployment of chromatin active domains will not only result in the discovery of many natural products but will also provide a lot of information about transcription and the local chromatin context.

6.5. Application Relevant Information for CRISPR/Cas-Methods

Next, we want to summarize insights regarding the deployment of CRISPR/Cas methods. This is aimed at helping with the experimental planning. It will focus on the two main CAS systems that have been used for fungal research, Cas9, and Cas12a.
Cas and dCas approaches have different prerequisites in regard to the target site, residence time and off-targets. When using CRISPR/Cas for gene-editing purposes, the successful outcome depends on a one-time event (the cleavage of DNA). In the case of CRISRRa or CRISPRi however the effect is determined by a continuous interaction of the fusion complex with the proteome or DNA at the target locus [420,421]. Concluding, a sgRNA that proved to elicit DNA mutations via CRISPR/Cas might not be of use for experiments where a long residence time is essential (e.g., CRISPRa).
Off-target annealing also has different consequences depending on the application. Cas proteins can anneal to off-target sites but as long as the sgRNA does not match appropriately, no cut will occur. In the case of dCas approaches, however, this could theoretically lead to off-target effects mediated by the active domain that is fused to the dCas protein.
There are several tools that design gRNAs for different (d)Cas systems based on one or several of the following features: sequence composition, chromatin accessibility, gene expression, nucleotide position, RNA secondary structure, melting temperature, free energy, off-target possibility, and more [422]. Some of them also feature different application types such as DNA cleavage, CRISPRi and CRISPRa. This is important because the sgRNA/crRNA design for mutational Cas approaches is different from those for CRISPRa approaches because they target the gene body and not promoter regions [423]. Most of them are biased towards one or more organisms (mostly mammalian) or applications (e.g., mutational vs. CRISPRa). Some tools such as Cas-Designer (Cas9) [424] feature a broad spectrum of genomes that are continuously updated (also by user request). CASPER (Cas9 and Cas12a) is not biased towards an organism and was specifically designed for genome editing in non-model organisms [425]. Guide RNA design tools for non-model organism often calculate the gRNA efficiency based only on sequence features which can lead to reduced sgRNA performance if other, unconsidered parameters, such as nucleosome occupancy, prevents target site accessibility.
In the case of CRISPRa approaches, the location of deployment within the promoter (in plants) was shown to have a distinctly different effect on the activation of the target gene even if the sgRNAs were only as little as 50 base-pairs apart from each other [426]. For more detailed information, read the following recent reviews which describe available prediction tools and guide RNA optimization [422,427].
Cas efficiency depends on multiple factors including Cas-sgRNA complex formation and concentration, target site accessibility, residence time, and codon optimization. By codon optimization alone, Cas efficiency could be increased from 35 to 70% [428].
Cas stabilizes the sgRNA, but the complex formation can be hindered by other RNAs and some divalent ions. To be functional, the Cas protein and the sgRNA/crRNA have to form a complex. Unbound sgRNAs get degraded very quickly. Upon complex formation, the Cas9 protein stabilizes the sgRNA, undergoes conformational changes, and is released from the inactive apo-state [429,430]. Random RNA molecules that are present in the cell also can bind unspecific to Cas9 and compete with sgRNAs which delays or prevents complex formation [431]. After complex formation however, unspecific RNAs have no significant impact on the ribonucleotide stability anymore [429]. It is therefore important to express the sgRNA/crRNA at high levels so that the Cas protein has a high chance of collision before another RNA molecule can interfere or the sgRNA/crRNA is degraded [429]. The following publications give in-depth information about Cas9 [432] and Cas12a [433] kinetics.
It also is important to note that in the apo-state Cas9 can bind DNA in a non-specific manner while the protein remains catalytically inactive, so no unguided “rogue”-activity is happening under normal conditions [430]. It was, however, shown that the presence of Mn2+ or Co2+ ions can lead to unspecific DNA cleavage even without bound crRNA (Cas12a) or sgRNA (Cas9), respectively, which makes media composition potentially also an important brick in the wall of CRISPR/Cas applications [434]. Processing of the pre-crRNA by Cas12a variants does not seem to be Mg2+-dependent but it was shown that in the case of LbCas12a and FnCas12a, affinity to the crRNA is dependent on Mg2+ ions which were found to be coordinated in the RNP complex [435,436]. It can be concluded that divalent ions can have a significant impact on Cas activity and specificity which should be kept in mind when deciding on culture conditions.
The determinants of the residence time of Cas at its target locus are the proper pairing with the PAM, the subsequent annealing to the seed region, the local proteome and its fluctuations (i.e., accessibility and dislocation), and the concentration of Cas-sgRNA complex [421]. The impact of mutations on the annealing, residence time and cleavage efficiency also follow this order (PAM > seed sequence > PAM-distal sgRNA sequence (=5′-end)) with the proteomic occupancy around the target locus having a varying influence [429,437]. The residence time is significantly shorter the less complementary the annealing sequence is which was shown in human cell lines where residence times of less than two minutes to up to more than three hours have been observed [429].
The proteome of the target site can prevent binding of Cas-sgRNA complex or even dislodge bound Cas. In vitro experiments confirmed that cleavage by Cas9 is nearly abolished when targeted to DNA occupied by nucleosomes. At the edges of nucleosomes or within linker regions, however, cleavage was largely unaffected. Interestingly, the position of the PAM was the essential determinant of cleavage efficiency when sequences close to the border of a nucleosome were targeted. This was verified in vitro by designing two nearly perfectly overlapping sgRNAs (shifted by four nts) but with opposite orientations. The sgRNA that had the PAM located proximal to the nucleosome showed drastically reduced cleavage efficiency compared to the sgRNA that had the PAM located distal [416]. This observation was also confirmed in vivo in human cell lines [438,439,440]. The magnitude of this effect is, however, also dependent on the strength of the interaction between nucleosome and DNA. Upon comparing two different DNA sequences that are known to have different affinities to nucleosomes, it was shown that breathing chromatin structures (fluctuations of DNA occupancy of the nucleosome) show a much higher cleaving permissiveness towards Cas9 than nucleosomes with strong interactions to DNA. Interestingly, the cleaving efficiency around the nucleosome dyad was comparably low in both cases. Furthermore, when repeating the same experiments with the addition of the chromatin remodeler SNF2H or RSC, Cas9 cleavage efficiency was significantly enhanced [441].
The obstruction of CRISPR/Cas by proteins is not limited to nucleosomes. In the case of SpCas9 (Streptococcus pyogenes), which is known to remain bound to both DNA ends after cleavage [442], RNA polymerase was able to dislodge the Cas9 protein when targeted to the template strand. This resulted in increased editing efficiency due to the release of both cleaved DNA ends and the onset of the DNA repair machinery [443]. This might however only be true for SpCas9 because in the case of SaCas9 (Staphylococcus aureus) [444] and AsCas12a (Acidaminococcus sp.) [445] it was shown that these proteins automatically release one DNA strand some hours after the catalytic event.
Apart from the sgRNA annealing site itself, post-PAM interactions were found to be important for stable interactions between the Cas protein and the DNA. SpCas9 was not able to remain in a place of the target site when the DNA sequence 14 base pairs downstream was occupied by another protein. It was observed that this post-PAM interaction is weak in force but very crucial for Cas9 binding and cleavage [446]. Displacement mediated by disturbance of the post-PAM accessibility can also happen after the association of SpCas9 to the DNA. In the case of SaCas9, the post-PAM interaction was only around seven base-pairs and had a much lesser effect than on SpCas9 [446]. SaCas9, in stark contrast to SpCas9, represented a significant barrier for other proteins such as DNA-tracking motors [444]. Zhang and co-workers give a nice comparison of interactions strength between SaCas9 and SpCas9 mediated by different regions across the footprint [444].
Methylated DNA could be cleaved without problems by Cas9 [447] and also CRISPRa approaches had been successful when the target area had DNA-methylated sites [442].
The examples from above show that the proteome of the target site can have a profound impact on the binding and residence time and therefore cleavage efficiency of CRISPR/Cas. Structures that statically occupy and block the DNA such as nucleosomes should be avoided when choosing a target site for CRISPR/Cas applications.
Off-target annealing has different consequences, depending on the application (i.e., Cas vs. dead-Cas). Cleavage at off-target locations depends on PAM sequence and guide-RNA homology. Two major concerns about the deployment of CRISPR/Cas systems are the possibility of off-target effects (i.e., annealing to a not intended stretch of DNA) and the restriction that CRISPR/Cas systems need a specific PAM sequence for annealing. An interesting finding in case of off-target activity of Cas9 was, that the RNP complex could bind to many off-targets, even with a seed sequence complementarity of only five nts (seed: 6–12 nts or 5 nts proximal to the PAM sequence for SpCas9 [448] or FnCas12a [394,435] respectively). However, for proper catalytic function, a complementary gRNA of at least 17 nts was found to be necessary. In the case of Cas9, proper initial binding of the Cas9-sgRNA complex to the DNA is mainly mediated by PAM interaction and complementary forces between the DNA and the seed region of the sgRNA. In a second step, Cas9 cleaves after sufficient pairing of the 5′-end of the guide RNA [449]. This model would explain why binding of Cas9 to the target site can even occur in case of little homology between sgRNA and target site but cleavage is nevertheless impaired due to low overall homology between sgRNA and target sequence.
The PAMS and, therefore, the possible annealing/cutting sites are different for Cas9 and Cas12a and are one of the biggest limitations of these tools. As already mentioned, Cas9 has a G-rich while Cas12a variants have T-rich PAMs. To circumvent these limitations, protein engineering performed on Cas9 and Cas12a led to the development of near- PAMless [450] or PAM-flexible [451,452,453,454] variants with increased off-target activity. But also high fidelity (HF) [455] variants have been engineered. Even combinations of both variants exist which compensate for the increased off-target activity of near-PAM-less or PAM-flexible variants [453,454]. It is of importance that the sgRNA/crRNA design has to account for the flexible PAM. It must not enable the Cas protein to anneal to the gRNA expression cassette itself which could lead to cleavage or transcriptional blockage. To assess the (off-target) cleavage pattern of a specific Cas-sgRNA complex in vivo, the GUIDE-SEQ approach can be applied which performs an unbiased detection of CRISPR/Cas-mediated double-strand breaks (DSBs) by in vivo integration of an oligo at sites that have DSBs (i.e., on- and off-target loci) and subsequent amplification and next-generation sequencing [456].
The information above should illustrate the many application forms and variants of CRISPR/Cas systems. There is not a one for all experimental strategy because the prerequisites of the experimental question determine not only the variant of CRISPR/Cas to be deployed but also the target genetic features (i.e., gene body in case of gene knockouts but promoters in case of activation via CRISPRa approaches) and their connected properties (i.e., PAM, proteomic occupancy, sequence bias) that have to be considered during gRNA design. In fungal organisms, mutational CRISPR approaches are already widely implemented. In the case of CRISPRa or targeted chromatin remodeling, however, this technology is still in its infancy but harbors great potential of the activation of BGCs and the discovery of new bioactive metabolites in the future.

7. Perspectives

Genome mining is a core component in the revelation of unknown BGCs in fungal genomes. The data sources that are used for predictions are very diverse across the platforms and include not only DNA or protein sequences but also transcript data, phylogenetic data, and more. Deploying more and more predictive parameters that are connected to BGCs and their regulation will surely promote the discovery of new BGCs in the future. Increasing insights into novel BGC types (i.e., gene types, protein domains, genomic arrangement, backbone genes, etc.) will also lead to a transition from “high novelty/low reliability” to “low novelty/high reliability” BGC prediction algorithms which, in turn, again increases their discovery rate. Approaches for the detection of novel BGC types are very diverse and will surely enable us to discover new and currently unknown BGC types in heavily researched model organisms.
The development of these tools, however, is similar to a feedback loop between prediction and experimental verification. Predictions of known BGCs largely depend on the detection of patterns from already verified BGC types and so, the provision of high-quality experimental verification data of BGCs is a crucial component of reliable predictions. Curated databases fulfill these requirements but need the regular attention of professionals in the field. The guided deposition of such data by the research group itself into curated databases will hopefully become a standard procedure within the publication process in the future. This would increase the reliability of predictions and also accelerate the discovery process of BGCs in fungal genomes and, hence, the development of novel pharmaceutics.
CRISPR/Cas technology is emerging as a new tool and already has proven to deliver new molecules by the activation of silent BGCs in a number of organisms. Its functionality as a transcriptional activator was already experimentally shown in fungi by fusing the trans-activator VPR to dCas9 and dCas12a. Heterochromatic regions, however, could get problematic because of target site accessibility and the sole deployment of trans-activator domains.
The targeted local histone modifications could represent the next step in the toolbox for the activation of silent BGCs. If it is possible to relieve heterochromatic regions of silent marks (i.e., H3K27me3, H3K9me3), the targeted and locus-specific transition to transcriptionally active euchromatin could be within reach. The choice of the right histone-modification domain, however, is still a topic that needs thorough research. Many observations that connect transcriptional activation and histone modification indicate that H3K9 acetylation (gene activation especially in the case of BGCs), deployed by Rtt109 and GcnE, and H3K27 acetylation (hallmark for open chromatin and de-repression in N. crassa), also deployed by GcnE, within the promoter region correlate best with gene activation across several fungal species. As there are often nucleosome-free regions in promoters, even within heterochromatin, nucleosome positioning maps assist with the choice of a proper entry point of dCas-fusion proteins. Depending on the target locus and the prevalent histone marks, additional activators (e.g., VPR) could be necessary to induce target genes. H3S28ph also seems a promising way of activating single genes but more research, especially in the fungal context is needed in this regard. Furthermore, if DNA motifs are also proven to be responsible for occurring histone marks in fungi, this would be an additional opportunity to relieve whole BGCs from heterochromatin. This way, the programmable histone modifier can be used in combination with these mutants to activate silent BGCs and reveal their biosynthetic product.

Author Contributions

Conceptualization, A.S., L.S.-R. and J.S.; Writing—original draft preparation, A.S.; writing—review and editing, A.S., L.S.-R. and J.S.; visualization, A.S.; supervision, L.S.-R. and J.S.; funding acquisition, J.S. All authors have read and agreed to the published version of the manuscript.

Funding

Open access funding provided by Austrian Science Fund (FWF). Work was funded by grant LS19-009 “TASSMATA” from the NFB Lower Austria Science Fund and grant P32790 “ChroCosm” from the FWF Austrian Science Fund to J.S.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Thanks to Harald Berger for giving feedback and evaluating the “Genome Mining” section. Open Access Funding by the Austrian Science Fund (FWF).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pham, J.V.; Yilma, M.A.; Feliz, A.; Majid, M.T.; Maffetone, N.; Walker, J.R.; Kim, E.; Cho, H.J.; Reynolds, J.M.; Song, M.C.; et al. A Review of the Microbial Production of Bioactive Natural Products and Biologics. Front. Microbiol. 2019, 10, 1404. [Google Scholar] [CrossRef]
  2. Fleming, A. On the antibacterial action of cultures of a penicillium, with special reference to their use in the isolation of B. influenzae. Br. J. Exp. Pathol. 1929, 10, 226. [Google Scholar] [CrossRef]
  3. Houbraken, J.; Frisvad, J.C.; Samson, R.A. Fleming’s penicillin producing strain is not Penicillium chrysogenum but P. rubens. IMA Fungus 2011, 2, 87–95. [Google Scholar] [CrossRef]
  4. Olsen, A. The Antibiotic Revolution. Kuhn’s Philos. Sci. 2013, 17. Available online: https://digitalcommons.providence.edu/kuhn_2013/6 (accessed on 28 July 2022).
  5. Dhillion, S.S.; Svarstad, H.; Amundsen, C.; Bugge, H.C. Bioprospecting: Effects on Environment and Development. AMBIO A J. Hum. Environ. 2002, 31, 491–493. [Google Scholar] [CrossRef]
  6. Laupacis, A.; Keown, P.A.; Ulan, R.A.; McKenzie, N.; Stiller, C.R. Cyclosporin A: A powerful immunosuppressant. Can. Med. Assoc. J. 1982, 126, 1041–1046. [Google Scholar]
  7. Moore, R.N.; Bigam, G.; Chan, J.K.; Hogg, A.M.; Nakashima, T.T.; Vederas, J.C. Biosynthesis of the hypocholesterolemic agent mevinolin by Aspergillus terreus. Determination of the origin of carbon, hydrogen, and oxygen atoms by carbon-13 NMR and mass spectrometry. J. Am. Chem. Soc. 1985, 107, 3694–3701. [Google Scholar] [CrossRef]
  8. Goldman, G.H.; Osmani, S.A. The Aspergilli: Genomics, Medical Aspects, Biotechnology, and Research Methods; CRC press: Boca Raton, FL, USA, 2007. [Google Scholar]
  9. Nakkeeran, S.; Rajamanickam, S.; Karthikeyan, M.; Mahendra, K.; Renukadevi, P.; Johnson, I. 11—Antimicrobial secondary metabolites from Trichoderma spp. as next generation fungicides. In Biocontrol Agents and Secondary Metabolites; Jogaiah, S., Ed.; Woodhead Publishing: Sawston, UK, 2021; pp. 257–282. [Google Scholar]
  10. Savita; Sharma, A. Fungi as Biological Control Agents. In Biofertilizers for Sustainable Agriculture and Environment; Giri, B., Prasad, R., Wu, Q.-S., Varma, A., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 395–411. [Google Scholar]
  11. Segaran, G.; Sathiavelu, M. Fungal endophytes: A potent biocontrol agent and a bioactive metabolites reservoir. Biocatal. Agric. Biotechnol. 2019, 21, 101284. [Google Scholar] [CrossRef]
  12. Shi, T.Q.; Peng, H.; Zeng, S.Y.; Ji, R.Y.; Shi, K.; Huang, H.; Ji, X.J. Microbial production of plant hormones: Opportunities and challenges. Bioengineered 2017, 8, 124–128. [Google Scholar] [CrossRef]
  13. Xu, K.; Li, X.Q.; Zhao, D.L.; Zhang, P. Antifungal Secondary Metabolites Produced by the Fungal Endophytes: Chemical Diversity and Potential Use in the Development of Biopesticides. Front. Microbiol. 2021, 12, 689527. [Google Scholar] [CrossRef]
  14. Du, F.Y.; Li, X.M.; Sun, Z.C.; Meng, L.H.; Wang, B.G. Secondary Metabolites with Agricultural Antagonistic Potentials from Beauveria felina, a Marine-Derived Entomopathogenic Fungus. J. Agric. Food Chem. 2020, 68, 14824–14831. [Google Scholar] [CrossRef]
  15. Sanchez, S.; Demain, A.L. Bioactive Products from Fungi. In Food Bioactives: Extraction and Biotechnology Applications; Puri, M., Ed.; Springer International Publishing: Cham, Switzerland, 2017; pp. 59–87. [Google Scholar]
  16. Copetti, M.V. Fungi as industrial producers of food ingredients. Curr. Opin. Food Sci. 2019, 25, 52–56. [Google Scholar] [CrossRef]
  17. Tacconelli, E.; Carrara, E.; Savoldi, A.; Harbarth, S.; Mendelson, M.; Monnet, D.L.; Pulcini, C.; Kahlmeter, G.; Kluytmans, J.; Carmeli, Y.; et al. Discovery, research, and development of new antibiotics: The WHO priority list of antibiotic-resistant bacteria and tuberculosis. Lancet Infect. Dis. 2018, 18, 318–327. [Google Scholar] [CrossRef]
  18. Vylkova, S. Environmental pH modulation by pathogenic fungi as a strategy to conquer the host. PLoS Pathog. 2017, 13, e1006149. [Google Scholar] [CrossRef]
  19. Liu, Y.-N.; Zhang, T.-J.; Lu, X.-X.; Ma, B.-L.; Ren, A.; Shi, L.; Jiang, A.-L.; Yu, H.-S.; Zhao, M.-W. Membrane fluidity is involved in the regulation of heat stress induced secondary metabolism in Ganoderma lucidum. Environ. Microbiol. 2017, 19, 1653–1668. [Google Scholar] [CrossRef]
  20. Sandmann, G. Carotenoids and Their Biosynthesis in Fungi. Molecules 2022, 27, 1431. [Google Scholar] [CrossRef]
  21. Wong, H.J.; Mohamad-Fauzi, N.; Rizman-Idid, M.; Convey, P.; Alias, S.A. Protective mechanisms and responses of micro-fungi towards ultraviolet-induced cellular damage. Polar Sci. 2019, 20, 19–34. [Google Scholar] [CrossRef]
  22. Batey, S.F.D.; Greco, C.; Hutchings, M.I.; Wilkinson, B. Chemical warfare between fungus-growing ants and their pathogens. Curr. Opin. Chem. Biol. 2020, 59, 172–181. [Google Scholar] [CrossRef]
  23. Preda, V.G.; Săndulescu, O. Communication is the key: Biofilms, quorum sensing, formation and prevention. Discoveries 2019, 7, e100. [Google Scholar] [CrossRef] [PubMed]
  24. Perincherry, L.; Lalak-Kańczugowska, J.; Stępień, Ł. Fusarium-Produced Mycotoxins in Plant-Pathogen Interactions. Toxins 2019, 11, 664. [Google Scholar] [CrossRef] [PubMed]
  25. Pusztahelyi, T.; Holb, I.; Pócsi, I. Secondary metabolites in fungus-plant interactions. Front. Plant Sci. 2015, 6, 573. [Google Scholar] [CrossRef]
  26. Schrettl, M.; Haas, H. Iron homeostasis—Achilles’ heel of Aspergillus fumigatus? Curr. Opin. Microbiol. 2011, 14, 400–405. [Google Scholar] [CrossRef] [Green Version]
  27. Haas, H. How to trigger a fungal weapon. eLife 2015, 4, e10504. [Google Scholar] [CrossRef]
  28. Robey, M.T.; Caesar, L.K.; Drott, M.T.; Keller, N.P.; Kelleher, N.L. An interpreted atlas of biosynthetic gene clusters from 1,000 fungal genomes. Proc. Natl. Acad. Sci. USA 2021, 118, e2020230118. [Google Scholar] [CrossRef]
  29. Keller, N.P. Fungal secondary metabolism: Regulation, function and drug discovery. Nat. Rev. Microbiol. 2019, 17, 167–180. [Google Scholar] [CrossRef]
  30. Heo, Y.M.; Oh, S.-Y.; Kim, K.; Han, S.-I.; Kwon, S.L.; Yoo, Y.; Kim, D.; Khim, J.S.; Kang, S.; Lee, H.; et al. Comparative Genomics and Transcriptomics Depict Marine Algicolous Arthrinium Species as Endosymbionts That Help Regulate Oxidative Stress in Brown Algae. Front. Mar. Sci. 2021, 8, 753222. [Google Scholar] [CrossRef]
  31. Martín, J.F. Transport systems, intracellular traffic of intermediates and secretion of β-lactam antibiotics in fungi. Fungal Biol. Biotechnol. 2020, 7, 6. [Google Scholar] [CrossRef]
  32. Keller, N.P. Translating biosynthetic gene clusters into fungal armor and weaponry. Nat. Chem. Biol. 2015, 11, 671–677. [Google Scholar] [CrossRef]
  33. Westphal, K.R.; Muurmann, A.T.; Paulsen, I.E.; Nørgaard, K.T.H.; Overgaard, M.L.; Dall, S.M.; Aalborg, T.; Wimmer, R.; Sørensen, J.L.; Sondergaard, T.E. Who Needs Neighbors? PKS8 Is a Stand-Alone Gene in Fusarium graminearum Responsible for Production of Gibepyrones and Prolipyrone B. Molecules 2018, 23, 2232. [Google Scholar] [CrossRef]
  34. Terabayashi, Y.; Sano, M.; Yamane, N.; Marui, J.; Tamano, K.; Sagara, J.; Dohmoto, M.; Oda, K.; Ohshima, E.; Tachibana, K.; et al. Identification and characterization of genes responsible for biosynthesis of kojic acid, an industrially important compound from Aspergillus oryzae. Fungal Genet. Biol. 2010, 47, 953–961. [Google Scholar] [CrossRef]
  35. Wiemann, P.; Guo, C.J.; Palmer, J.M.; Sekonyela, R.; Wang, C.C.; Keller, N.P. Prototype of an intertwined secondary-metabolite supercluster. Proc. Natl. Acad. Sci. USA 2013, 110, 17065–17070. [Google Scholar] [CrossRef]
  36. Dai, G.; Shen, Q.; Zhang, Y.; Bian, X. Biosynthesis of Fungal Natural Products Involving Two Separate Pathway Crosstalk. J. Fungi 2022, 8, 320. [Google Scholar] [CrossRef]
  37. Gluck-Thaler, E.; Haridas, S.; Binder, M.; Grigoriev, I.V.; Crous, P.W.; Spatafora, J.W.; Bushley, K.; Slot, J.C. The Architecture of Metabolism Maximizes Biosynthetic Diversity in the Largest Class of Fungi. Mol. Biol. Evol. 2020, 37, 2838–2856. [Google Scholar] [CrossRef]
  38. Doroghazi, J.R.; Albright, J.C.; Goering, A.W.; Ju, K.S.; Haines, R.R.; Tchalukov, K.A.; Labeda, D.P.; Kelleher, N.L.; Metcalf, W.W. A roadmap for natural product discovery based on large-scale genomics and metabolomics. Nat. Chem. Biol. 2014, 10, 963–968. [Google Scholar] [CrossRef]
  39. Lind, A.L.; Wisecaver, J.H.; Lameiras, C.; Wiemann, P.; Palmer, J.M.; Keller, N.P.; Rodrigues, F.; Goldman, G.H.; Rokas, A. Drivers of genetic diversity in secondary metabolic gene clusters within a fungal species. PLOS Biol. 2017, 15, e2003583. [Google Scholar] [CrossRef]
  40. Fedorova, N.D.; Khaldi, N.; Joardar, V.S.; Maiti, R.; Amedeo, P.; Anderson, M.J.; Crabtree, J.; Silva, J.C.; Badger, J.H.; Albarraq, A.; et al. Genomic Islands in the Pathogenic Filamentous Fungus Aspergillus fumigatus. PLOS Genet. 2008, 4, e1000046. [Google Scholar] [CrossRef]
  41. Andersen, M.R.; Salazar, M.P.; Schaap, P.J.; van de Vondervoort, P.J.I.; Culley, D.; Thykaer, J.; Frisvad, J.C.; Nielsen, K.F.; Albang, R.; Albermann, K.; et al. Comparative genomics of citric-acid-producing Aspergillus niger ATCC 1015 versus enzyme-producing CBS 513.88. Genome Res. 2011, 21, 885–897. [Google Scholar] [CrossRef]
  42. Gibbons, J.G.; Salichos, L.; Slot, J.C.; Rinker, D.C.; McGary, K.L.; King, J.G.; Klich, M.A.; Tabb, D.L.; McDonald, W.H.; Rokas, A. The Evolutionary Imprint of Domestication on Genome Variation and Function of the Filamentous Fungus Aspergillus oryzae. Curr. Biol. 2012, 22, 1403–1409. [Google Scholar] [CrossRef]
  43. Rokas, A.; Wisecaver, J.H.; Lind, A.L. The birth, evolution and death of metabolic gene clusters in fungi. Nat. Rev. Microbiol. 2018, 16, 731–744. [Google Scholar] [CrossRef]
  44. Cheng, K.; Rong, X.; Pinto-Tomás, A.A.; Fernández-Villalobos, M.; Murillo-Cruz, C.; Huang, Y.; Voordouw, G. Population Genetic Analysis of Streptomyces albidoflavus Reveals Habitat Barriers to Homologous Recombination in the Diversification of Streptomycetes. Appl. Environ. Microbiol. 2015, 81, 966–975. [Google Scholar] [CrossRef]
  45. Dal Grande, F.; Jamilloux, V.; Choisne, N.; Calchera, A.; Rolshausen, G.; Petersen, M.; Schulz, M.; Nilsson, M.A.; Schmitt, I. Transposable Elements in the Genome of the Lichen-Forming Fungus Umbilicaria pustulata and Their Distribution in Different Climate Zones along Elevation. Biology 2022, 11, 24. [Google Scholar] [CrossRef]
  46. Machida, M.; Asai, K.; Sano, M.; Tanaka, T.; Kumagai, T.; Terai, G.; Kusumoto, K.; Arima, T.; Akita, O.; Kashiwagi, Y.; et al. Genome sequencing and analysis of Aspergillus oryzae. Nature 2005, 438, 1157–1161. [Google Scholar] [CrossRef] [Green Version]
  47. Takeda, I.; Umemura, M.; Koike, H.; Asai, K.; Machida, M. Motif-independent prediction of a secondary metabolism gene cluster using comparative genomics: Application to sequenced genomes of Aspergillus and ten other filamentous fungal species. DNA Res. Int. J. Rapid Publ. Rep. Genes Genomes 2014, 21, 447–457. [Google Scholar] [CrossRef]
  48. Brown, C.A.; Murray, A.W.; Verstrepen, K.J. Rapid expansion and functional divergence of subtelomeric gene families in yeasts. Curr. Biol. 2010, 20, 895–903. [Google Scholar] [CrossRef] [PubMed]
  49. Graham-Taylor, C.; Kamphuis, L.G.; Derbyshire, M.C. A detailed in silico analysis of secondary metabolite biosynthesis clusters in the genome of the broad host range plant pathogenic fungus Sclerotinia sclerotiorum. BMC Genom. 2020, 21, 7. [Google Scholar] [CrossRef] [PubMed]
  50. Rehmeyer, C.; Li, W.; Kusaba, M.; Kim, Y.-S.; Brown, D.; Staben, C.; Dean, R.; Farman, M. Organization of chromosome ends in the rice blast fungus, Magnaporthe oryzae. Nucleic Acids Res. 2006, 34, 4685–4701. [Google Scholar] [CrossRef] [PubMed]
  51. Farman, M.L. Telomeres in the rice blast fungus Magnaporthe oryzae: The world of the end as we know it. FEMS Microbiol. Lett. 2007, 273, 125–132. [Google Scholar] [CrossRef]
  52. McDonagh, A.; Fedorova, N.D.; Crabtree, J.; Yu, Y.; Kim, S.; Chen, D.; Loss, O.; Cairns, T.; Goldman, G.; Armstrong-James, D.; et al. Sub-Telomere Directed Gene Expression during Initiation of Invasive Aspergillosis. PLOS Pathog. 2008, 4, e1000154. [Google Scholar] [CrossRef]
  53. Cairns, T.; Meyer, V. In silico prediction and characterization of secondary metabolite biosynthetic gene clusters in the wheat pathogen Zymoseptoria tritici. BMC Genom. 2017, 18, 631. [Google Scholar] [CrossRef]
  54. Morris, J.J.; Lenski, R.E.; Zinser, E.R. The Black Queen Hypothesis: Evolution of dependencies through adaptive gene loss. mBio 2012, 3. [Google Scholar] [CrossRef]
  55. Burkhardt, I.; Siemon, T.; Henrot, M.; Studt, L.; Rosler, S.; Tudzynski, B.; Christmann, M.; Dickschat, J.S. Mechanistic Characterisation of Two Sesquiterpene Cyclases from the Plant Pathogenic Fungus Fusarium fujikuroi. Angew. Chem. Int. Ed. Engl. 2016, 55, 8748–8751. [Google Scholar] [CrossRef] [PubMed]
  56. Torres, D.E.; Thomma, B.P.H.J.; Seidl, M.F. Transposable Elements Contribute to Genome Dynamics and Gene Expression Variation in the Fungal Plant Pathogen Verticillium dahliae. Genome Biol. Evol. 2021, 13, evab135. [Google Scholar] [CrossRef] [PubMed]
  57. Valero-Jiménez, C.A.; Steentjes, M.B.F.; Slot, J.C.; Shi-Kunne, X.; Scholten, O.E.; van Kan, J.A.L. Dynamics in Secondary Metabolite Gene Clusters in Otherwise Highly Syntenic and Stable Genomes in the Fungal Genus Botrytis. Genome Biol. Evol. 2020, 12, 2491–2507. [Google Scholar] [CrossRef]
  58. Porquier, A.; Morgant, G.; Moraga, J.; Dalmais, B.; Luyten, I.; Simon, A.; Pradier, J.-M.; Amselem, J.; Collado, I.G.; Viaud, M. The botrydial biosynthetic gene cluster of Botrytis cinerea displays a bipartite genomic structure and is positively regulated by the putative Zn(II)2Cys6 transcription factor BcBot6. Fungal Genet. Biol. 2016, 96, 33–46. [Google Scholar] [CrossRef] [PubMed]
  59. Rokas, A.; Mead, M.E.; Steenwyk, J.L.; Raja, H.A.; Oberlies, N.H. Biosynthetic gene clusters and the evolution of fungal chemodiversity. Nat. Prod. Rep. 2020, 37, 868–878. [Google Scholar] [CrossRef]
  60. Slot, J.C.; Rokas, A. Horizontal Transfer of a Large and Highly Toxic Secondary Metabolic Gene Cluster between Fungi. Curr. Biol. 2011, 21, 134–139. [Google Scholar] [CrossRef]
  61. Campbell, M.A.; Rokas, A.; Slot, J.C. Horizontal transfer and death of a fungal secondary metabolic gene cluster. Genome Biol. Evol. 2012, 4, 289–293. [Google Scholar] [CrossRef]
  62. Dhillon, B.; Feau, N.; Aerts, A.L.; Beauseigle, S.; Bernier, L.; Copeland, A.; Foster, A.; Gill, N.; Henrissat, B.; Herath, P.; et al. Horizontal gene transfer and gene dosage drives adaptation to wood colonization in a tree pathogen. Proc. Natl. Acad. Sci. USA 2015, 112, 3451–3456. [Google Scholar] [CrossRef]
  63. Jia, Q.; Chen, X.; Köllner, T.G.; Rinkel, J.; Fu, J.; Labbé, J.; Xiong, W.; Dickschat, J.S.; Gershenzon, J.; Chen, F. Terpene Synthase Genes Originated from Bacteria through Horizontal Gene Transfer Contribute to Terpenoid Diversity in Fungi. Sci. Rep. 2019, 9, 9223. [Google Scholar] [CrossRef]
  64. Proctor, R.H.; Van Hove, F.; Susca, A.; Stea, G.; Busman, M.; van der Lee, T.; Waalwijk, C.; Moretti, A.; Ward, T.J. Birth, death and horizontal transfer of the fumonisin biosynthetic gene cluster during the evolutionary diversification of Fusarium. Mol. Microbiol. 2013, 90, 290–306. [Google Scholar] [CrossRef]
  65. Reynolds, H.T.; Vijayakumar, V.; Gluck-Thaler, E.; Korotkin, H.B.; Matheny, P.B.; Slot, J.C. Horizontal gene cluster transfer increased hallucinogenic mushroom diversity. Evol. Lett. 2018, 2, 88–101. [Google Scholar] [CrossRef]
  66. Schumacher, J.; Gautier, A.; Morgant, G.; Studt, L.; Ducrot, P.-H.; Le Pêcheur, P.; Azeddine, S.; Fillinger, S.; Leroux, P.; Tudzynski, B.; et al. A Functional Bikaverin Biosynthesis Gene Cluster in Rare Strains of Botrytis cinerea Is Positively Controlled by VELVET. PLoS ONE 2013, 8, e53729. [Google Scholar] [CrossRef] [Green Version]
  67. Kohli, G.S.; John, U.; Van Dolah, F.M.; Murray, S.A. Evolutionary distinctiveness of fatty acid and polyketide synthesis in eukaryotes. ISME J. 2016, 10, 1877–1890. [Google Scholar] [CrossRef] [PubMed]
  68. Alexander, N.J.; Proctor, R.H.; McCormick, S.P. Genes, gene clusters, and biosynthesis of trichothecenes and fumonisins in Fusarium. Toxin Rev. 2009, 28, 198–215. [Google Scholar] [CrossRef]
  69. Min, B.; Grigoriev, I.V.; Choi, I.-G. FunGAP: Fungal Genome Annotation Pipeline using evidence-based gene model evaluation. Bioinformatics 2017, 33, 2936–2937. [Google Scholar] [CrossRef] [PubMed]
  70. Blin, K.; Kim, H.U.; Medema, M.H.; Weber, T. Recent development of antiSMASH and other computational approaches to mine secondary metabolite biosynthetic gene clusters. Brief Bioinform 2019, 20, 1103–1113. [Google Scholar] [CrossRef]
  71. Cruz-Morales, P.; Kopp, J.F.; Martínez-Guerrero, C.; Yáñez-Guerra, L.A.; Selem-Mojica, N.; Ramos-Aboites, H.; Feldmann, J.; Barona-Gómez, F. Phylogenomic Analysis of Natural Products Biosynthetic Gene Clusters Allows Discovery of Arseno-Organic Metabolites in Model Streptomycetes. Genome Biol. Evol. 2016, 8, 1906–1916. [Google Scholar] [CrossRef]
  72. Vignolle, G.A.; Schaffer, D.; Zehetner, L.; Mach, R.L.; Mach-Aigner, A.R.; Derntl, C. FunOrder: A robust and semi-automated method for the identification of essential biosynthetic genes through computational molecular co-evolution. PLOS Comput. Biol. 2021, 17, e1009372. [Google Scholar] [CrossRef]
  73. Walker, A.S.; Clardy, J. A Machine Learning Bioinformatics Method to Predict Biological Activity from Biosynthetic Gene Clusters. J. Chem. Inf. Modeling 2021, 61, 2560–2571. [Google Scholar] [CrossRef]
  74. Qiang, B.; Lai, J.; Jin, H.; Zhang, L.; Liu, Z. Target Prediction Model for Natural Products using Transfer Learning. Int. J. Mol. Sci. 2021, 22, 4632. [Google Scholar] [CrossRef]
  75. Moumbock, A.F.A.; Li, J.; Mishra, P.; Gao, M.; Günther, S. Current computational methods for predicting protein interactions of natural products. Comput. Struct. Biotechnol. J. 2019, 17, 1367–1376. [Google Scholar] [CrossRef] [PubMed]
  76. Mohimani, H.; Gurevich, A.; Shlemov, A.; Mikheenko, A.; Korobeynikov, A.; Cao, L.; Shcherbin, E.; Nothias, L.-F.; Dorrestein, P.C.; Pevzner, P.A. Dereplication of microbial metabolites through database search of mass spectra. Nat. Commun. 2018, 9, 4035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Medema, M.H.; Fischbach, M.A. Computational approaches to natural product discovery. Nat. Chem. Biol. 2015, 11, 639–648. [Google Scholar] [CrossRef] [PubMed]
  78. Blin, K.; Shaw, S.; Kloosterman, A.M.; Charlop-Powers, Z.; van Wezel, G.P.; Medema, M.H.; Weber, T. antiSMASH 6.0: Improving cluster detection and comparison capabilities. Nucleic Acids Res. 2021, 49, W29–W35. [Google Scholar] [CrossRef]
  79. Khaldi, N.; Seifuddin, F.T.; Turner, G.; Haft, D.; Nierman, W.C.; Wolfe, K.H.; Fedorova, N.D. SMURF: Genomic mapping of fungal secondary metabolite clusters. Fungal Genet. Biol. 2010, 47, 736–741. [Google Scholar] [CrossRef]
  80. Sélem-Mojica, N.; Aguilar, C.; Gutiérrez-García, K.; Martínez-Guerrero, C.E.; Barona-Gómez, F. EvoMining reveals the origin and fate of natural product biosynthetic enzymes. Microb. Genom. 2019, 5, e000260. [Google Scholar] [CrossRef]
  81. Blin, K.; Wolf, T.; Chevrette, M.G.; Lu, X.; Schwalen, C.J.; Kautsar, S.A.; Suarez Duran, H.G.; de los Santos, E.L.C.; Kim, H.U.; Nave, M.; et al. antiSMASH 4.0—improvements in chemistry prediction and gene cluster boundary identification. Nucleic Acids Res. 2017, 45, W36–W41. [Google Scholar] [CrossRef]
  82. Cimermancic, P.; Medema, M.H.; Claesen, J.; Kurita, K.; Wieland Brown, L.C.; Mavrommatis, K.; Pati, A.; Godfrey, P.A.; Koehrsen, M.; Clardy, J.; et al. Insights into secondary metabolism from a global analysis of prokaryotic biosynthetic gene clusters. Cell 2014, 158, 412–421. [Google Scholar] [CrossRef]
  83. Mungan, M.D.; Alanjary, M.; Blin, K.; Weber, T.; Medema, M.H.; Ziemert, N. ARTS 2.0: Feature updates and expansion of the Antibiotic Resistant Target Seeker for comparative genome mining. Nucleic Acids Res. 2020, 48, W546–W552. [Google Scholar] [CrossRef]
  84. Alanjary, M.; Kronmiller, B.; Adamek, M.; Blin, K.; Weber, T.; Huson, D.; Philmus, B.; Ziemert, N. The Antibiotic Resistant Target Seeker (ARTS), an exploration engine for antibiotic cluster prioritization and novel drug target discovery. Nucleic Acids Res. 2017, 45, W42–W48. [Google Scholar] [CrossRef]
  85. Kjærbølling, I.; Vesth, T.; Andersen, M.R.; Sessions, P.F.D. Resistance Gene-Directed Genome Mining of 50 Aspergillus Species. mSystems 2019, 4, e00085-19. [Google Scholar] [CrossRef]
  86. Umemura, M.; Koike, H.; Nagano, N.; Ishii, T.; Kawano, J.; Yamane, N.; Kozone, I.; Horimoto, K.; Shin-ya, K.; Asai, K.; et al. MIDDAS-M: Motif-Independent De Novo Detection of Secondary Metabolite Gene Clusters through the Integration of Genome Sequencing and Transcriptome Data. PLoS ONE 2014, 8, e84028. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Harrison, P.W.; Ahamed, A.; Aslam, R.; Alako, B.T.F.; Burgin, J.; Buso, N.; Courtot, M.; Fan, J.; Gupta, D.; Haseeb, M.; et al. The European Nucleotide Archive in 2020. Nucleic Acids Res. 2021, 49, D82–D85. [Google Scholar] [CrossRef] [PubMed]
  88. Kanz, C.; Aldebert, P.; Althorpe, N.; Baker, W.; Baldwin, A.; Bates, K.; Browne, P.; van den Broek, A.; Castro, M.; Cochrane, G.; et al. The EMBL Nucleotide Sequence Database. Nucleic Acids Res. 2005, 33, D29–D33. [Google Scholar] [CrossRef] [PubMed]
  89. Benson, D.A.; Cavanaugh, M.; Clark, K.; Karsch-Mizrachi, I.; Lipman, D.J.; Ostell, J.; Sayers, E.W. GenBank. Nucleic Acids Res. 2012, 41, D36–D42. [Google Scholar] [CrossRef] [PubMed]
  90. Wang, M.; Carver, J.J.; Phelan, V.V.; Sanchez, L.M.; Garg, N.; Peng, Y.; Nguyen, D.D.; Watrous, J.; Kapono, C.A.; Luzzatto-Knaan, T.; et al. Sharing and community curation of mass spectrometry data with Global Natural Products Social Molecular Networking. Nat. Biotechnol. 2016, 34, 828–837. [Google Scholar] [CrossRef]
  91. Van Santen, J.A.; Poynton, E.F.; Iskakova, D.; McMann, E.; Alsup, T.A.; Clark, T.N.; Fergusson, C.H.; Fewer, D.P.; Hughes, A.H.; McCadden, C.A.; et al. The Natural Products Atlas 2.0: A database of microbially-derived natural products. Nucleic Acids Res. 2021, 50, D1317–D1323. [Google Scholar] [CrossRef]
  92. Haft, D.H.; Selengut, J.D.; Richter, R.A.; Harkins, D.; Basu, M.K.; Beck, E. TIGRFAMs and Genome Properties in 2013. Nucleic Acids Res. 2012, 41, D387–D395. [Google Scholar] [CrossRef]
  93. Finn, R.D.; Tate, J.; Mistry, J.; Coggill, P.C.; Sammut, S.J.; Hotz, H.-R.; Ceric, G.; Forslund, K.; Eddy, S.R.; Sonnhammer, E.L.L.; et al. The Pfam protein families database. Nucleic Acids Res. 2008, 36, D281–D288. [Google Scholar] [CrossRef]
  94. Eddy, S.R. Accelerated Profile HMM Searches. PLOS Comput. Biol. 2011, 7, e1002195. [Google Scholar] [CrossRef]
  95. Almeida, H.; Palys, S.; Tsang, A.; Diallo, A.B. TOUCAN: A framework for fungal biosynthetic gene cluster discovery. NAR Genom. Bioinform. 2020, 2, lqaa098. [Google Scholar] [CrossRef] [PubMed]
  96. Blin, K.; Shaw, S.; Kautsar, S.A.; Medema, M.H.; Weber, T. The antiSMASH database version 3: Increased taxonomic coverage and new query features for modular enzymes. Nucleic Acids Res. 2020, 49, D639–D643. [Google Scholar] [CrossRef] [PubMed]
  97. Flissi, A.; Ricart, E.; Campart, C.; Chevalier, M.; Dufresne, Y.; Michalik, J.; Jacques, P.; Flahaut, C.; Lisacek, F.; Leclère, V.; et al. Norine: Update of the nonribosomal peptide resource. Nucleic Acids Res. 2019, 48, D465–D469. [Google Scholar] [CrossRef]
  98. Navarro-Muñoz, J.C.; Selem-Mojica, N.; Mullowney, M.W.; Kautsar, S.A.; Tryon, J.H.; Parkinson, E.I.; De Los Santos, E.L.C.; Yeong, M.; Cruz-Morales, P.; Abubucker, S.; et al. A computational framework to explore large-scale biosynthetic diversity. Nat. Chem. Biol. 2020, 16, 60–68. [Google Scholar] [CrossRef] [PubMed]
  99. Kautsar, S.A.; Blin, K.; Shaw, S.; Navarro-Muñoz, J.C.; Terlouw, B.R.; van der Hooft, J.J.J.; van Santen, J.A.; Tracanna, V.; Suarez Duran, H.G.; Pascal Andreu, V.; et al. MIBiG 2.0: A repository for biosynthetic gene clusters of known function. Nucleic Acids Res. 2019, 48, D454–D458. [Google Scholar] [CrossRef]
  100. Palaniappan, K.; Chen, I.-M.A.; Chu, K.; Ratner, A.; Seshadri, R.; Kyrpides, N.C.; Ivanova, N.N.; Mouncey, N.J. IMG-ABC v.5.0: An update to the IMG/Atlas of Biosynthetic Gene Clusters Knowledgebase. Nucleic Acids Res. 2019, 48, D422–D430. [Google Scholar] [CrossRef]
  101. Hannigan, G.D.; Prihoda, D.; Palicka, A.; Soukup, J.; Klempir, O.; Rampula, L.; Durcak, J.; Wurst, M.; Kotowski, J.; Chang, D.; et al. A deep learning genome-mining strategy for biosynthetic gene cluster prediction. Nucleic Acids Res. 2019, 47, e110. [Google Scholar] [CrossRef]
  102. Agrawal, P.; Khater, S.; Gupta, M.; Sain, N.; Mohanty, D. RiPPMiner: A bioinformatics resource for deciphering chemical structures of RiPPs based on prediction of cleavage and cross-links. Nucleic Acids Res. 2017, 45, W80–W88. [Google Scholar] [CrossRef]
  103. Khater, S.; Gupta, M.; Agrawal, P.; Sain, N.; Prava, J.; Gupta, P.; Grover, M.; Kumar, N.; Mohanty, D. SBSPKSv2: Structure-based sequence analysis of polyketide synthases and non-ribosomal peptide synthetases. Nucleic Acids Res. 2017, 45, W72–W79. [Google Scholar] [CrossRef]
  104. Anand, S.; Prasad, M.V.R.; Yadav, G.; Kumar, N.; Shehara, J.; Ansari, M.Z.; Mohanty, D. SBSPKS: Structure based sequence analysis of polyketide synthases. Nucleic Acids Res. 2010, 38, W487–W496. [Google Scholar] [CrossRef]
  105. Zierep, P.F.; Padilla, N.; Yonchev, D.G.; Telukunta, K.K.; Klementz, D.; Günther, S. SeMPI: A genome-based secondary metabolite prediction and identification web server. Nucleic Acids Res. 2017, 45, W64–W71. [Google Scholar] [CrossRef] [PubMed]
  106. Wolf, T.; Shelest, V.; Nath, N.; Shelest, E. CASSIS and SMIPS: Promoter-based prediction of secondary metabolite gene clusters in eukaryotic genomes. Bioinformatics 2015, 32, 1138–1143. [Google Scholar] [CrossRef] [Green Version]
  107. Vesth, T.C.; Brandl, J.; Andersen, M.R. FunGeneClusterS: Predicting fungal gene clusters from genome and transcriptome data. Synth. Syst. Biotechnol. 2016, 1, 122–129. [Google Scholar] [CrossRef]
  108. Andersen, M.R.; Nielsen, J.B.; Klitgaard, A.; Petersen, L.M.; Zachariasen, M.; Hansen, T.J.; Blicher, L.H.; Gotfredsen, C.H.; Larsen, T.O.; Nielsen, K.F.; et al. Accurate prediction of secondary metabolite gene clusters in filamentous fungi. Proc. Natl. Acad. Sci. USA 2013, 110, E99–E107. [Google Scholar] [CrossRef] [PubMed]
  109. Bachmann, B.O.; Ravel, J. Chapter 8 Methods for In Silico Prediction of Microbial Polyketide and Nonribosomal Peptide Biosynthetic Pathways from DNA Sequence Data. In Methods in Enzymology; Academic Press: Cambridge, MA, USA, 2009; Volume 458, pp. 181–217. [Google Scholar]
  110. Johnston, C.W.; Skinnider, M.A.; Wyatt, M.A.; Li, X.; Ranieri, M.R.M.; Yang, L.; Zechel, D.L.; Ma, B.; Magarvey, N.A. An automated Genomes-to-Natural Products platform (GNP) for the discovery of modular natural products. Nat. Commun. 2015, 6, 8421. [Google Scholar] [CrossRef] [PubMed]
  111. Dufresne, Y.; Noé, L.; Leclère, V.; Pupin, M. Smiles2Monomers: A link between chemical and biological structures for polymers. J. Cheminform. 2015, 7, 62. [Google Scholar] [CrossRef]
  112. Conway, K.R.; Boddy, C.N. ClusterMine360: A database of microbial PKS/NRPS biosynthesis. Nucleic Acids Res. 2012, 41, D402–D407. [Google Scholar] [CrossRef]
  113. Röttig, M.; Medema, M.H.; Blin, K.; Weber, T.; Rausch, C.; Kohlbacher, O. NRPSpredictor2—A web server for predicting NRPS adenylation domain specificity. Nucleic Acids Res. 2011, 39, W362–W367. [Google Scholar] [CrossRef]
  114. Weber, T.; Rausch, C.; Lopez, P.; Hoof, I.; Gaykova, V.; Huson, D.H.; Wohlleben, W. CLUSEAN: A computer-based framework for the automated analysis of bacterial secondary metabolite biosynthetic gene clusters. J. Biotechnol. 2009, 140, 13–17. [Google Scholar] [CrossRef]
  115. Bode, H.B.; Bethe, B.; Höfs, R.; Zeeck, A. Big Effects from Small Changes: Possible Ways to Explore Nature’s Chemical Diversity. ChemBioChem 2002, 3, 619–627. [Google Scholar] [CrossRef]
  116. Romano, S.; Jackson, S.A.; Patry, S.; Dobson, A.D.W. Extending the “One Strain Many Compounds” (OSMAC) Principle to Marine Microorganisms. Mar. Drugs 2018, 16, 244. [Google Scholar] [CrossRef] [PubMed]
  117. Braga, G.U.; Rangel, D.E.; Fernandes, É.K.; Flint, S.D.; Roberts, D.W. Molecular and physiological effects of environmental UV radiation on fungal conidia. Curr. Genet. 2015, 61, 405–425. [Google Scholar] [CrossRef] [PubMed]
  118. Tian, F.; Lee, S.Y.; Woo, S.Y.; Choi, H.Y.; Heo, S.; Nah, G.; Chun, H.S. Transcriptomic responses of Aspergillus flavus to temperature and oxidative stresses during aflatoxin production. Sci. Rep. 2021, 11, 2803. [Google Scholar] [CrossRef]
  119. Gressler, M.; Meyer, F.; Heine, D.; Hortschansky, P.; Hertweck, C.; Brock, M. Phytotoxin production in Aspergillus terreus is regulated by independent environmental signals. eLife 2015, 4, e07861. [Google Scholar] [CrossRef] [PubMed]
  120. Iyer, M.S.; Pal, A.; Srinivasan, S.; Somvanshi, P.R.; Venkatesh, K.V. Global Transcriptional Regulators Fine-Tune the Translational and Metabolic Efficiency for Optimal Growth of Escherichia coli. mSystems 2021, 6, e00001-21. [Google Scholar] [CrossRef]
  121. Lind, A.L.; Wisecaver, J.H.; Smith, T.D.; Feng, X.; Calvo, A.M.; Rokas, A. Examining the Evolution of the Regulatory Circuit Controlling Secondary Metabolism and Development in the Fungal Genus Aspergillus. PLOS Genet. 2015, 11, e1005096. [Google Scholar] [CrossRef]
  122. 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]
  123. Leiter, É.; Emri, T.; Pákozdi, K.; Hornok, L.; Pócsi, I. The impact of bZIP Atf1ortholog global regulators in fungi. Appl. Microbiol. Biotechnol. 2021, 105, 5769–5783. [Google Scholar] [CrossRef]
  124. Pfannmüller, A.; Boysen, J.M.; Tudzynski, B. Nitrate Assimilation in Fusarium fujikuroi Is Controlled by Multiple Levels of Regulation. Front. Microbiol. 2017, 8, 381. [Google Scholar] [CrossRef]
  125. Wiemann, P.; Sieber, C.M.K.; von Bargen, K.W.; Studt, L.; Niehaus, E.-M.; Espino, J.J.; Huß, K.; Michielse, C.B.; Albermann, S.; Wagner, D.; et al. Deciphering the Cryptic Genome: Genome-wide Analyses of the Rice Pathogen Fusarium fujikuroi Reveal Complex Regulation of Secondary Metabolism and Novel Metabolites. PLOS Pathog. 2013, 9, e1003475. [Google Scholar] [CrossRef]
  126. Tudzynski, B. Nitrogen regulation of fungal secondary metabolism in fungi. Front. Microbiol. 2014, 5, 656. [Google Scholar] [CrossRef] [PubMed]
  127. Pfannmüller, A.; Leufken, J.; Studt, L.; Michielse, C.B.; Sieber, C.M.K.; Güldener, U.; Hawat, S.; Hippler, M.; Fufezan, C.; Tudzynski, B. Comparative transcriptome and proteome analysis reveals a global impact of the nitrogen regulators AreA and AreB on secondary metabolism in Fusarium fujikuroi. PLoS ONE 2017, 12, e0176194. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Scherlach, K.; Sarkar, A.; Schroeckh, V.; Dahse, H.-M.; Roth, M.; Brakhage, A.A.; Horn, U.; Hertweck, C. Two Induced Fungal Polyketide Pathways Converge into Antiproliferative Spiroanthrones. ChemBioChem 2011, 12, 1836–1839. [Google Scholar] [CrossRef]
  129. Giese, H.; Sondergaard, T.E.; Sørensen, J.L. The AreA transcription factor in Fusarium graminearum regulates the use of some nonpreferred nitrogen sources and secondary metabolite production. Fungal Biol. 2013, 117, 814–821. [Google Scholar] [CrossRef]
  130. Li, J.; Pan, Y.; Liu, G. Disruption of the nitrogen regulatory gene AcareA in Acremonium chrysogenum leads to reduction of cephalosporin production and repression of nitrogen metabolism. Fungal Genet. Biol. 2013, 61, 69–79. [Google Scholar] [CrossRef]
  131. López-Berges, M.S.; Schäfer, K.; Hera, C.; Di Pietro, A. Combinatorial function of velvet and AreA in transcriptional regulation of nitrate utilization and secondary metabolism. Fungal Genet. Biol. 2014, 62, 78–84. [Google Scholar] [CrossRef] [PubMed]
  132. Akhberdi, O.; Zhang, Q.; Wang, D.; Wang, H.; Hao, X.; Liu, Y.; Wei, D.; Zhu, X. Distinct Roles of Velvet Complex in the Development, Stress Tolerance, and Secondary Metabolism in Pestalotiopsis microspora, a Taxol Producer. Genes 2018, 9, 164. [Google Scholar] [CrossRef]
  133. Bayram, Ö.S.; Dettmann, A.; Karahoda, B.; Moloney, N.M.; Ormsby, T.; McGowan, J.; Cea-Sánchez, S.; Miralles-Durán, A.; Brancini, G.T.P.; Luque, E.M.; et al. Control of Development, Secondary Metabolism and Light-Dependent Carotenoid Biosynthesis by the Velvet Complex of Neurospora crassa. Genetics 2019, 212, 691–710. [Google Scholar] [CrossRef]
  134. Lan, N.; Yue, Q.; An, Z.; Bills, G.F. Apc.LaeA and Apc.VeA of the velvet complex govern secondary metabolism and morphological development in the echinocandin-producing fungus Aspergillus pachycristatus. J. Ind. Microbiol. Biotechnol. 2020, 47, 155–168. [Google Scholar] [CrossRef]
  135. Monroy, A.A.; Stappler, E.; Schuster, A.; Sulyok, M.; Schmoll, M. A CRE1—Regulated cluster is responsible for light dependent production of dihydrotrichotetronin in Trichoderma reesei. PLoS ONE 2017, 12, e0182530. [Google Scholar] [CrossRef]
  136. Wang, G.; Zhang, H.; Wang, Y.; Liu, F.; Li, E.; Ma, J.; Yang, B.; Zhang, C.; Li, L.; Liu, Y. Requirement of LaeA, VeA, and VelB on Asexual Development, Ochratoxin A Biosynthesis, and Fungal Virulence in Aspergillus ochraceus. Front. Microbiol. 2019, 10, 2759. [Google Scholar] [CrossRef] [PubMed]
  137. 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] [PubMed]
  138. 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]
  139. Martín, J.F. Key role of LaeA and velvet complex proteins on expression of β-lactam and PR-toxin genes in Penicillium chrysogenum: Cross-talk regulation of secondary metabolite pathways. J. Ind. Microbiol. Biotechnol. 2017, 44, 525–535. [Google Scholar] [CrossRef]
  140. Sarikaya-Bayram, Ö.; Palmer, J.M.; Keller, N.; Braus, G.H.; Bayram, Ö. One Juliet and four Romeos: VeA and its methyltransferases. Front. Microbiol. 2015, 6, 1. [Google Scholar] [CrossRef]
  141. Wiemann, P.; Brown, D.W.; Kleigrewe, K.; Bok, J.W.; Keller, N.P.; Humpf, H.-U.; Tudzynski, B. FfVel1 and FfLae1, components of a velvet-like complex in Fusarium fujikuroi, affect differentiation, secondary metabolism and virulence. Mol. Microbiol. 2010, 77, 972–994. [Google Scholar] [CrossRef]
  142. Bok, J.W.; Keller, N.P. LaeA, a Regulator of Secondary Metabolism in Aspergillus spp. Eukaryot. Cell 2004, 3, 527–535. [Google Scholar] [CrossRef]
  143. Feng, Y.; Yin, Z.; Wu, Y.; Xu, L.; Du, H.; Wang, N.; Huang, L. LaeA Controls Virulence and Secondary Metabolism in Apple Canker Pathogen Valsa mali. Front. Microbiol. 2020, 11, 581203. [Google Scholar] [CrossRef]
  144. Kumar, D.; Barad, S.; Chen, Y.; Luo, X.; Tannous, J.; Dubey, A.; Glam Matana, N.; Tian, S.; Li, B.; Keller, N.; et al. LaeA regulation of secondary metabolism modulates virulence in Penicillium expansum and is mediated by sucrose. Mol. Plant Pathol. 2017, 18, 1150–1163. [Google Scholar] [CrossRef]
  145. Jiang, T.; Wang, M.; Li, L.; Si, J.; Song, B.; Zhou, C.; Yu, M.; Wang, X.; Zhang, Y.; Ding, G.; et al. Overexpression of the Global Regulator LaeA in Chaetomium globosum Leads to the Biosynthesis of Chaetoglobosin Z. J. Nat. Prod. 2016, 79, 2487–2494. [Google Scholar] [CrossRef]
  146. Butchko, R.A.E.; Brown, D.W.; Busman, M.; Tudzynski, B.; Wiemann, P. Lae1 regulates expression of multiple secondary metabolite gene clusters in Fusarium verticillioides. Fungal Genet. Biol. 2012, 49, 602–612. [Google Scholar] [CrossRef] [PubMed]
  147. Niehaus, E.M.; Rindermann, L.; Janevska, S.; Munsterkotter, M.; Guldener, U.; Tudzynski, B. Analysis of the global regulator Lae1 uncovers a connection between Lae1 and the histone acetyltransferase HAT1 in Fusarium fujikuroi. Appl. Microbiol. Biotechnol. 2018, 102, 279–295. [Google Scholar] [CrossRef] [PubMed]
  148. Barda, O.; Maor, U.; Sadhasivam, S.; Bi, Y.; Zakin, V.; Prusky, D.; Sionov, E. The pH-Responsive Transcription Factor PacC Governs Pathogenicity and Ochratoxin A Biosynthesis in Aspergillus carbonarius. Front. Microbiol. 2020, 11, 210. [Google Scholar] [CrossRef] [PubMed]
  149. Gao, H.; Wang, Y.; Luo, Q.; Yang, L.; He, X.; Wu, J.; Kachanuban, K.; Wilaipun, P.; Zhu, W.; Wang, Y. Bioactive Metabolites from Acid-Tolerant Fungi in a Thai Mangrove Sediment. Front. Microbiol. 2021, 11, 60995. [Google Scholar] [CrossRef]
  150. 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]
  151. Shah, A.J.; Tilburn, J.; Adlard, M.W.; Arst, H.N., Jr. pH regulation of penicillin production in Aspergillus nidulans. FEMS Microbiol. Lett. 1991, 61, 209–212. [Google Scholar] [CrossRef]
  152. Trushina, N.; Levin, M.; Mukherjee, P.K.; Horwitz, B.A. PacC and pH–dependent transcriptome of the mycotrophic fungus Trichoderma virens. BMC Genom. 2013, 14, 138. [Google Scholar] [CrossRef]
  153. Wiemann, P.; Willmann, A.; Straeten, M.; Kleigrewe, K.; Beyer, M.; Humpf, H.U.; Tudzynski, B. Biosynthesis of the red pigment bikaverin in Fusarium fujikuroi: Genes, their function and regulation. Mol. Microbiol. 2009, 72, 931–946. [Google Scholar] [CrossRef]
  154. Li, B.; Chen, Y.; Tian, S. Function of pH-dependent transcription factor PacC in regulating development, pathogenicity, and mycotoxin biosynthesis of phytopathogenic fungi. FEBS J. 2022, 289, 1723–1730. [Google Scholar] [CrossRef]
  155. Merhej, J.; Richard-Forget, F.; Barreau, C. The pH regulatory factor Pac1 regulates Tri gene expression and trichothecene production in Fusarium graminearum. Fungal Genet. Biol. 2011, 48, 275–284. [Google Scholar] [CrossRef]
  156. Tilburn, J.; Sarkar, S.; Widdick, D.A.; Espeso, E.A.; Orejas, M.; Mungroo, J.; Peñalva, M.A.; Arst, H.N., Jr. 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]
  157. Lind, A.L.; Smith, T.D.; Saterlee, T.; Calvo, A.M.; Rokas, A. Regulation of Secondary Metabolism by the Velvet Complex Is Temperature-Responsive in Aspergillus. G3 2016, 6, 4023–4033. [Google Scholar] [CrossRef] [Green Version]
  158. Yogabaanu, U.; Weber, J.-F.F.; Convey, P.; Rizman-Idid, M.; Alias, S.A. Antimicrobial properties and the influence of temperature on secondary metabolite production in cold environment soil fungi. Polar Sci. 2017, 14, 60–67. [Google Scholar] [CrossRef]
  159. Overy, D.; Correa, H.; Roullier, C.; Chi, W.-C.; Pang, K.-L.; Rateb, M.; Ebel, R.; Shang, Z.; Capon, R.; Bills, G.; et al. Does Osmotic Stress Affect Natural Product Expression in Fungi? Mar. Drugs 2017, 15, 254. [Google Scholar] [CrossRef] [PubMed]
  160. Huang, J.; Lu, C.; Qian, X.; Huang, Y.; Zheng, Z.; Shen, Y. Effect of salinity on the growth, biological activity and secondary metabolites of some marine fungi. Acta Oceanol. Sin. 2011, 30, 118. [Google Scholar] [CrossRef]
  161. Sarri, E.; Termentzi, A.; Abraham, E.M.; Papadopoulos, G.K.; Baira, E.; Machera, K.; Loukas, V.; Komaitis, F.; Tani, E. Salinity Stress Alters the Secondary Metabolic Profile of M. sativa, M. arborea and Their Hybrid (Alborea). Int. J. Mol. Sci. 2021, 22, 4882. [Google Scholar] [CrossRef] [PubMed]
  162. Wang, Y.; Lu, Z.; Sun, K.; Zhu, W. Effects of High Salt Stress on Secondary Metabolite Production in the Marine-Derived Fungus Spicaria elegans. Mar. Drugs 2011, 9, 535–542. [Google Scholar] [CrossRef] [PubMed]
  163. Fil, T.P.; Pallini, H.F.; Amaral, L.D.S.; Silva, J.V.D.; Bidóia, D.L.; Peron, F.; Garcia, F.P.; Nakamura, C.V.; Rodrigues-Filho, E. Copper and Manganese Cations Alter Secondary Metabolism in the Fungus Penicillium brasilianum. J. Braz. Chem. Soc. 2016, 27, 1444–1451. [Google Scholar] [CrossRef]
  164. Scott, R.E.; Jones, A.; Lam, K.S.; Gaucher, G.M. Manganese and antibiotic biosynthesis. I. A specific manganese requirement for patulin production in Penicillium urticae. Can. J. Microbiol. 1986, 32, 259–267. [Google Scholar] [CrossRef] [PubMed]
  165. Wiemann, P.; Lechner, B.E.; Baccile, J.A.; Velk, T.A.; Yin, W.B.; Bok, J.W.; Pakala, S.; Losada, L.; Nierman, W.C.; Schroeder, F.C.; et al. Perturbations in small molecule synthesis uncovers an iron-responsive secondary metabolite network in Aspergillus fumigatus. Front. Microbiol. 2014, 5, 530. [Google Scholar] [CrossRef] [PubMed]
  166. Raffa, N.; Won, T.H.; Sukowaty, A.; Candor, K.; Cui, C.; Halder, S.; Dai, M.; Landero-Figueroa, J.A.; Schroeder, F.C.; Keller, N.P. Dual-purpose isocyanides produced by Aspergillus fumigatus contribute to cellular copper sufficiency and exhibit antimicrobial activity. Proc. Natl. Acad. Sci. USA 2021, 118, e2015224118. [Google Scholar] [CrossRef]
  167. Gressler, M.; Zaehle, C.; Scherlach, K.; Hertweck, C.; Brock, M. Multifactorial induction of an orphan PKS-NRPS gene cluster in Aspergillus terreus. Chem. Biol. 2011, 18, 198–209. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Ruiz-Villafán, B.; Cruz-Bautista, R.; Manzo-Ruiz, M.; Passari, A.K.; Villarreal-Gómez, K.; Rodríguez-Sanoja, R.; Sánchez, S. Carbon catabolite regulation of secondary metabolite formation, an old but not well-established regulatory system. Microb. Biotechnol. 2022, 15, 1058–1072. [Google Scholar] [CrossRef] [PubMed]
  169. Eisendle, M.; Schrettl, M.; Kragl, C.; Müller, D.; Illmer, P.; Haas, H. The intracellular siderophore ferricrocin is involved in iron storage, oxidative-stress resistance, germination, and sexual development in Aspergillus nidulans. Eukaryot. Cell 2006, 5, 1596–1603. [Google Scholar] [CrossRef]
  170. Hong, S.Y.; Roze, L.V.; Linz, J.E. Oxidative stress-related transcription factors in the regulation of secondary metabolism. Toxins 2013, 5, 683–702. [Google Scholar] [CrossRef] [PubMed]
  171. 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]
  172. Zhang, F.; Xu, G.; Geng, L.; Lu, X.; Yang, K.; Yuan, J.; Nie, X.; Zhuang, Z.; Wang, S. The Stress Response Regulator AflSkn7 Influences Morphological Development, Stress Response, and Pathogenicity in the Fungus Aspergillus flavus. Toxins 2016, 8, 202. [Google Scholar] [CrossRef]
  173. 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] [PubMed]
  174. Reverberi, M.; Gazzetti, K.; Punelli, F.; Scarpari, M.; Zjalic, S.; Ricelli, A.; Fabbri, A.A.; Fanelli, C. Aoyap1 regulates OTA synthesis by controlling cell redox balance in Aspergillus ochraceus. Appl. Microbiol. Biotechnol. 2012, 95, 1293–1304. [Google Scholar] [CrossRef]
  175. Oakley, C.E.; Ahuja, M.; Sun, W.W.; Entwistle, R.; Akashi, T.; Yaegashi, J.; Guo, C.J.; Cerqueira, G.C.; Russo Wortman, J.; Wang, C.C.; et al. Discovery of McrA, a master regulator of Aspergillus secondary metabolism. Mol. Microbiol. 2017, 103, 347–365. [Google Scholar] [CrossRef]
  176. Niehaus, E.-M.; Schumacher, J.; Burkhardt, I.; Rabe, P.; Münsterkötter, M.; Güldener, U.; Sieber, C.M.K.; Dickschat, J.S.; Tudzynski, B. The GATA-Type Transcription Factor Csm1 Regulates Conidiation and Secondary Metabolism in Fusarium fujikuroi. Front. Microbiol. 2017, 8, 1175. [Google Scholar] [CrossRef]
  177. Kim, H.Y.; Heo, D.Y.; Park, H.M.; Singh, D.; Lee, C.H. Metabolomic and Transcriptomic Comparison of Solid-State and Submerged Fermentation of Penicillium expansum KACC 40815. PLoS ONE 2016, 11, e0149012. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Son, S.Y.; Lee, S.; Singh, D.; Lee, N.R.; Lee, D.Y.; Lee, C.H. Comprehensive Secondary Metabolite Profiling Toward Delineating the Solid and Submerged-State Fermentation of Aspergillus oryzae KCCM 12698. Front. Microbiol. 2018, 9, 1076. [Google Scholar] [CrossRef] [PubMed]
  179. Thammajaruk, N.; Sriubolmas, N.; Israngkul, D.; Meevootisom, V.; Wiyakrutta, S. Optimization of culture conditions for mycoepoxydiene production by Phomopsis sp. Hant25. J. Ind. Microbiol. Biotechnol. 2011, 38, 679–685. [Google Scholar] [CrossRef]
  180. Bigelis, R.; He, H.; Yang, H.Y.; Chang, L.-P.; Greenstein, M. Production of fungal antibiotics using polymeric solid supports in solid-state and liquid fermentation. J. Ind. Microbiol. Biotechnol. 2006, 33, 815–826. [Google Scholar] [CrossRef] [PubMed]
  181. Nahlik, K.; Dumkow, M.; Bayram, O.; Helmstaedt, K.; Busch, S.; Valerius, O.; Gerke, J.; Hoppert, M.; Schwier, E.; Opitz, L.; et al. The COP9 signalosome mediates transcriptional and metabolic response to hormones, oxidative stress protection and cell wall rearrangement during fungal development. Mol. Microbiol. 2010, 78, 964–979. [Google Scholar] [CrossRef] [PubMed]
  182. Gerke, J.; Bayram, O.; Feussner, K.; Landesfeind, M.; Shelest, E.; Feussner, I.; Braus, G.H. Breaking the silence: Protein stabilization uncovers silenced biosynthetic gene clusters in the fungus Aspergillus nidulans. Appl. Environ. Microbiol. 2012, 78, 8234–8244. [Google Scholar] [CrossRef] [PubMed]
  183. Szewczyk, E.; Chiang, Y.M.; Oakley, C.E.; Davidson, A.D.; Wang, C.C.; Oakley, B.R. Identification and characterization of the asperthecin gene cluster of Aspergillus nidulans. Appl. Environ. Microbiol. 2008, 74, 7607–7612. [Google Scholar] [CrossRef]
  184. Berger, H.; Bacher, M.; Labuda, R.; Eppel, I.M.; Bayer, F.; Sulyok, M.; Gasparotto, E.; Zehetbauer, F.; Doppler, M.; Gratzl, H.; et al. Polaramycin B, and not physical interaction, is the signal that rewires fungal metabolism in the Streptomyces–Aspergillus interaction. Environ. Microbiol. 2022. [Google Scholar] [CrossRef]
  185. Gu, Y.; Ding, P.; Liang, Z.; Song, Y.; Liu, Y.; Chen, G.; Li, J.L. Activated production of silent metabolites from marine-derived fungus Penicillium citrinum. Fitoterapia 2018, 127, 207–211. [Google Scholar] [CrossRef]
  186. Gila, B.C.; Antal, K.; Birko, Z.; Keseru, J.S.; Pocsi, I.; Emri, T. Strategies Shaping the Transcription of Carbohydrate-Active Enzyme Genes in Aspergillus nidulans. J. Fungi 2022, 8, 79. [Google Scholar] [CrossRef]
  187. Li, T.; Jiang, G.; Qu, H.; Wang, Y.; Xiong, Y.; Jian, Q.; Wu, Y.; Duan, X.; Zhu, X.; Hu, W.; et al. Comparative Transcriptome Analysis of Penicillium citrinum Cultured with Different Carbon Sources Identifies Genes Involved in Citrinin Biosynthesis. Toxins 2017, 9, 69. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Lotfy, M.M.; Sayed, A.M.; AboulMagd, A.M.; Hassan, H.M.; El Amir, D.; Abouzid, S.F.; El-Gendy, A.O.; Rateb, M.E.; Abdelmohsen, U.R.; Alhadrami, H.; et al. Metabolomic profiling, biological evaluation of Aspergillus awamori, the river Nile-derived fungus using epigenetic and OSMAC approaches. RSC Adv. 2021, 11, 6709–6719. [Google Scholar] [CrossRef] [PubMed]
  189. Bok, J.W.; Hoffmeister, D.; Maggio-Hall, L.A.; Murillo, R.; Glasner, J.D.; Keller, N.P. Genomic mining for Aspergillus natural products. Chem. Biol. 2006, 13, 31–37. [Google Scholar] [CrossRef]
  190. Fountain, J.C.; Bajaj, P.; Pandey, M.; Nayak, S.N.; Yang, L.; Kumar, V.; Jayale, A.S.; Chitikineni, A.; Zhuang, W.; Scully, B.T.; et al. Oxidative stress and carbon metabolism influence Aspergillus flavus transcriptome composition and secondary metabolite production. Sci. Rep. 2016, 6, 38747. [Google Scholar] [CrossRef]
  191. Joffe, A.Z.; Lisker, N. Effects of light, temperature, and pH value on aflatoxin production in vitro. Appl. Microbiol. 1969, 18, 517–518. [Google Scholar] [CrossRef]
  192. Hagiwara, D.; Sakai, K.; Suzuki, S.; Umemura, M.; Nogawa, T.; Kato, N.; Osada, H.; Watanabe, A.; Kawamoto, S.; Gonoi, T.; et al. Temperature during conidiation affects stress tolerance, pigmentation, and trypacidin accumulation in the conidia of the airborne pathogen Aspergillus fumigatus. PLoS ONE 2017, 12, e0177050. [Google Scholar] [CrossRef] [PubMed]
  193. Berthier, E.; Lim, F.Y.; Deng, Q.; Guo, C.J.; Kontoyiannis, D.P.; Wang, C.C.; Rindy, J.; Beebe, D.J.; Huttenlocher, A.; Keller, N.P. Low-volume toolbox for the discovery of immunosuppressive fungal secondary metabolites. PLoS Pathog 2013, 9, e1003289. [Google Scholar] [CrossRef]
  194. Jyoti, S.; Singh, D.P. Production of Secondary Metabolites from Two Penicillium Strains Adapted to Different Temperature Conditions: A Study on DifferentialResponse of Fungal Strains to Temperature Stress. Cell. Mol. Biol. 2016. [Google Scholar] [CrossRef]
  195. Yaegashi, J.; Praseuth, M.B.; Tyan, S.W.; Sanchez, J.F.; Entwistle, R.; Chiang, Y.M.; Oakley, B.R.; Wang, C.C. Molecular genetic characterization of the biosynthesis cluster of a prenylated isoindolinone alkaloid aspernidine A in Aspergillus nidulans. Org. Lett. 2013, 15, 2862–2865. [Google Scholar] [CrossRef]
  196. Michielse, C.B.; Pfannmuller, A.; Macios, M.; Rengers, P.; Dzikowska, A.; Tudzynski, B. The interplay between the GATA transcription factors AreA, the global nitrogen regulator and AreB in Fusarium fujikuroi. Mol. Microbiol. 2014, 91, 472–493. [Google Scholar] [CrossRef] [PubMed]
  197. Niehaus, E.M.; von Bargen, K.W.; Espino, J.J.; Pfannmuller, A.; Humpf, H.U.; Tudzynski, B. Characterization of the fusaric acid gene cluster in Fusarium fujikuroi. Appl. Microbiol. Biotechnol. 2014, 98, 1749–1762. [Google Scholar] [CrossRef] [PubMed]
  198. Kim, H.; Son, H.; Lee, Y.-W. Effects of light on secondary metabolism and fungal development of Fusarium graminearum. J. Appl. Microbiol. 2014, 116, 380–389. [Google Scholar] [CrossRef] [PubMed]
  199. Tudzynski, B.; Homann, V.; Feng, B.; Marzluf, G.A. Isolation, characterization and disruption of the areA nitrogen regulatory gene of Gibberella fujikuroi. Mol. Genet. Genom. 1999, 261, 106–114. [Google Scholar] [CrossRef] [PubMed]
  200. Mihlan, M.; Homann, V.; Liu, T.W.; Tudzynski, B. AREA directly mediates nitrogen regulation of gibberellin biosynthesis in Gibberella fujikuroi, but its activity is not affected by NMR. Mol. Microbiol. 2003, 47, 975–991. [Google Scholar] [CrossRef]
  201. Nazari, L.; Manstretta, V.; Rossi, V. A non-linear model for temperature-dependent sporulation and T-2 and HT-2 production of Fusarium langsethiae and Fusarium sporotrichioides. Fungal Biol. 2016, 120, 562–571. [Google Scholar] [CrossRef]
  202. Sharma, V.K.; Kumar, J.; Singh, D.K.; Mishra, A.; Verma, S.K.; Gond, S.K.; Kumar, A.; Singh, N.; Kharwar, R.N. Induction of Cryptic and Bioactive Metabolites through Natural Dietary Components in an Endophytic Fungus Colletotrichum gloeosporioides (Penz.) Sacc. Front. Microbiol. 2017, 8, 1126. [Google Scholar] [CrossRef]
  203. Lin, Z.; Zhu, T.; Wei, H.; Zhang, G.; Wang, H.; Gu, Q. Spicochalasin A and New Aspochalasins from the Marine-Derived Fungus Spicaria elegans. Eur. J. Org. Chem. 2009, 2009, 3045–3051. [Google Scholar] [CrossRef]
  204. Smetanina, O.F.; Yurchenko, A.N.; Ivanets, E.V.; Kalinovsky, A.I.; Khudyakova, Y.V.; Dyshlovoy, S.A.; von Amsberg, G.; Yurchenko, E.A.; Afiyatullov, S.S. Unique prostate cancer-toxic polyketides from marine sediment-derived fungus Isaria felina. J. Antibiot. 2017, 70, 856–858. [Google Scholar] [CrossRef]
  205. Pruss, S.; Fetzner, R.; Seither, K.; Herr, A.; Pfeiffer, E.; Metzler, M.; Lawrence, C.B.; Fischer, R. Role of the Alternaria alternata blue-light receptor LreA (white-collar 1) in spore formation and secondary metabolism. Appl. Environ. Microbiol. 2014, 80, 2582–2591. [Google Scholar] [CrossRef]
  206. Bazafkan, H.; Beier, S.; Stappler, E.; Böhmdorfer, S.; Oberlerchner, J.T.; Sulyok, M.; Schmoll, M. SUB1 has photoreceptor dependent and independent functions in sexual development and secondary metabolism in Trichoderma reesei. Mol. Microbiol. 2017, 106, 742–759. [Google Scholar] [CrossRef] [PubMed]
  207. Shang, Z.; Li, X.-M.; Li, C.-S.; Wang, B.-G. Diverse Secondary Metabolites Produced by Marine-Derived Fungus Nigrospora sp. MA75 on Various Culture Media. Chem. Biodivers. 2012, 9, 1338–1348. [Google Scholar] [CrossRef]
  208. Siless, G.E.; Gallardo, G.L.; Rodriguez, M.A.; Rincon, Y.A.; Godeas, A.M.; Cabrera, G.M. Metabolites from the Dark Septate Endophyte Drechslera sp. Evaluation by LC/MS and Principal Component Analysis of Culture Extracts with Histone Deacetylase Inhibitors. Chem. Biodivers. 2018, 15, e1800133. [Google Scholar] [CrossRef] [PubMed]
  209. Khan, A.A.; Bacha, N.; Ahmad, B.; Bakht, J.; Lutfullah, G.; Ali, J. Synthesis of secondary metabolites by Cladosporium resinae (NRL-6437) under different growth media and chemical inducers and their pharmaceutical activity. Pak. J. Pharm. Sci. 2017, 30, 1617–1624. [Google Scholar] [PubMed]
  210. Morishita, Y.; Okazaki, Y.; Luo, Y.Y.; Nunoki, J.; Taniguchi, T.; Oshima, Y.; Asai, T. Use of plant hormones to activate silent polyketide biosynthetic pathways in Arthrinium sacchari, a fungus isolated from a spider. Org. Biomol. Chem. 2019, 17, 780–784. [Google Scholar] [CrossRef]
  211. Netzker, T.; Fischer, J.; Weber, J.; Mattern, D.J.; König, C.C.; Valiante, V.; Schroeckh, V.; Brakhage, A.A. Microbial communication leading to the activation of silent fungal secondary metabolite gene clusters. Front. Microbiol. 2015, 6, 299. [Google Scholar] [CrossRef]
  212. Tan, Z.Q.; Leow, H.Y.; Lee, D.C.W.; Karisnan, K.; Song, A.A.L.; Mai, C.W.; Yap, W.S.; Lim, S.H.E.; Lai, K.S. Co-Culture Systems for the Production of Secondary Metabolites: Current and Future Prospects. Open Biotechnol. J. 2019, 13, 18–26. [Google Scholar] [CrossRef]
  213. Bertrand, S.; Schumpp, O.; Bohni, N.; Bujard, A.; Azzollini, A.; Monod, M.; Gindro, K.; Wolfender, J.-L. Detection of metabolite induction in fungal co-cultures on solid media by high-throughput differential ultra-high pressure liquid chromatography–time-of-flight mass spectrometry fingerprinting. J. Chromatogr. A 2013, 1292, 219–228. [Google Scholar] [CrossRef]
  214. Chen, Y.; Wang, J.; Yang, N.; Wen, Z.; Sun, X.; Chai, Y.; Ma, Z. Wheat microbiome bacteria can reduce virulence of a plant pathogenic fungus by altering histone acetylation. Nat. Commun. 2018, 9, 3429. [Google Scholar] [CrossRef]
  215. Schroeckh, V.; Scherlach, K.; Nützmann, H.W.; Shelest, E.; Schmidt-Heck, W.; Schuemann, J.; Martin, K.; Hertweck, C.; Brakhage, A.A. Intimate bacterial-fungal interaction triggers biosynthesis of archetypal polyketides in Aspergillus nidulans. Proc. Natl. Acad. Sci. USA 2009, 106, 14558–14563. [Google Scholar] [CrossRef]
  216. Cueto, M.; Jensen, P.R.; Kauffman, C.; Fenical, W.; Lobkovsky, E.; Clardy, J. Pestalone, a New Antibiotic Produced by a Marine Fungus in Response to Bacterial Challenge. J. Nat. Prod. 2001, 64, 1444–1446. [Google Scholar] [CrossRef]
  217. Xu, C.; Sun, X.; Jin, M.; Zhang, X. A Novel Benzoquinone Compound Isolated from Deep-Sea Hydrothermal Vent Triggers Apoptosis of Tumor Cells. Mar. Drugs 2017, 15, 200. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Mandelare, P.E.; Adpressa, D.A.; Kaweesa, E.N.; Zakharov, L.N.; Loesgen, S. Coculture of Two Developmental Stages of a Marine-Derived Aspergillus alliaceus Results in the Production of the Cytotoxic Bianthrone Allianthrone A. J. Nat. Prod. 2018, 81, 1014–1022. [Google Scholar] [CrossRef] [PubMed]
  219. Yu, L.; Ding, W.; Ma, Z. Induced production of cytochalasans in co-culture of marine fungus Aspergillus flavipes and actinomycete Streptomyces sp. Nat. Prod. Res. 2016, 30, 1718–1723. [Google Scholar] [CrossRef] [PubMed]
  220. Wakefield, J.; Hassan, H.M.; Jaspars, M.; Ebel, R.; Rateb, M.E. Dual Induction of New Microbial Secondary Metabolites by Fungal Bacterial Co-cultivation. Front. Microbiol. 2017, 8, 1284. [Google Scholar] [CrossRef]
  221. Stroe, M.C.; Netzker, T.; Scherlach, K.; Kruger, T.; Hertweck, C.; Valiante, V.; Brakhage, A.A. Targeted induction of a silent fungal gene cluster encoding the bacteria-specific germination inhibitor fumigermin. Elife 2020, 9, e52541. [Google Scholar] [CrossRef]
  222. Gerke, J.; Köhler, A.M.; Wennrich, J.-P.; Große, V.; Shao, L.; Heinrich, A.K.; Bode, H.B.; Chen, W.; Surup, F.; Braus, G.H. Biosynthesis of Antibacterial Iron-Chelating Tropolones in Aspergillus nidulans as Response to Glycopeptide-Producing Streptomycetes. Front. Fungal Biol. 2022, 2, 777474. [Google Scholar] [CrossRef]
  223. Abdelwahab, M.F.; Kurtán, T.; Mándi, A.; Müller, W.E.G.; Fouad, M.A.; Kamel, M.S.; Liu, Z.; Ebrahim, W.; Daletos, G.; Proksch, P. Induced secondary metabolites from the endophytic fungus Aspergillus versicolor through bacterial co-culture and OSMAC approaches. Tetrahedron Lett. 2018, 59, 2647–2652. [Google Scholar] [CrossRef]
  224. Moree, W.J.; Yang, J.Y.; Zhao, X.; Liu, W.-T.; Aparicio, M.; Atencio, L.; Ballesteros, J.; Sánchez, J.; Gavilán, R.G.; Gutiérrez, M.; et al. Imaging Mass Spectrometry of a Coral Microbe Interaction with Fungi. J. Chem. Ecol. 2013, 39, 1045–1054. [Google Scholar] [CrossRef]
  225. Oh, D.-C.; Kauffman, C.A.; Jensen, P.R.; Fenical, W. Induced Production of Emericellamides A and B from the Marine-Derived Fungus Emericella sp. in Competing Co-culture. J. Nat. Prod. 2007, 70, 515–520. [Google Scholar] [CrossRef]
  226. Shang, Z.; Salim, A.A.; Capon, R.J. Chaunopyran A: Co-Cultivation of Marine Mollusk-Derived Fungi Activates a Rare Class of 2-Alkenyl-Tetrahydropyran. J. Nat. Prod. 2017, 80, 1167–1172. [Google Scholar] [CrossRef] [PubMed]
  227. Yao, L.; Zhu, L.-P.; Xu, X.-Y.; Tan, L.-L.; Sadilek, M.; Fan, H.; Hu, B.; Shen, X.-T.; Yang, J.; Qiao, B.; et al. Discovery of novel xylosides in co-culture of basidiomycetes Trametes versicolor and Ganoderma applanatum by integrated metabolomics and bioinformatics. Sci. Rep. 2016, 6, 33237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Oh, D.-C.; Jensen, P.R.; Kauffman, C.A.; Fenical, W. Libertellenones A–D: Induction of cytotoxic diterpenoid biosynthesis by marine microbial competition. Bioorganic Med. Chem. 2005, 13, 5267–5273. [Google Scholar] [CrossRef] [PubMed]
  229. Wang, Y.; Wang, L.; Zhuang, Y.; Kong, F.; Zhang, C.; Zhu, W. Phenolic Polyketides from the Co-Cultivation of Marine-Derived Penicillium sp. WC-29-5 and Streptomyces fradiae 007. Mar. Drugs 2014, 12, 2079–2088. [Google Scholar] [CrossRef] [PubMed]
  230. Yu, L.; Ding, W.; Wang, Q.; Ma, Z.; Xu, X.; Zhao, X.; Chen, Z. Induction of cryptic bioactive 2,5-diketopiperazines in fungus Penicillium sp. DT-F29 by microbial co-culture. Tetrahedron 2017, 73, 907–914. [Google Scholar] [CrossRef]
  231. Chen, H.; Aktas, N.; Konuklugil, B.; Mándi, A.; Daletos, G.; Lin, W.; Dai, H.; Kurtán, T.; Proksch, P. A new fusarielin analogue from Penicillium sp. isolated from the Mediterranean sponge Ircinia oros. Tetrahedron Lett. 2015, 56, 5317–5320. [Google Scholar] [CrossRef]
  232. Zhang, Z.; He, X.; Zhang, G.; Che, Q.; Zhu, T.; Gu, Q.; Li, D. Inducing Secondary Metabolite Production by Combined Culture of Talaromyces aculeatus and Penicillium variabile. J. Nat. Prod. 2017, 80, 3167–3171. [Google Scholar] [CrossRef]
  233. Khalil, Z.G.; Cruz-Morales, P.; Licona-Cassani, C.; Marcellin, E.; Capon, R.J. Inter-Kingdom beach warfare: Microbial chemical communication activates natural chemical defences. ISME J. 2019, 13, 147–158. [Google Scholar] [CrossRef]
  234. Sulyok, M.; Stadler, D.; Steiner, D.; Krska, R. Validation of an LC-MS/MS-based dilute-and-shoot approach for the quantification of > 500 mycotoxins and other secondary metabolites in food crops: Challenges and solutions. Anal. Bioanal. Chem. 2020, 412, 2607–2620. [Google Scholar] [CrossRef]
  235. Skinnider, M.A.; Johnston, C.W.; Gunabalasingam, M.; Merwin, N.J.; Kieliszek, A.M.; MacLellan, R.J.; Li, H.; Ranieri, M.R.M.; Webster, A.L.H.; Cao, M.P.T.; et al. Comprehensive prediction of secondary metabolite structure and biological activity from microbial genome sequences. Nat. Commun. 2020, 11, 6058. [Google Scholar] [CrossRef]
  236. 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] [PubMed]
  237. Bergmann, S.; Schümann, J.; Scherlach, K.; Lange, C.; Brakhage, A.A.; Hertweck, C. Genomics-driven discovery of PKS-NRPS hybrid metabolites from Aspergillus nidulans. Nat. Chem. Biol. 2007, 3, 213–217. [Google Scholar] [CrossRef]
  238. Chiang, Y.M.; Szewczyk, E.; Davidson, A.D.; Keller, N.; Oakley, B.R.; Wang, C.C. A gene cluster containing two fungal polyketide synthases encodes the biosynthetic pathway for a polyketide, asperfuranone, in Aspergillus nidulans. J. Am. Chem. Soc. 2009, 131, 2965–2970. [Google Scholar] [CrossRef] [PubMed]
  239. Janevska, S.; Arndt, B.; Baumann, L.; Apken, L.H.; Mauriz Marques, L.M.; Humpf, H.U.; Tudzynski, B. Establishment of the Inducible Tet-On System for the Activation of the Silent Trichosetin Gene Cluster in Fusarium fujikuroi. Toxins 2017, 9, 126. [Google Scholar] [CrossRef]
  240. Inglis, D.O.; Binkley, J.; Skrzypek, M.S.; Arnaud, M.B.; Cerqueira, G.C.; Shah, P.; Wymore, F.; Wortman, J.R.; Sherlock, G. Comprehensive annotation of secondary metabolite biosynthetic genes and gene clusters of Aspergillus nidulans, A. fumigatus, A. niger and A. oryzae. BMC Microbiol. 2013, 13, 91. [Google Scholar] [CrossRef] [PubMed]
  241. Chang, P.-K.; Ehrlich, K.C. Genome-wide analysis of the Zn(II)2Cys6 zinc cluster-encoding gene family in Aspergillus flavus. Appl. Microbiol. Biotechnol. 2013, 97, 4289–4300. [Google Scholar] [CrossRef]
  242. Kwon, M.J.; Steiniger, C.; Cairns, T.C.; Wisecaver, J.H.; Lind, A.L.; Pohl, C.; Regner, C.; Rokas, A.; Meyer, V.; Goldman, G.H. Beyond the Biosynthetic Gene Cluster Paradigm: Genome-Wide Coexpression Networks Connect Clustered and Unclustered Transcription Factors to Secondary Metabolic Pathways. Microbiol. Spectr. 2021, 9, e00898-21. [Google Scholar] [CrossRef]
  243. Ahuja, M.; Chiang, Y.-M.; Chang, S.-L.; Praseuth, M.B.; Entwistle, R.; Sanchez, J.F.; Lo, H.-C.; Yeh, H.-H.; Oakley, B.R.; Wang, C.C.C. Illuminating the Diversity of Aromatic Polyketide Synthases in Aspergillus nidulans. J. Am. Chem. Soc. 2012, 134, 8212–8221. [Google Scholar] [CrossRef]
  244. 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] [CrossRef]
  245. Zehetbauer, F.; Seidl, A.; Berger, H.; Sulyok, M.; Kastner, F.; Strauss, J. RimO (SrrB) is required for carbon starvation signaling and production of secondary metabolites in Aspergillus nidulans. Fungal Genet. Biol. 2022, 162, 103726. [Google Scholar] [CrossRef]
  246. Grau, M.F.; Entwistle, R.; Chiang, Y.-M.; Ahuja, M.; Oakley, C.E.; Akashi, T.; Wang, C.C.C.; Todd, R.B.; Oakley, B.R. Hybrid Transcription Factor Engineering Activates the Silent Secondary Metabolite Gene Cluster for (+)-Asperlin in Aspergillus nidulans. ACS Chem. Biol. 2018, 13, 3193–3205. [Google Scholar] [CrossRef] [PubMed]
  247. Sung, C.T.; Chang, S.-L.; Entwistle, R.; Ahn, G.; Lin, T.-S.; Petrova, V.; Yeh, H.-H.; Praseuth, M.B.; Chiang, Y.-M.; Oakley, B.R.; et al. Overexpression of a three-gene conidial pigment biosynthetic pathway in Aspergillus nidulans reveals the first NRPS known to acetylate tryptophan. Fungal Genet. Biol. 2017, 101, 1–6. [Google Scholar] [CrossRef] [PubMed]
  248. Yeh, H.-H.; Ahuja, M.; Chiang, Y.-M.; Oakley, C.E.; Moore, S.; Yoon, O.; Hajovsky, H.; Bok, J.-W.; Keller, N.P.; Wang, C.C.C.; et al. Resistance Gene-Guided Genome Mining: Serial Promoter Exchanges in Aspergillus nidulans Reveal the Biosynthetic Pathway for Fellutamide B, a Proteasome Inhibitor. ACS Chem. Biol. 2016, 11, 2275–2284. [Google Scholar] [CrossRef] [PubMed]
  249. Baccile, J.A.; Spraker, J.E.; Le, H.H.; Brandenburger, E.; Gomez, C.; Bok, J.W.; Macheleidt, J.; Brakhage, A.A.; Hoffmeister, D.; Keller, N.P.; et al. Plant-like biosynthesis of isoquinoline alkaloids in Aspergillus fumigatus. Nat. Chem. Biol. 2016, 12, 419–424. [Google Scholar] [CrossRef]
  250. Lim, F.Y.; Won, T.H.; Raffa, N.; Baccile, J.A.; Wisecaver, J.; Rokas, A.; Schroeder, F.C.; Keller, N.P.; Goldman, G.H. Fungal Isocyanide Synthases and Xanthocillin Biosynthesis in Aspergillus fumigatus. mBio 2018, 9, e00785-18. [Google Scholar] [CrossRef]
  251. Schüller, A.; Wolansky, L.; Berger, H.; Studt, L.; Gacek-Matthews, A.; Sulyok, M.; Strauss, J. A novel fungal gene regulation system based on inducible VPR-dCas9 and nucleosome map-guided sgRNA positioning. Appl. Microbiol. Biotechnol. 2020, 104, 9801–9822. [Google Scholar] [CrossRef]
  252. Li, X.; Huang, L.; Pan, L.; Wang, B.; Pan, L. CRISPR/dCas9-mediated epigenetic modification reveals differential regulation of histone acetylation on Aspergillus niger secondary metabolite. Microbiol. Res. 2021, 245, 126694. [Google Scholar] [CrossRef]
  253. Roux, I.; Woodcraft, C.; Hu, J.; Wolters, R.; Gilchrist, C.L.M.; Chooi, Y.-H. CRISPR-Mediated Activation of Biosynthetic Gene Clusters for Bioactive Molecule Discovery in Filamentous Fungi. ACS Synth. Biol. 2020, 9, 1843–1854. [Google Scholar] [CrossRef]
  254. Bromann, K.; Toivari, M.; Viljanen, K.; Vuoristo, A.; Ruohonen, L.; Nakari-Setälä, T. Identification and Characterization of a Novel Diterpene Gene Cluster in Aspergillus nidulans. PLoS ONE 2012, 7, e35450. [Google Scholar] [CrossRef]
  255. Yeh, H.-H.; Chiang, Y.-M.; Entwistle, R.; Ahuja, M.; Lee, K.-H.; Bruno, K.S.; Wu, T.-K.; Oakley, B.R.; Wang, C.C.C. Molecular genetic analysis reveals that a nonribosomal peptide synthetase-like (NRPS-like) gene in Aspergillus nidulans is responsible for microperfuranone biosynthesis. Appl. Microbiol. Biotechnol. 2012, 96, 739–748. [Google Scholar] [CrossRef]
  256. Ishikawa, N.; Tanaka, H.; Koyama, F.; Noguchi, H.; Wang, C.C.C.; Hotta, K.; Watanabe, K. Non-Heme Dioxygenase Catalyzes Atypical Oxidations of 6,7-Bicyclic Systems To Form the 6,6-Quinolone Core of Viridicatin-Type Fungal Alkaloids. Angew. Chem. Int. Ed. 2014, 53, 12880–12884. [Google Scholar] [CrossRef] [PubMed]
  257. Lin, T.S.; Chen, B.; Chiang, Y.M.; Wang, C.C.C. Discovery and Elucidation of the Biosynthesis of Aspernidgulenes: Novel Polyenes from Aspergillus nidulans by using Serial Promoter Replacement. Chembiochem 2019, 20, 329–334. [Google Scholar] [CrossRef] [PubMed]
  258. Zabala, A.O.; Xu, W.; Chooi, Y.-H.; Tang, Y. Characterization of a Silent Azaphilone Gene Cluster from Aspergillus niger ATCC 1015 Reveals a Hydroxylation-Mediated Pyran-Ring Formation. Chem. Biol. 2012, 19, 1049–1059. [Google Scholar] [CrossRef] [PubMed]
  259. 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]
  260. Tang, S.; Men, P.; Zhang, W.; Li, H.; Li, Z.; Huang, X.; Lu, X. Identification of a polyketide biosynthesis gene cluster by transcriptional regulator activation in Aspergillus terreus. Fungal Genet. Biol. 2022, 160, 103690. [Google Scholar] [CrossRef]
  261. Dong, Y.; Cui, C.B.; Li, C.W.; Hua, W.; Wu, C.J.; Zhu, T.J.; Gu, Q.Q. Activation of dormant secondary metabolite production by introducing neomycin resistance into the deep-sea fungus, Aspergillus versicolor ZBY-3. Mar. Drugs 2014, 12, 4326–4352. [Google Scholar] [CrossRef]
  262. Niehaus, E.M.; Janevska, S.; von Bargen, K.W.; Sieber, C.M.; Harrer, H.; Humpf, H.U.; Tudzynski, B. Apicidin F: Characterization and genetic manipulation of a new secondary metabolite gene cluster in the rice pathogen Fusarium fujikuroi. PLoS ONE 2014, 9, e103336. [Google Scholar] [CrossRef]
  263. Von Bargen, K.W.; Niehaus, E.M.; Krug, I.; Bergander, K.; Wurthwein, E.U.; Tudzynski, B.; Humpf, H.U. Isolation and Structure Elucidation of Fujikurins A-D: Products of the PKS19 Gene Cluster in Fusarium fujikuroi. J. Nat. Prod. 2015, 78, 1809–1815. [Google Scholar] [CrossRef]
  264. Arndt, B.; Studt, L.; Wiemann, P.; Osmanov, H.; Kleigrewe, K.; Köhler, J.; Krug, I.; Tudzynski, B.; Humpf, H.-U. Genetic engineering, high resolution mass spectrometry and nuclear magnetic resonance spectroscopy elucidate the bikaverin biosynthetic pathway in Fusarium fujikuroi. Fungal Genet. Biol. 2015, 84, 26–36. [Google Scholar] [CrossRef]
  265. Arndt, B.; Janevska, S.; Schmid, R.; Hubner, F.; Tudzynski, B.; Humpf, H.U. A Fungal N-Dimethylallyltryptophan Metabolite from Fusarium fujikuroi. Chembiochem 2017, 18, 899–904. [Google Scholar] [CrossRef]
  266. Sorensen, J.L.; Hansen, F.T.; Sondergaard, T.E.; Staerk, D.; Lee, T.V.; Wimmer, R.; Klitgaard, L.G.; Purup, S.; Giese, H.; Frandsen, R.J. Production of novel fusarielins by ectopic activation of the polyketide synthase 9 cluster in Fusarium graminearum. Environ. Microbiol. 2012, 14, 1159–1170. [Google Scholar] [CrossRef] [PubMed]
  267. Jia, L.J.; Tang, H.Y.; Wang, W.Q.; Yuan, T.L.; Wei, W.Q.; Pang, B.; Gong, X.M.; Wang, S.F.; Li, Y.J.; Zhang, D.; et al. A linear nonribosomal octapeptide from Fusarium graminearum facilitates cell-to-cell invasion of wheat. Nat. Commun. 2019, 10, 922. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Kakule, T.B.; Sardar, D.; Lin, Z.; Schmidt, E.W. Two related pyrrolidinedione synthetase loci in Fusarium heterosporum ATCC 74349 produce divergent metabolites. ACS Chem. Biol. 2013, 8, 1549–1557. [Google Scholar] [CrossRef] [PubMed]
  269. Wang, W.; Drott, M.; Greco, C.; Luciano-Rosario, D.; Wang, P.; Keller, N.P. Transcription Factor Repurposing Offers Insights into Evolution of Biosynthetic Gene Cluster Regulation. mBio 2021, 12, e0139921. [Google Scholar] [CrossRef]
  270. Wang, N.; Cui, C.-B.; Li, C.-W. A new cyclic dipeptide penicimutide: The activated production of cyclic dipeptides by introduction of neomycin-resistance in the marine-derived fungus Penicillium purpurogenum G59. Arch. Pharmacal Res. 2016, 39, 762–770. [Google Scholar] [CrossRef]
  271. Mózsik, L.; Hoekzema, M.; de Kok, N.A.W.; Bovenberg, R.A.L.; Nygård, Y.; Driessen, A.J.M. CRISPR-based transcriptional activation tool for silent genes in filamentous fungi. Sci. Rep. 2021, 11, 1118. [Google Scholar] [CrossRef]
  272. Zhou, H.; Zhan, J.; Watanabe, K.; Xie, X.; Tang, Y. A polyketide macrolactone synthase from the filamentous fungus Gibberella zeae. Proc. Natl. Acad. Sci. USA 2008, 105, 6249–6254. [Google Scholar] [CrossRef]
  273. Haynes, S.W.; Ames, B.D.; Gao, X.; Tang, Y.; Walsh, C.T. Unraveling Terminal C-Domain-Mediated Condensation in Fungal Biosynthesis of Imidazoindolone Metabolites. Biochemistry 2011, 50, 5668–5679. [Google Scholar] [CrossRef]
  274. Meng, X.; Fang, Y.; Ding, M.; Zhang, Y.; Jia, K.; Li, Z.; Collemare, J.; Liu, W. Developing fungal heterologous expression platforms to explore and improve the production of natural products from fungal biodiversity. Biotechnol. Adv. 2022, 54, 107866. [Google Scholar] [CrossRef]
  275. Vassaux, A.; Meunier, L.; Vandenbol, M.; Baurain, D.; Fickers, P.; Jacques, P.; Leclère, V. Nonribosomal peptides in fungal cell factories: From genome mining to optimized heterologous production. Biotechnol. Adv. 2019, 37, 107449. [Google Scholar] [CrossRef]
  276. Liu, C.; Minami, A.; Ozaki, T.; Wu, J.; Kawagishi, H.; Maruyama, J.-i.; Oikawa, H. Efficient Reconstitution of Basidiomycota Diterpene Erinacine Gene Cluster in Ascomycota Host Aspergillus oryzae Based on Genomic DNA Sequences. J. Am. Chem. Soc. 2019, 141, 15519–15523. [Google Scholar] [CrossRef] [PubMed]
  277. Okorafor, I.C.; Chen, M.; Tang, Y. High-Titer Production of Olivetolic Acid and Analogs in Engineered Fungal Host using a Nonplant Biosynthetic Pathway. ACS Synth. Biol. 2021, 10, 2159–2166. [Google Scholar] [CrossRef] [PubMed]
  278. Boecker, S.; Grätz, S.; Kerwat, D.; Adam, L.; Schirmer, D.; Richter, L.; Schütze, T.; Petras, D.; Süssmuth, R.D.; Meyer, V. Aspergillus niger is a superior expression host for the production of bioactive fungal cyclodepsipeptides. Fungal Biol. Biotechnol. 2018, 5, 4. [Google Scholar] [CrossRef]
  279. Nielsen, M.R.; Sørensen, J.L. Investigating Fungal Biosynthetic Pathways using Heterologous Gene Expression: Fusarium sp. as a Heterologous Host. In Engineering Natural Product Biosynthesis: Methods and Protocols; Skellam, E., Ed.; Springer: New York, NY, USA, 2022; pp. 53–74. [Google Scholar]
  280. Nielsen, M.R.; Wollenberg, R.D.; Westphal, K.R.; Sondergaard, T.E.; Wimmer, R.; Gardiner, D.M.; Sørensen, J.L. Heterologous expression of intact biosynthetic gene clusters in Fusarium graminearum. Fungal Genet. Biol. 2019, 132, 103248. [Google Scholar] [CrossRef]
  281. Kindinger, F.; Nies, J.; Becker, A.; Zhu, T.; Li, S.-M. Genomic Locus of a Penicillium crustosum Pigment as an Integration Site for Secondary Metabolite Gene Expression. ACS Chem. Biol. 2019, 14, 1227–1234. [Google Scholar] [CrossRef] [PubMed]
  282. Pohl, C.; Polli, F.; Schütze, T.; Viggiano, A.; Mózsik, L.; Jung, S.; de Vries, M.; Bovenberg, R.A.L.; Meyer, V.; Driessen, A.J.M. A Penicillium rubens platform strain for secondary metabolite production. Sci. Rep. 2020, 10, 7630. [Google Scholar] [CrossRef]
  283. Bok, J.W.; Ye, R.; Clevenger, K.D.; Mead, D.; Wagner, M.; Krerowicz, A.; Albright, J.C.; Goering, A.W.; Thomas, P.M.; Kelleher, N.L.; et al. Fungal artificial chromosomes for mining of the fungal secondary metabolome. BMC Genom. 2015, 16, 343. [Google Scholar] [CrossRef]
  284. Clevenger, K.D.; Bok, J.W.; Ye, R.; Miley, G.P.; Verdan, M.H.; Velk, T.; Chen, C.; Yang, K.; Robey, M.T.; Gao, P.; et al. A scalable platform to identify fungal secondary metabolites and their gene clusters. Nat. Chem. Biol. 2017, 13, 895–901. [Google Scholar] [CrossRef]
  285. Harvey, C.J.B.; Tang, M.; Schlecht, U.; Horecka, J.; Fischer, C.R.; Lin, H.C.; Li, J.; Naughton, B.; Cherry, J.; Miranda, M.; et al. HEx: A heterologous expression platform for the discovery of fungal natural products. Sci. Adv. 2018, 4, eaar5459. [Google Scholar] [CrossRef]
  286. Wiemann, P.; Soukup, A.A.; Folz, J.S.; Wang, P.M.; Noack, A.; Keller, N.P. CoIN: Co-inducible nitrate expression system for secondary metabolites in Aspergillus nidulans. Fungal Biol. Biotechnol. 2018, 5, 6. [Google Scholar] [CrossRef]
  287. Robey, M.T.; Ye, R.; Bok, J.W.; Clevenger, K.D.; Islam, M.N.; Chen, C.; Gupta, R.; Swyers, M.; Wu, E.; Gao, P.; et al. Identification of the First Diketomorpholine Biosynthetic Pathway using FAC-MS Technology. ACS Chem. Biol. 2018, 13, 1142–1147. [Google Scholar] [CrossRef] [PubMed]
  288. Maiya, S.; Grundmann, A.; Li, S.-M.; Turner, G. The Fumitremorgin Gene Cluster of Aspergillus fumigatus: Identification of a Gene Encoding Brevianamide F Synthetase. ChemBioChem 2006, 7, 1062–1069. [Google Scholar] [CrossRef] [PubMed]
  289. Itoh, T.; Tokunaga, K.; Matsuda, Y.; Fujii, I.; Abe, I.; Ebizuka, Y.; Kushiro, T. Reconstitution of a fungal meroterpenoid biosynthesis reveals the involvement of a novel family of terpene cyclases. Nat. Chem. 2010, 2, 858–864. [Google Scholar] [CrossRef] [PubMed]
  290. Watanabe, A.; Fujii, I.; Sankawa, U.; Mayorga, M.E.; Timberlake, W.E.; Ebizuka, Y. Re-identification of Aspergillus nidulans wA gene to code for a polyketide synthase of naphthopyrone. Tetrahedron Lett. 1999, 40, 91–94. [Google Scholar] [CrossRef]
  291. Caesar, L.K.; Robey, M.T.; Swyers, M.; Islam, M.N.; Ye, R.; Vagadia, P.P.; Schiltz, G.E.; Thomas, P.M.; Wu, C.C.; Kelleher, N.L.; et al. Heterologous Expression of the Unusual Terreazepine Biosynthetic Gene Cluster Reveals a Promising Approach for Identifying New Chemical Scaffolds. mBio 2020, 11. [Google Scholar] [CrossRef]
  292. Yuan, Y.; Cheng, S.; Bian, G.; Yan, P.; Ma, Z.; Dai, W.; Chen, R.; Fu, S.; Huang, H.; Chi, H.; et al. Efficient exploration of terpenoid biosynthetic gene clusters in filamentous fungi. Nat. Catal. 2022, 5, 277–287. [Google Scholar] [CrossRef]
  293. Bian, G.; Hou, A.; Yuan, Y.; Hu, B.; Cheng, S.; Ye, Z.; Di, Y.; Deng, Z.; Liu, T. Metabolic Engineering-Based Rapid Characterization of a Sesquiterpene Cyclase and the Skeletons of Fusariumdiene and Fusagramineol from Fusarium graminearum. Org. Lett. 2018, 20, 1626–1629. [Google Scholar] [CrossRef]
  294. Pedersen, T.B.; Nielsen, M.R.; Kristensen, S.B.; Spedtsberg, E.M.L.; Yasmine, W.; Matthiesen, R.; Kaniki, S.E.K.; Sørensen, T.; Petersen, C.; Muff, J.; et al. Heterologous Expression of the Core Genes in the Complex Fusarubin Gene Cluster of Fusarium Solani. Int. J. Mol. Sci. 2020, 21, 7601. [Google Scholar] [CrossRef]
  295. Tan, D.; Jamieson, C.S.; Ohashi, M.; Tang, M.C.; Houk, K.N.; Tang, Y. Genome-Mined Diels-Alderase Catalyzes Formation of the cis-Octahydrodecalins of Varicidin A and B. J. Am. Chem. Soc. 2019, 141, 769–773. [Google Scholar] [CrossRef]
  296. Hoefgen, S.; Lin, J.; Fricke, J.; Stroe, M.C.; Mattern, D.J.; Kufs, J.E.; Hortschansky, P.; Brakhage, A.A.; Hoffmeister, D.; Valiante, V. Facile assembly and fluorescence-based screening method for heterologous expression of biosynthetic pathways in fungi. Metab. Eng. 2018, 48, 44–51. [Google Scholar] [CrossRef]
  297. Yin, W.B.; Chooi, Y.H.; Smith, A.R.; Cacho, R.A.; Hu, Y.; White, T.C.; Tang, Y. Discovery of cryptic polyketide metabolites from dermatophytes using heterologous expression in Aspergillus nidulans. ACS Synth. Biol. 2013, 2, 629–634. [Google Scholar] [CrossRef] [PubMed]
  298. Noll, M.; Kornberg, R.D. Action of micrococcal nuclease on chromatin and the location of histone H1. J. Mol. Biol. 1977, 109, 393–404. [Google Scholar] [CrossRef]
  299. Richmond, T.J.; Finch, J.T.; Rushton, B.; Rhodes, D.; Klug, A. Structure of the nucleosome core particle at 7 Å resolution. Nature 1984, 311, 532–537. [Google Scholar] [CrossRef] [PubMed]
  300. Freitag, M. 5 Fungal Chromatin and Its Role in Regulation of Gene Expression. In Fungal Genomics; Nowrousian, M., Ed.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 99–120. [Google Scholar]
  301. Buscaino, A. Chromatin-Mediated Regulation of Genome Plasticity in Human Fungal Pathogens. Genes 2019, 10, 855. [Google Scholar] [CrossRef]
  302. Mivelaz, M.; Cao, A.-M.; Kubik, S.; Zencir, S.; Hovius, R.; Boichenko, I.; Stachowicz, A.M.; Kurat, C.F.; Shore, D.; Fierz, B. Chromatin Fiber Invasion and Nucleosome Displacement by the Rap1 Transcription Factor. Mol. Cell 2020, 77, 488–500. [Google Scholar] [CrossRef]
  303. Zaret, K.S. Pioneer Transcription Factors Initiating Gene Network Changes. Annu. Rev. Genet. 2020, 54, 367–385. [Google Scholar] [CrossRef]
  304. Allis, C.D.; Berger, S.L.; Cote, J.; Dent, S.; Jenuwien, T.; Kouzarides, T.; Pillus, L.; Reinberg, D.; Shi, Y.; Shiekhattar, R.; et al. New Nomenclature for Chromatin-Modifying Enzymes. Cell 2007, 131, 633–636. [Google Scholar] [CrossRef]
  305. Reyes-Dominguez, Y.; Bok, J.W.; Berger, H.; Shwab, E.K.; Basheer, A.; Gallmetzer, A.; Scazzocchio, C.; Keller, N.; Strauss, J. Heterochromatic marks are associated with the repression of secondary metabolism clusters in Aspergillus nidulans. Mol. Microbiol. 2010, 76, 1376–1386. [Google Scholar] [CrossRef]
  306. Tamaru, H.; Zhang, X.; McMillen, D.; Singh, P.B.; Nakayama, J.-i.; Grewal, S.I.; Allis, C.D.; Cheng, X.; Selker, E.U. Trimethylated lysine 9 of histone H3 is a mark for DNA methylation in Neurospora crassa. Nat. Genet. 2003, 34, 75–79. [Google Scholar] [CrossRef]
  307. Chujo, T.; Scott, B. Histone H3K9 and H3K27 methylation regulates fungal alkaloid biosynthesis in a fungal endophyte–plant symbiosis. Mol. Microbiol. 2014, 92, 413–434. [Google Scholar] [CrossRef]
  308. Atanasoff-Kardjalieff, A.K.; Lünne, F.; Kalinina, S.; Strauss, J.; Humpf, H.-U.; Studt, L. Biosynthesis of Fusapyrone Depends on the H3K9 Methyltransferase, FmKmt1, in Fusarium mangiferae. Front. Fungal Biol. 2021, 2, 671796. [Google Scholar] [CrossRef]
  309. Zhang, X.; Noberini, R.; Bonaldi, T.; Collemare, J.; Seidl, M.F. The histone code of the fungal genus Aspergillus uncovered by evolutionary and proteomic analyses. bioRxiv 2022. [Google Scholar] [CrossRef]
  310. Elías-Villalobos, A.; Barrales, R.R.; Ibeas, J.I. Chromatin modification factors in plant pathogenic fungi: Insights from Ustilago maydis. Fungal Genet. Biol. 2019, 129, 52–64. [Google Scholar] [CrossRef] [PubMed]
  311. Lachner, M.; Sengupta, R.; Schotta, G.; Jenuwein, T. Trilogies of histone lysine methylation as epigenetic landmarks of the eukaryotic genome. Cold Spring Harb. Symp. Quant. Biol. 2004, 69, 209–218. [Google Scholar] [CrossRef]
  312. Connolly, L.R.; Smith, K.M.; Freitag, M. The Fusarium graminearum histone H3 K27 methyltransferase KMT6 regulates development and expression of secondary metabolite gene clusters. PLoS Genet 2013, 9, e1003916. [Google Scholar] [CrossRef]
  313. Jamieson, K.; Rountree, M.R.; Lewis, Z.A.; Stajich, J.E.; Selker, E.U. Regional control of histone H3 lysine 27 methylation in Neurospora. Proc. Natl. Acad. Sci. USA 2013, 110, 6027–6032. [Google Scholar] [CrossRef]
  314. Schwartz, Y.B.; Kahn, T.G.; Nix, D.A.; Li, X.-Y.; Bourgon, R.; Biggin, M.; Pirrotta, V. Genome-wide analysis of Polycomb targets in Drosophila melanogaster. Nat. Genet. 2006, 38, 700–705. [Google Scholar] [CrossRef]
  315. Studt, L.; Rösler, S.M.; Burkhardt, I.; Arndt, B.; Freitag, M.; Humpf, H.U.; Dickschat, J.S.; Tudzynski, B. Knock-down of the methyltransferase Kmt6 relieves H3K27me3 and results in induction of cryptic and otherwise silent secondary metabolite gene clusters in Fusarium fujikuroi. Environ. Microbiol. 2016, 18, 4037–4054. [Google Scholar] [CrossRef]
  316. Kramer, H.M.; Seidl, M.F.; Thomma, B.; Cook, D.E. Local Rather than Global H3K27me3 Dynamics Are Associated with Differential Gene Expression in Verticillium dahliae. mBio 2022, 13, e0356621. [Google Scholar] [CrossRef]
  317. Jamieson, K.; McNaught, K.J.; Ormsby, T.; Leggett, N.A.; Honda, S.; Selker, E.U. Telomere repeats induce domains of H3K27 methylation in Neurospora. eLife 2018, 7, e31216. [Google Scholar] [CrossRef]
  318. Zhang, W.; Huang, J.; Cook, D.E. Histone modification dynamics at H3K27 are associated with altered transcription of in planta induced genes in Magnaporthe oryzae. PLOS Genet. 2021, 17, e1009376. [Google Scholar] [CrossRef] [PubMed]
  319. Ngo, V.; Chen, Z.; Zhang, K.; Whitaker, J.W.; Wang, M.; Wang, W. Epigenomic analysis reveals DNA motifs regulating histone modifications in human and mouse. Proc. Natl. Acad. Sci. USA 2019, 116, 3668–3677. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Demers, D.H.; Knestrick, M.A.; Fleeman, R.; Tawfik, R.; Azhari, A.; Souza, A.; Vesely, B.; Netherton, M.; Gupta, R.; Colon, B.L.; et al. Exploitation of Mangrove Endophytic Fungi for Infectious Disease Drug Discovery. Mar. Drugs 2018, 16, 376. [Google Scholar] [CrossRef] [PubMed]
  321. Aldholmi, M.; Wilkinson, B.; Ganesan, A. Epigenetic modulation of secondary metabolite profiles in Aspergillus calidoustus and Aspergillus westerdijkiae through histone deacetylase (HDAC) inhibition by vorinostat. J. Antibiot. 2020, 73, 410–413. [Google Scholar] [CrossRef]
  322. Zutz, C.; Gacek, A.; Sulyok, M.; Wagner, M.; Strauss, J.; Rychli, K. Small Chemical Chromatin Effectors Alter Secondary Metabolite Production in Aspergillus clavatus. Toxins 2013, 5, 1723–1741. [Google Scholar] [CrossRef]
  323. Bok, J.W.; Chiang, Y.M.; Szewczyk, E.; Reyes-Dominguez, Y.; Davidson, A.D.; Sanchez, J.F.; Lo, H.C.; Watanabe, K.; Strauss, J.; Oakley, B.R.; et al. Chromatin-level regulation of biosynthetic gene clusters. Nat. Chem. Biol. 2009, 5, 462–464. [Google Scholar] [CrossRef]
  324. Pidroni, A.; Faber, B.; Brosch, G.; Bauer, I.; Graessle, S. A Class 1 Histone Deacetylase as Major Regulator of Secondary Metabolite Production in Aspergillus nidulans. Front. Microbiol. 2018, 9, 2212. [Google Scholar] [CrossRef]
  325. Henke, M.T.; Soukup, A.A.; Goering, A.W.; McClure, R.A.; Thomson, R.J.; Keller, N.P.; Kelleher, N.L. New Aspercryptins, Lipopeptide Natural Products, Revealed by HDAC Inhibition in Aspergillus nidulans. ACS Chem. Biol. 2016, 11, 2117–2123. [Google Scholar] [CrossRef]
  326. Albright, J.C.; Henke, M.T.; Soukup, A.A.; McClure, R.A.; Thomson, R.J.; Keller, N.P.; Kelleher, N.L. Large-Scale Metabolomics Reveals a Complex Response of Aspergillus nidulans to Epigenetic Perturbation. ACS Chem. Biol. 2015, 10, 1535–1541. [Google Scholar] [CrossRef]
  327. Henrikson, J.C.; Hoover, A.R.; Joyner, P.M.; Cichewicz, R.H. A chemical epigenetics approach for engineering the in situbiosynthesis of a cryptic natural product from Aspergillus niger. Org. Biomol. Chem. 2009, 7, 435–438. [Google Scholar] [CrossRef]
  328. Wang, B.; Li, X.; Yu, D.; Chen, X.; Tabudravu, J.; Deng, H.; Pan, L. Deletion of the epigenetic regulator GcnE in Aspergillus niger FGSC A1279 activates the production of multiple polyketide metabolites. Microbiol. Res. 2018, 217, 101–107. [Google Scholar] [CrossRef] [PubMed]
  329. Li, X.; Xia, Z.; Tang, J.; Wu, J.; Tong, J.; Li, M.; Ju, J.; Chen, H.; Wang, L. Identification and Biological Evaluation of Secondary Metabolites from Marine Derived Fungi-Aspergillus sp. SCSIOW3, Cultivated in the Presence of Epigenetic Modifying Agents. Molecules 2017, 22, 1302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  330. Triastuti, A.; Vansteelandt, M.; Barakat, F.; Trinel, M.; Jargeat, P.; Fabre, N.; Amasifuen Guerra, C.A.; Mejia, K.; Valentin, A.; Haddad, M. How Histone Deacetylase Inhibitors Alter the Secondary Metabolites of Botryosphaeria mamane, an Endophytic Fungus Isolated from Bixa orellana. Chem. Biodivers 2019, 16, e1800485. [Google Scholar] [CrossRef] [PubMed]
  331. Mao, X.-M.; Xu, W.; Li, D.; Yin, W.-B.; Chooi, Y.-H.; Li, Y.-Q.; Tang, Y.; Hu, Y. Epigenetic Genome Mining of an Endophytic Fungus Leads to the Pleiotropic Biosynthesis of Natural Products. Angew. Chem. Int. Ed. 2015, 54, 7592–7596. [Google Scholar] [CrossRef]
  332. Adpressa, D.A.; Stalheim, K.J.; Proteau, P.J.; Loesgen, S. Unexpected Biotransformation of the HDAC Inhibitor Vorinostat Yields Aniline-Containing Fungal Metabolites. ACS Chem. Biol. 2017, 12, 1842–1847. [Google Scholar] [CrossRef]
  333. Liu, S.L.; Zhou, L.; Chen, H.P.; Liu, J.K. Sesquiterpenes with diverse skeletons from histone deacetylase inhibitor modified cultures of the basidiomycete Cyathus stercoreus (Schwein.) de Toni HFG134. Phytochemistry 2022, 195, 113048. [Google Scholar] [CrossRef]
  334. Chettri, P.; Dupont, P.-Y.; Bradshaw, R.E. Chromatin-level regulation of the fragmented dothistromin gene cluster in the forest pathogen Dothistroma septosporum. Mol. Microbiol. 2018, 107, 508–522. [Google Scholar] [CrossRef]
  335. Niehaus, E.-M.; Studt, L.; von Bargen, K.W.; Kummer, W.; Humpf, H.-U.; Reuter, G.; Tudzynski, B. Sound of silence: The beauvericin cluster in Fusarium fujikuroi is controlled by cluster-specific and global regulators mediated by H3K27 modification. Environ. Microbiol. 2016, 18, 4282–4302. [Google Scholar] [CrossRef]
  336. Janevska, S.; Baumann, L.; Sieber, C.M.K.; Munsterkotter, M.; Ulrich, J.; Kamper, J.; Guldener, U.; Tudzynski, B. Elucidation of the Two H3K36me3 Histone Methyltransferases Set2 and Ash1 in Fusarium fujikuroi Unravels Their Different Chromosomal Targets and a Major Impact of Ash1 on Genome Stability. Genetics 2018, 208, 153–171. [Google Scholar] [CrossRef]
  337. Janevska, S.; Güldener, U.; Sulyok, M.; Tudzynski, B.; Studt, L. Set1 and Kdm5 are antagonists for H3K4 methylation and regulators of the major conidiation-specific transcription factor gene ABA1 in Fusarium fujikuroi. Environ. Microbiol. 2018, 20, 3343–3362. [Google Scholar] [CrossRef]
  338. Adpressa, D.A.; Connolly, L.R.; Konkel, Z.M.; Neuhaus, G.F.; Chang, X.L.; Pierce, B.R.; Smith, K.M.; Freitag, M.; Loesgen, S. A metabolomics-guided approach to discover Fusarium graminearum metabolites after removal of a repressive histone modification. Fungal Genet. Biol. 2019, 132, 103256. [Google Scholar] [CrossRef] [PubMed]
  339. Gu, Q.; Ji, T.; Sun, X.; Huang, H.; Zhang, H.; Lu, X.; Wu, L.; Huo, R.; Wu, H.; Gao, X. Histone H3 lysine 9 methyltransferase FvDim5 regulates fungal development, pathogenicity and osmotic stress responses in Fusarium verticillioides. FEMS Microbiol. Lett. 2017, 364. [Google Scholar] [CrossRef] [Green Version]
  340. Fan, A.; Mi, W.; Liu, Z.; Zeng, G.; Zhang, P.; Hu, Y.; Fang, W.; Yin, W.B. Deletion of a Histone Acetyltransferase Leads to the Pleiotropic Activation of Natural Products in Metarhizium robertsii. Org. Lett. 2017, 19, 1686–1689. [Google Scholar] [CrossRef] [PubMed]
  341. Qadri, M.; Nalli, Y.; Jain, S.K.; Chaubey, A.; Ali, A.; Strobel, G.A.; Vishwakarma, R.A.; Riyaz-Ul-Hassan, S. An Insight into the Secondary Metabolism of Muscodor yucatanensis: Small-Molecule Epigenetic Modifiers Induce Expression of Secondary Metabolism-Related Genes and Production of New Metabolites in the Endophyte. Microb. Ecol. 2017, 73, 954–965. [Google Scholar] [CrossRef]
  342. Ramesha, K.P.; Chandra Mohana, N.; Chandra Nayaka, S.; Satish, S. Epigenetic Modifiers Revamp Secondary Metabolite Production in Endophytic Nigrospora sphaerica. Front. Microbiol. 2021, 12, 730355. [Google Scholar] [CrossRef]
  343. El-Hawary, S.S.; Sayed, A.M.; Mohammed, R.; Hassan, H.M.; Zaki, M.A.; Rateb, M.E.; Mohammed, T.A.; Amin, E.; Abdelmohsen, U.R. Epigenetic Modifiers Induce Bioactive Phenolic Metabolites in the Marine-Derived Fungus Penicillium brevicompactum. Mar. Drugs 2018, 16, 253. [Google Scholar] [CrossRef] [PubMed]
  344. Guzman-Chavez, F.; Salo, O.; Samol, M.; Ries, M.; Kuipers, J.; Bovenberg, R.A.L.; Vreeken, R.J.; Driessen, A.J.M. Deregulation of secondary metabolism in a histone deacetylase mutant of Penicillium chrysogenum. MicrobiologyOpen 2018, 7, e00598. [Google Scholar] [CrossRef]
  345. Gubiani, J.R.; Wijeratne, E.M.K.; Shi, T.; Araujo, A.R.; Arnold, A.E.; Chapman, E.; Gunatilaka, A.A.L. An epigenetic modifier induces production of (10′S)-verruculide B, an inhibitor of protein tyrosine phosphatases by Phoma sp. nov. LG0217, a fungal endophyte of Parkinsonia microphylla. Bioorganic Med. Chem. 2017, 25, 1860–1866. [Google Scholar] [CrossRef]
  346. Shi, T.; Shao, C.L.; Liu, Y.; Zhao, D.L.; Cao, F.; Fu, X.M.; Yu, J.Y.; Wu, J.S.; Zhang, Z.K.; Wang, C.Y. Terpenoids from the Coral-Derived Fungus Trichoderma harzianum (XS-20090075) Induced by Chemical Epigenetic Manipulation. Front. Microbiol. 2020, 11, 572. [Google Scholar] [CrossRef]
  347. Jeon, J.; Kwon, S.; Lee, Y.-H. Histone acetylation in fungal pathogens of plants. Plant Pathol. J. 2014, 30, 1–9. [Google Scholar] [CrossRef]
  348. Gacek, A.; Strauss, J. The chromatin code of fungal secondary metabolite gene clusters. Appl. Microbiol. Biotechnol. 2012, 95, 1389–1404. [Google Scholar] [CrossRef] [PubMed]
  349. Lan, H.; Wu, L.; Sun, R.; Keller, N.P.; Yang, K.; Ye, L.; He, S.; Zhang, F.; Wang, S. The HosA Histone Deacetylase Regulates Aflatoxin Biosynthesis Through Direct Regulation of Aflatoxin Cluster Genes. Mol. Plant-Microbe Interact. 2019, 32, 1210–1228. [Google Scholar] [CrossRef] [PubMed]
  350. Shwab, E.K.; Bok, J.W.; Tribus, M.; Galehr, J.; Graessle, S.; Keller, N.P. Histone Deacetylase Activity Regulates Chemical Diversity in Aspergillus. Eukaryot. Cell 2007, 6, 1656–1664. [Google Scholar] [CrossRef] [PubMed]
  351. Price, R.J.; Weindling, E.; Berman, J.; Buscaino, A.; Pietro, A.D. Chromatin Profiling of the Repetitive and Nonrepetitive Genomes of the Human Fungal Pathogen Candida albicans. mBio 2019, 10, e01376-19. [Google Scholar] [CrossRef]
  352. Ho, T.C.S.; Chan, A.H.Y.; Ganesan, A. Thirty Years of HDAC Inhibitors: 2020 Insight and Hindsight. J. Med. Chem. 2020, 63, 12460–12484. [Google Scholar] [CrossRef]
  353. Freitag, M. Histone Methylation by SET Domain Proteins in Fungi. Annu. Rev. Microbiol. 2017, 71, 413–439. [Google Scholar] [CrossRef]
  354. Gacek-Matthews, A.; Berger, H.; Sasaki, T.; Wittstein, K.; Gruber, C.; Lewis, Z.A.; Strauss, J. KdmB, a Jumonji Histone H3 Demethylase, Regulates Genome-Wide H3K4 Trimethylation and Is Required for Normal Induction of Secondary Metabolism in Aspergillus nidulans. PLOS Genet. 2016, 12, e1006222. [Google Scholar] [CrossRef]
  355. Bachleitner, S.; Sørensen, J.L.; Gacek-Matthews, A.; Sulyok, M.; Studt, L.; Strauss, J. Evidence of a Demethylase-Independent Role for the H3K4-Specific Histone Demethylases in Aspergillus nidulans and Fusarium graminearum Secondary Metabolism. Front. Microbiol. 2019, 10, 1759. [Google Scholar] [CrossRef]
  356. Morgan, M.A.J.; Shilatifard, A. Reevaluating the roles of histone-modifying enzymes and their associated chromatin modifications in transcriptional regulation. Nat. Genet. 2020, 52, 1271–1281. [Google Scholar] [CrossRef]
  357. Giles, S.S.; Soukup, A.A.; Lauer, C.; Shaaban, M.; Lin, A.; Oakley, B.R.; Wang, C.C.C.; Keller, N.P. Cryptic Aspergillus nidulans Antimicrobials. Appl. Environ. Microbiol. 2011, 77, 3669–3675. [Google Scholar] [CrossRef]
  358. Krogan, N.J.; Dover, J.; Khorrami, S.; Greenblatt, J.F.; Schneider, J.; Johnston, M.; Shilatifard, A. COMPASS, a histone H3 (Lysine 4) methyltransferase required for telomeric silencing of gene expression. J. Biol. Chem. 2002, 277, 10753–10755. [Google Scholar] [CrossRef]
  359. Palmer, J.M.; Bok, J.W.; Lee, S.; Dagenais, T.R.T.; Andes, D.R.; Kontoyiannis, D.P.; Keller, N.P. Loss of CclA, required for histone 3 lysine 4 methylation, decreases growth but increases secondary metabolite production in Aspergillus fumigatus. PeerJ 2013, 1, e4. [Google Scholar] [CrossRef] [Green Version]
  360. Studt, L.; Janevska, S.; Arndt, B.; Boedi, S.; Sulyok, M.; Humpf, H.-U.; Tudzynski, B.; Strauss, J. Lack of the COMPASS Component Ccl1 Reduces H3K4 Trimethylation Levels and Affects Transcription of Secondary Metabolite Genes in Two Plant–Pathogenic Fusarium Species. Front. Microbiol. 2017, 7, 2144. [Google Scholar] [CrossRef] [PubMed]
  361. Soukup, A.A.; Chiang, Y.-M.; Bok, J.W.; Reyes-Dominguez, Y.; Oakley, B.R.; Wang, C.C.C.; Strauss, J.; Keller, N.P. Overexpression of the Aspergillus nidulans histone 4 acetyltransferase EsaA increases activation of secondary metabolite production. Mol. Microbiol. 2012, 86, 314–330. [Google Scholar] [CrossRef] [PubMed]
  362. Roze, L.V.; Arthur, A.E.; Hong, S.Y.; Chanda, A.; Linz, J.E. The initiation and pattern of spread of histone H4 acetylation parallel the order of transcriptional activation of genes in the aflatoxin cluster. Mol. Microbiol. 2007, 66, 713–726. [Google Scholar] [CrossRef]
  363. Niehaus, E.-M.; Kleigrewe, K.; Wiemann, P.; Studt, L.; Sieber, C.M.K.; Connolly, L.R.; Freitag, M.; Güldener, U.; Tudzynski, B.; Humpf, H.-U. Genetic Manipulation of the Fusarium fujikuroi Fusarin Gene Cluster Yields Insight into the Complex Regulation and Fusarin Biosynthetic Pathway. Chem. Biol. 2013, 20, 1055–1066. [Google Scholar] [CrossRef] [PubMed]
  364. Studt, L.; Schmidt, F.J.; Jahn, L.; Sieber, C.M.K.; Connolly, L.R.; Niehaus, E.-M.; Freitag, M.; Humpf, H.-U.; Tudzynski, B. Two Histone Deacetylases, FfHda1 and FfHda2, Are Important for Fusarium fujikuroi Secondary Metabolism and Virulence. Appl. Environ. Microbiol. 2013, 79, 7719–7734. [Google Scholar] [CrossRef]
  365. Kong, X.; van Diepeningen, A.D.; van der Lee, T.A.J.; Waalwijk, C.; Xu, J.; Xu, J.; Zhang, H.; Chen, W.; Feng, J. The Fusarium graminearum Histone Acetyltransferases Are Important for Morphogenesis, DON Biosynthesis, and Pathogenicity. Front. Microbiol. 2018, 9, 654. [Google Scholar] [CrossRef]
  366. Rösler, S.M.; Kramer, K.; Finkemeier, I.; Humpf, H.-U.; Tudzynski, B. The SAGA complex in the rice pathogen Fusarium fujikuroi: Structure and functional characterization. Mol. Microbiol. 2016, 102, 951–974. [Google Scholar] [CrossRef]
  367. Fischer, J.; Müller, S.Y.; Netzker, T.; Jäger, N.; Gacek-Matthews, A.; Scherlach, K.; Stroe, M.C.; García-Altares, M.; Pezzini, F.; Schoeler, H.; et al. Chromatin mapping identifies BasR, a key regulator of bacteria-triggered production of fungal secondary metabolites. eLife 2018, 7, e40969. [Google Scholar] [CrossRef]
  368. Nützmann, H.-W.; Reyes-Dominguez, Y.; Scherlach, K.; Schroeckh, V.; Horn, F.; Gacek, A.; Schümann, J.; Hertweck, C.; Strauss, J.; Brakhage, A.A. Bacteria-induced natural product formation in the fungus Aspergillus nidulans requires Saga/Ada-mediated histone acetylation. Proc. Natl. Acad. Sci. USA 2011, 108, 14282–14287. [Google Scholar] [CrossRef] [PubMed]
  369. Atanasoff-Kardjalieff, A.K.; Studt, L. Secondary Metabolite Gene Regulation in Mycotoxigenic Fusarium Species: A Focus on Chromatin. Toxins 2022, 14, 96. [Google Scholar] [CrossRef]
  370. Sun, R.; Wen, M.; Wu, L.; Lan, H.; Yuan, J.; Wang, S. The Fungi-specific histone Acetyltransferase Rtt109 mediates morphogenesis, Aflatoxin synthesis and pathogenicity in Aspergillus flavus by acetylating H3K9. IMA Fungus 2021, 12, 9. [Google Scholar] [CrossRef] [PubMed]
  371. Ferraro, A.R.; Ameri, A.J.; Lu, Z.; Kamei, M.; Schmitz, R.J.; Lewis, Z.A. Chromatin accessibility profiling in Neurospora crassa reveals molecular features associated with accessible and inaccessible chromatin. BMC Genom. 2021, 22, 459. [Google Scholar] [CrossRef] [PubMed]
  372. Zhao, W.; Xu, Y.; Wang, Y.; Gao, D.; King, J.; Xu, Y.; Liang, F.S. Investigating crosstalk between H3K27 acetylation and H3K4 trimethylation in CRISPR/dCas-based epigenome editing and gene activation. Sci. Rep. 2021, 11, 15912. [Google Scholar] [CrossRef]
  373. Nai, Y.-S.; Huang, Y.-C.; Yen, M.-R.; Chen, P.-Y. Diversity of Fungal DNA Methyltransferases and Their Association with DNA Methylation Patterns. Front. Microbiol. 2021, 11, 616922. [Google Scholar] [CrossRef]
  374. Pillay, L.C.; Nekati, L.; Makhwitine, P.J.; Ndlovu, S.I. Epigenetic Activation of Silent Biosynthetic Gene Clusters in Endophytic Fungi using Small Molecular Modifiers. Front. Microbiol. 2022, 13, 815008. [Google Scholar] [CrossRef]
  375. Basenko, E.Y.; Sasaki, T.; Ji, L.; Prybol, C.J.; Burckhardt, R.M.; Schmitz, R.J.; Lewis, Z.A. Genome-wide redistribution of H3K27me3 is linked to genotoxic stress and defective growth. Proc. Natl. Acad. Sci. USA 2015, 112, E6339–E6348. [Google Scholar] [CrossRef]
  376. Freitag, M.; Hickey, P.C.; Khlafallah, T.K.; Read, N.D.; Selker, E.U. HP1 is essential for DNA methylation in neurospora. Mol. Cell 2004, 13, 427–434. [Google Scholar] [CrossRef]
  377. Zhang, X.; Liu, X.; Zhao, Y.; Cheng, J.; Xie, J.; Fu, Y.; Jiang, D.; Chen, T. Histone H3 Lysine 9 Methyltransferase DIM5 Is Required for the Development and Virulence of Botrytis cinerea. Front. Microbiol. 2016, 7, 1289. [Google Scholar] [CrossRef]
  378. Djoumbou-Feunang, Y.; Fiamoncini, J.; Gil-de-la-Fuente, A.; Greiner, R.; Manach, C.; Wishart, D.S. BioTransformer: A comprehensive computational tool for small molecule metabolism prediction and metabolite identification. J. Cheminform. 2019, 11, 2. [Google Scholar] [CrossRef] [PubMed]
  379. Barrangou, R.; Fremaux, C.; Deveau, H.; Richards, M.; Boyaval, P.; Moineau, S.; Romero, D.A.; Horvath, P. CRISPR provides acquired resistance against viruses in prokaryotes. Science 2007, 315, 1709–1712. [Google Scholar] [CrossRef] [PubMed]
  380. Makarova, K.S.; Wolf, Y.I.; Iranzo, J.; Shmakov, S.A.; Alkhnbashi, O.S.; Brouns, S.J.J.; Charpentier, E.; Cheng, D.; Haft, D.H.; Horvath, P.; et al. Evolutionary classification of CRISPR–Cas systems: A burst of class 2 and derived variants. Nat. Rev. Microbiol. 2020, 18, 67–83. [Google Scholar] [CrossRef] [PubMed]
  381. Ouedraogo, J.-P.; Tsang, A. CRISPR_Cas systems for fungal research. Fungal Biol. Rev. 2020, 34, 189–201. [Google Scholar] [CrossRef]
  382. Jiang, C.; Lv, G.; Tu, Y.; Cheng, X.; Duan, Y.; Zeng, B.; He, B. Applications of CRISPR/Cas9 in the Synthesis of Secondary Metabolites in Filamentous Fungi. Front. Microbiol. 2021, 12, 638096. [Google Scholar] [CrossRef]
  383. Chavez, A.; Tuttle, M.; Pruitt, B.W.; Ewen-Campen, B.; Chari, R.; Ter-Ovanesyan, D.; Haque, S.J.; Cecchi, R.J.; Kowal, E.J.K.; Buchthal, J.; et al. Comparison of Cas9 activators in multiple species. Nat. Methods 2016, 13, 563–567. [Google Scholar] [CrossRef]
  384. Wensing, L.; Sharma, J.; Uthayakumar, D.; Proteau, Y.; Chavez, A.; Shapiro, R.S. A CRISPR Interference Platform for Efficient Genetic Repression in Candida albicans. mSphere 2019, 4. [Google Scholar] [CrossRef]
  385. Wensing, L.; Shapiro, R.S. Design and Generation of a CRISPR Interference System for Genetic Repression and Essential Gene Analysis in the Fungal Pathogen Candida albicans. Methods Mol. Biol. 2022, 2377, 69–88. [Google Scholar] [CrossRef]
  386. Schmidtmann, E.; Anton, T.; Rombaut, P.; Herzog, F.; Leonhardt, H. Determination of local chromatin composition by CasID. Nucleus 2016, 7, 476–484. [Google Scholar] [CrossRef]
  387. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A programmable dual-RNA-guided DNA endonuclease in adaptive bacterial immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef]
  388. Nødvig, C.S.; Nielsen, J.B.; Kogle, M.E.; Mortensen, U.H. A CRISPR-Cas9 System for Genetic Engineering of Filamentous Fungi. PLoS ONE 2015, 10, e0133085. [Google Scholar] [CrossRef] [PubMed]
  389. Nødvig, C.S.; Hoof, J.B.; Kogle, M.E.; Jarczynska, Z.D.; Lehmbeck, J.; Klitgaard, D.K.; Mortensen, U.H. Efficient oligo nucleotide mediated CRISPR-Cas9 gene editing in Aspergilli. Fungal Genet. Biol. 2018, 115, 78–89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  390. Verwaal, R.; Buiting-Wiessenhaan, N.; Dalhuijsen, S.; Roubos, J.A. CRISPR/Cpf1 enables fast and simple genome editing of Saccharomyces cerevisiae. Yeast 2018, 35, 201–211. [Google Scholar] [CrossRef]
  391. Fonfara, I.; Richter, H.; Bratovič, M.; Le Rhun, A.; Charpentier, E. The CRISPR-associated DNA-cleaving enzyme Cpf1 also processes precursor CRISPR RNA. Nature 2016, 532, 517–521. [Google Scholar] [CrossRef] [PubMed]
  392. Zetsche, B.; Heidenreich, M.; Mohanraju, P.; Fedorova, I.; Kneppers, J.; DeGennaro, E.M.; Winblad, N.; Choudhury, S.R.; Abudayyeh, O.O.; Gootenberg, J.S.; et al. Multiplex gene editing by CRISPR–Cpf1 using a single crRNA array. Nat. Biotechnol. 2017, 35, 31–34. [Google Scholar] [CrossRef]
  393. Bandyopadhyay, A.; Kancharla, N.; Javalkote, V.S.; Dasgupta, S.; Brutnell, T.P. CRISPR-Cas12a (Cpf1): A Versatile Tool in the Plant Genome Editing Tool Box for Agricultural Advancement. Front. Plant Sci. 2020, 11, 584151. [Google Scholar] [CrossRef]
  394. Zetsche, B.; Gootenberg, J.S.; Abudayyeh, O.O.; Slaymaker, I.M.; Makarova, K.S.; Essletzbichler, P.; Volz, S.E.; Joung, J.; van der Oost, J.; Regev, A.; et al. Cpf1 Is a Single RNA-Guided Endonuclease of a Class 2 CRISPR-Cas System. Cell 2015, 163, 759–771. [Google Scholar] [CrossRef]
  395. Song, R.; Zhai, Q.; Sun, L.; Huang, E.; Zhang, Y.; Zhu, Y.; Guo, Q.; Tian, Y.; Zhao, B.; Lu, H. CRISPR/Cas9 genome editing technology in filamentous fungi: Progress and perspective. Appl. Microbiol. Biotechnol. 2019, 103, 6919–6932. [Google Scholar] [CrossRef]
  396. Kim, H.K.; Song, M.; Lee, J.; Menon, A.V.; Jung, S.; Kang, Y.M.; Choi, J.W.; Woo, E.; Koh, H.C.; Nam, J.W.; et al. In vivo high-throughput profiling of CRISPR-Cpf1 activity. Nat. Methods 2017, 14, 153–159. [Google Scholar] [CrossRef]
  397. Jacobsen, T.; Ttofali, F.; Liao, C.; Manchalu, S.; Gray, B.N.; Beisel, C.L. Characterization of Cas12a nucleases reveals diverse PAM profiles between closely-related orthologs. Nucleic Acids Res. 2020, 48, 5624–5638. [Google Scholar] [CrossRef]
  398. Gasiunas, G.; Barrangou, R.; Horvath, P.; Siksnys, V. Cas9-crRNA ribonucleoprotein complex mediates specific DNA cleavage for adaptive immunity in bacteria. Proc. Natl. Acad. Sci. USA 2012, 109, E2579–E2586. [Google Scholar] [CrossRef] [PubMed]
  399. Komor, A.C.; Kim, Y.B.; Packer, M.S.; Zuris, J.A.; Liu, D.R. Programmable editing of a target base in genomic DNA without double-stranded DNA cleavage. Nature 2016, 533, 420–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  400. Yamano, T.; Nishimasu, H.; Zetsche, B.; Hirano, H.; Slaymaker, I.M.; Li, Y.; Fedorova, I.; Nakane, T.; Makarova, K.S.; Koonin, E.V.; et al. Crystal Structure of Cpf1 in Complex with Guide RNA and Target DNA. Cell 2016, 165, 949–962. [Google Scholar] [CrossRef] [PubMed]
  401. Cofsky, J.C.; Karandur, D.; Huang, C.J.; Witte, I.P.; Kuriyan, J.; Doudna, J.A. CRISPR-Cas12a exploits R-loop asymmetry to form double-strand breaks. eLife 2020, 9, e55143. [Google Scholar] [CrossRef]
  402. Qi, L.S.; Larson, M.H.; Gilbert, L.A.; Doudna, J.A.; Weissman, J.S.; Arkin, A.P.; Lim, W.A. Repurposing CRISPR as an RNA-Guided Platform for Sequence-Specific Control of Gene Expression. Cell 2013, 152, 1173–1183. [Google Scholar] [CrossRef]
  403. Smith, J.D.; Suresh, S.; Schlecht, U.; Wu, M.; Wagih, O.; Peltz, G.; Davis, R.W.; Steinmetz, L.M.; Parts, L.; St. Onge, R.P. Quantitative CRISPR interference screens in yeast identify chemical-genetic interactions and new rules for guide RNA design. Genome Biol. 2016, 17, 45. [Google Scholar] [CrossRef]
  404. Malzahn, A.A.; Tang, X.; Lee, K.; Ren, Q.; Sretenovic, S.; Zhang, Y.; Chen, H.; Kang, M.; Bao, Y.; Zheng, X.; et al. Application of CRISPR-Cas12a temperature sensitivity for improved genome editing in rice, maize, and Arabidopsis. BMC Biol. 2019, 17, 9. [Google Scholar] [CrossRef]
  405. Zhou, Y.; Liu, X.; Wu, J.; Zhao, G.; Wang, J. CRISPR-Cas12a-Assisted Genome Editing in Amycolatopsis mediterranei. Front. Bioeng. Biotechnol. 2020, 8, 698. [Google Scholar] [CrossRef]
  406. Liu, Q.; Zhang, Y.; Li, F.; Li, J.; Sun, W.; Tian, C. Upgrading of efficient and scalable CRISPR–Cas-mediated technology for genetic engineering in thermophilic fungus Myceliophthora thermophila. Biotechnol. Biofuels 2019, 12, 293. [Google Scholar] [CrossRef]
  407. Lightfoot, J.D.; Fuller, K.K. CRISPR/Cas9-Mediated Gene Replacement in the Fungal Keratitis Pathogen Fusarium solani var. petroliphilum. Microorganisms 2019, 7, 457. [Google Scholar] [CrossRef]
  408. Fuller, K.K.; Chen, S.; Loros, J.J.; Dunlap, J.C. Development of the CRISPR/Cas9 System for Targeted Gene Disruption in Aspergillus fumigatus. Eukaryot. Cell 2015, 14, 1073–1080. [Google Scholar] [CrossRef] [PubMed]
  409. Pachlinger, R.; Mitterbauer, R.; Adam, G.; Strauss, J. Metabolically independent and accurately adjustable Aspergillus sp. expression system. Appl. Environ. Microbiol. 2005, 71, 672–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  410. Weber, J.; Valiante, V.; Nødvig, C.S.; Mattern, D.J.; Slotkowski, R.A.; Mortensen, U.H.; Brakhage, A.A. Functional Reconstitution of a Fungal Natural Product Gene Cluster by Advanced Genome Editing. ACS Synth. Biol. 2017, 6, 62–68. [Google Scholar] [CrossRef] [PubMed]
  411. Huang, J.; Cook, D.E. CRISPR-Cas12a ribonucleoprotein-mediated gene editing in the plant pathogenic fungus Magnaporthe oryzae. STAR Protoc. 2022, 3, 101072. [Google Scholar] [CrossRef] [PubMed]
  412. Van Leeuwe, T.M.; Arentshorst, M.; Ernst, T.; Alazi, E.; Punt, P.J.; Ram, A.F.J. Efficient marker free CRISPR/Cas9 genome editing for functional analysis of gene families in filamentous fungi. Fungal Biol. Biotechnol. 2019, 6, 13. [Google Scholar] [CrossRef]
  413. Schuster, M.; Kahmann, R. CRISPR-Cas9 genome editing approaches in filamentous fungi and oomycetes. Fungal Genet. Biol. 2019, 130, 43–53. [Google Scholar] [CrossRef]
  414. Chavez, A.; Scheiman, J.; Vora, S.; Pruitt, B.W.; Tuttle, M.; Iye, E.P.R.; Lin, S.; Kiani, S.; Guzman, C.D.; Wiegand, D.J.; et al. Highly efficient Cas9-mediated transcriptional programming. Nat. Methods 2015, 12, 326–328. [Google Scholar] [CrossRef]
  415. Hall, D.B.; Struhl, K. The VP16 Activation Domain Interacts with Multiple Transcriptional Components as Determined by Protein-Protein Cross-linking in Vivo. J. Biol. Chem. 2002, 277, 46043–46050. [Google Scholar] [CrossRef]
  416. Hinz, J.M.; Laughery, M.F.; Wyrick, J.J. Nucleosomes Inhibit Cas9 Endonuclease Activity in Vitro. Biochemistry 2015, 54, 7063–7066. [Google Scholar] [CrossRef]
  417. Thakore, P.I.; D’Ippolito, A.M.; Song, L.; Safi, A.; Shivakumar, N.K.; Kabadi, A.M.; Reddy, T.E.; Crawford, G.E.; Gersbach, C.A. Highly specific epigenome editing by CRISPR-Cas9 repressors for silencing of distal regulatory elements. Nat. Methods 2015, 12, 1143–1149. [Google Scholar] [CrossRef]
  418. Klann, T.S.; Black, J.B.; Chellappan, M.; Safi, A.; Song, L.; Hilton, I.B.; Crawford, G.E.; Reddy, T.E.; Gersbach, C.A. CRISPR–Cas9 epigenome editing enables high-throughput screening for functional regulatory elements in the human genome. Nat. Biotechnol. 2017, 35, 561–568. [Google Scholar] [CrossRef]
  419. Li, J.; Mahata, B.; Escobar, M.; Goell, J.; Wang, K.; Khemka, P.; Hilton, I.B. Programmable human histone phosphorylation and gene activation using a CRISPR/Cas9-based chromatin kinase. Nat. Commun. 2021, 12, 896. [Google Scholar] [CrossRef] [PubMed]
  420. Lickwar, C.R.; Mueller, F.; Hanlon, S.E.; McNally, J.G.; Lieb, J.D. Genome-wide protein–DNA binding dynamics suggest a molecular clutch for transcription factor function. Nature 2012, 484, 251–255. [Google Scholar] [CrossRef] [PubMed]
  421. Feng, Y.; Liu, S.; Chen, R.; Xie, A. Target binding and residence: A new determinant of DNA double-strand break repair pathway choice in CRISPR/Cas9 genome editing. J. Zhejiang Univ.—Sci. B 2021, 22, 73–86. [Google Scholar] [CrossRef]
  422. Liu, G.; Zhang, Y.; Zhang, T. Computational approaches for effective CRISPR guide RNA design and evaluation. Comput. Struct. Biotechnol. J. 2020, 18, 35–44. [Google Scholar] [CrossRef]
  423. Xu, H.; Xiao, T.; Chen, C.H.; Li, W.; Meyer, C.A.; Wu, Q.; Wu, D.; Cong, L.; Zhang, F.; Liu, J.S.; et al. Sequence determinants of improved CRISPR sgRNA design. Genome Res. 2015, 25, 1147–1157. [Google Scholar] [CrossRef]
  424. Park, J.; Bae, S.; Kim, J.-S. Cas-Designer: A web-based tool for choice of CRISPR-Cas9 target sites. Bioinformatics 2015, 31, 4014–4016. [Google Scholar] [CrossRef]
  425. Mendoza, B.J.; Trinh, C.T. Enhanced guide-RNA design and targeting analysis for precise CRISPR genome editing of single and consortia of industrially relevant and non-model organisms. Bioinformatics 2017, 34, 16–23. [Google Scholar] [CrossRef]
  426. Li, Z.; Zhang, D.; Xiong, X.; Yan, B.; Xie, W.; Sheen, J.; Li, J.-F. A potent Cas9-derived gene activator for plant and mammalian cells. Nat. Plants 2017, 3, 930–936. [Google Scholar] [CrossRef]
  427. Hwang, G.-H.; Song, B.; Bae, S. Current widely-used web-based tools for CRISPR nucleases, base editors, and prime editors. Gene Genome Ed. 2021, 1, 100004. [Google Scholar] [CrossRef]
  428. Zhang, S.; Guo, F.; Yan, W.; Dai, Z.; Dong, W.; Zhou, J.; Zhang, W.; Xin, F.; Jiang, M. Recent Advances of CRISPR/Cas9-Based Genetic Engineering and Transcriptional Regulation in Industrial Biology. Front. Bioeng. Biotechnol. 2020, 7, 459. [Google Scholar] [CrossRef] [PubMed]
  429. Ma, H.; Tu, L.-C.; Naseri, A.; Huisman, M.; Zhang, S.; Grunwald, D.; Pederson, T. CRISPR-Cas9 nuclear dynamics and target recognition in living cells. J. Cell Biol. 2016, 214, 529–537. [Google Scholar] [CrossRef] [Green Version]
  430. Palermo, G.; Miao, Y.; Walker, R.C.; Jinek, M.; McCammon, J.A. CRISPR-Cas9 conformational activation as elucidated from enhanced molecular simulations. Proc. Natl. Acad. Sci. USA 2017, 114, 7260–7265. [Google Scholar] [CrossRef] [PubMed]
  431. Mekler, V.; Minakhin, L.; Semenova, E.; Kuznedelov, K.; Severinov, K. Kinetics of the CRISPR-Cas9 effector complex assembly and the role of 3′-terminal segment of guide RNA. Nucleic Acids Res. 2016, 44, 2837–2845. [Google Scholar] [CrossRef]
  432. Raper, A.T.; Stephenson, A.A.; Suo, Z. Functional Insights Revealed by the Kinetic Mechanism of CRISPR/Cas9. J. Am. Chem. Soc. 2018, 140, 2971–2984. [Google Scholar] [CrossRef]
  433. Strohkendl, I.; Saifuddin, F.A.; Rybarski, J.R.; Finkelstein, I.J.; Russell, R. Kinetic Basis for DNA Target Specificity of CRISPR-Cas12a. Mol. Cell 2018, 71, 816–824.e3. [Google Scholar] [CrossRef]
  434. Sundaresan, R.; Parameshwaran, H.P.; Yogesha, S.D.; Keilbarth, M.W.; Rajan, R. RNA-Independent DNA Cleavage Activities of Cas9 and Cas12a. Cell Rep. 2017, 21, 3728–3739. [Google Scholar] [CrossRef]
  435. Swarts, D.C.; van der Oost, J.; Jinek, M. Structural Basis for Guide RNA Processing and Seed-Dependent DNA Targeting by CRISPR-Cas12a. Mol. Cell 2017, 66, 221–233.e4. [Google Scholar] [CrossRef]
  436. Dong, D.; Ren, K.; Qiu, X.; Zheng, J.; Guo, M.; Guan, X.; Liu, H.; Li, N.; Zhang, B.; Yang, D.; et al. The crystal structure of Cpf1 in complex with CRISPR RNA. Nature 2016, 532, 522–526. [Google Scholar] [CrossRef]
  437. Mekler, V.; Kuznedelov, K.; Severinov, K. Quantification of the affinities of CRISPR-Cas9 nucleases for cognate protospacer adjacent motif (PAM) sequences. J. Biol. Chem. 2020, 295, 6509–6517. [Google Scholar] [CrossRef]
  438. Chung, C.H.; Allen, A.G.; Sullivan, N.T.; Atkins, A.; Nonnemacher, M.R.; Wigdahl, B.; Dampier, W. Computational Analysis Concerning the Impact of DNA Accessibility on CRISPR-Cas9 Cleavage Efficiency. Mol. Ther. 2020, 28, 19–28. [Google Scholar] [CrossRef]
  439. Daer, R.M.; Cutts, J.P.; Brafman, D.A.; Haynes, K.A. The Impact of Chromatin Dynamics on Cas9-Mediated Genome Editing in Human Cells. ACS Synth. Biol. 2017, 6, 428–438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  440. Verkuijl, S.A.N.; Rots, M.G. The influence of eukaryotic chromatin state on CRISPR–Cas9 editing efficiencies. Curr. Opin. Biotechnol. 2019, 55, 68–73. [Google Scholar] [CrossRef] [PubMed]
  441. Isaac, R.S.; Jiang, F.; Doudna, J.A.; Lim, W.A.; Narlikar, G.J.; Almeida, R. Nucleosome breathing and remodeling constrain CRISPR-Cas9 function. eLife 2016, 5, e13450. [Google Scholar] [CrossRef] [PubMed]
  442. Sternberg, S.H.; Redding, S.; Jinek, M.; Greene, E.C.; Doudna, J.A. DNA interrogation by the CRISPR RNA-guided endonuclease Cas9. Nature 2014, 507, 62–67. [Google Scholar] [CrossRef]
  443. Clarke, R.; Heler, R.; MacDougall, M.S.; Yeo, N.C.; Chavez, A.; Regan, M.; Hanakahi, L.; Church, G.M.; Marraffini, L.A.; Merrill, B.J. Enhanced Bacterial Immunity and Mammalian Genome Editing via RNA-Polymerase-Mediated Dislodging of Cas9 from Double-Strand DNA Breaks. Mol. Cell 2018, 71, 42–55.e8. [Google Scholar] [CrossRef]
  444. Zhang, S.; Zhang, Q.; Hou, X.-M.; Guo, L.; Wang, F.; Bi, L.; Zhang, X.; Li, H.-H.; Wen, F.; Xi, X.-G.; et al. Dynamics of Staphylococcus aureus Cas9 in DNA target Association and Dissociation. EMBO Rep. 2020, 21, e50184. [Google Scholar] [CrossRef]
  445. Jeon, Y.; Choi, Y.H.; Jang, Y.; Yu, J.; Goo, J.; Lee, G.; Jeong, Y.K.; Lee, S.H.; Kim, I.-S.; Kim, J.-S.; et al. Direct observation of DNA target searching and cleavage by CRISPR-Cas12a. Nat. Commun. 2018, 9, 2777. [Google Scholar] [CrossRef]
  446. Zhang, Q.; Wen, F.; Zhang, S.; Jin, J.; Bi, L.; Lu, Y.; Li, M.; Xi, X.-G.; Huang, X.; Shen, B.; et al. The post-PAM interaction of RNA-guided spCas9 with DNA dictates its target binding and dissociation. Sci. Adv. 2019, 5, eaaw9807. [Google Scholar] [CrossRef]
  447. Hsu, P.D.; Scott, D.A.; Weinstein, J.A.; Ran, F.A.; Konermann, S.; Agarwala, V.; Li, Y.; Fine, E.J.; Wu, X.; Shalem, O.; et al. DNA targeting specificity of RNA-guided Cas9 nucleases. Nat. Biotechnol. 2013, 31, 827–832. [Google Scholar] [CrossRef]
  448. Künne, T.; Swarts, D.C.; Brouns, S.J. Planting the seed: Target recognition of short guide RNAs. Trends Microbiol. 2014, 22, 74–83. [Google Scholar] [CrossRef] [PubMed]
  449. Wu, X.; Scott, D.A.; Kriz, A.J.; Chiu, A.C.; Hsu, P.D.; Dadon, D.B.; Cheng, A.W.; Trevino, A.E.; Konermann, S.; Chen, S.; et al. Genome-wide binding of the CRISPR endonuclease Cas9 in mammalian cells. Nat. Biotechnol. 2014, 32, 670–676. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  450. Walton, R.T.; Christie, K.A.; Whittaker, M.N.; Kleinstiver, B.P. Unconstrained genome targeting with near-PAMless engineered CRISPR-Cas9 variants. Science 2020, 368, 290–296. [Google Scholar] [CrossRef] [PubMed]
  451. Kleinstiver, B.P.; Prew, M.S.; Tsai, S.Q.; Topkar, V.V.; Nguyen, N.T.; Zheng, Z.; Gonzales, A.P.W.; Li, Z.; Peterson, R.T.; Yeh, J.-R.J.; et al. Engineered CRISPR-Cas9 nucleases with altered PAM specificities. Nature 2015, 523, 481–485. [Google Scholar] [CrossRef] [PubMed]
  452. Kleinstiver, B.P.; Prew, M.S.; Tsai, S.Q.; Nguyen, N.T.; Topkar, V.V.; Zheng, Z.; Joung, J.K. Broadening the targeting range of Staphylococcus aureus CRISPR-Cas9 by modifying PAM recognition. Nat. Biotechnol. 2015, 33, 1293–1298. [Google Scholar] [CrossRef]
  453. Tóth, E.; Varga, É.; Kulcsár, P.I.; Kocsis-Jutka, V.; Krausz, S.L.; Nyeste, A.; Welker, Z.; Huszár, K.; Ligeti, Z.; Tálas, A.; et al. Improved LbCas12a variants with altered PAM specificities further broaden the genome targeting range of Cas12a nucleases. Nucleic Acids Res. 2020, 48, 3722–3733. [Google Scholar] [CrossRef]
  454. Kleinstiver, B.P.; Sousa, A.A.; Walton, R.T.; Tak, Y.E.; Hsu, J.Y.; Clement, K.; Welch, M.M.; Horng, J.E.; Malagon-Lopez, J.; Scarfò, I.; et al. Engineered CRISPR–Cas12a variants with increased activities and improved targeting ranges for gene, epigenetic and base editing. Nat. Biotechnol. 2019, 37, 276–282. [Google Scholar] [CrossRef]
  455. Kleinstiver, B.P.; Pattanayak, V.; Prew, M.S.; Tsai, S.Q.; Nguyen, N.T.; Zheng, Z.; Joung, J.K. High-fidelity CRISPR–Cas9 nucleases with no detectable genome-wide off-target effects. Nature 2016, 529, 490–495. [Google Scholar] [CrossRef]
  456. Tsai, S.Q.; Zheng, Z.; Nguyen, N.T.; Liebers, M.; Topkar, V.V.; Thapar, V.; Wyvekens, N.; Khayter, C.; Iafrate, A.J.; Le, L.P.; et al. GUIDE-seq enables genome-wide profiling of off-target cleavage by CRISPR-Cas nucleases. Nat. Biotechnol. 2015, 33, 187–197. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Workflow for genome mining assisted BGC prediction and verification. The order and interplay of prediction tools, databases, and scientist is depicted. In orange letters, we point out pitfalls that could lead to incorrect predictions and can easily be overlooked. The correct order is depicted by the arrows and follows the numbered bold headlines. The light-yellow arrow represents a re-prediction of an outdated BGC annotation. Explanations regarding point 5. “Experimental verification”: The top part is the genomic locus with the putative BGC surrounded by a dotted outline; genes are numbered from 1–6, BB stands for “backbone- gene” (gene No. 3; orange) and TF stands for “pathway-specific transcription factor” (gene No. 4, green). Yellow lightning bolt symbols represent genetic manipulation (i.e., gene deletion in case of BB or overexpression in case of TF). The colors represent the three different samples/transformants: The control (“C”) represents the wildtype (wt) or recipient strain and is illustrated by a grey flask, expression bars, and HPLC spectra (i.e., wt-like synthesis of the investigated secondary metabolite. In this case it is produced but in case of a silent BGC, the peak on the far right would be absent). Transformant 1 (“T1”) has the backbone gene deleted and is represented by an orange flask, gene expression bars and HPLC peaks (orange flat line; i.e., no synthesis of the secondary metabolite); Transformant 2 (“T2”) has the TF of the BGC overexpressed. It is depicted by a green flask, green gene expression bars and a green HPLC peak (i.e., overproduction of the secondary metabolite). Diagrams have either the expression level (“Expr.”) or the absorption value (“Abs.”) as the y-axis and time (“min”) as the x-axis. The setup of 5. Experimental verification represents one possibility of linking a putative BGC to its metabolite. After deposition into a database (6.), the gathered information about the BGC can be utilized for the prediction of BGCs in other genomes (2.).
Figure 1. Workflow for genome mining assisted BGC prediction and verification. The order and interplay of prediction tools, databases, and scientist is depicted. In orange letters, we point out pitfalls that could lead to incorrect predictions and can easily be overlooked. The correct order is depicted by the arrows and follows the numbered bold headlines. The light-yellow arrow represents a re-prediction of an outdated BGC annotation. Explanations regarding point 5. “Experimental verification”: The top part is the genomic locus with the putative BGC surrounded by a dotted outline; genes are numbered from 1–6, BB stands for “backbone- gene” (gene No. 3; orange) and TF stands for “pathway-specific transcription factor” (gene No. 4, green). Yellow lightning bolt symbols represent genetic manipulation (i.e., gene deletion in case of BB or overexpression in case of TF). The colors represent the three different samples/transformants: The control (“C”) represents the wildtype (wt) or recipient strain and is illustrated by a grey flask, expression bars, and HPLC spectra (i.e., wt-like synthesis of the investigated secondary metabolite. In this case it is produced but in case of a silent BGC, the peak on the far right would be absent). Transformant 1 (“T1”) has the backbone gene deleted and is represented by an orange flask, gene expression bars and HPLC peaks (orange flat line; i.e., no synthesis of the secondary metabolite); Transformant 2 (“T2”) has the TF of the BGC overexpressed. It is depicted by a green flask, green gene expression bars and a green HPLC peak (i.e., overproduction of the secondary metabolite). Diagrams have either the expression level (“Expr.”) or the absorption value (“Abs.”) as the y-axis and time (“min”) as the x-axis. The setup of 5. Experimental verification represents one possibility of linking a putative BGC to its metabolite. After deposition into a database (6.), the gathered information about the BGC can be utilized for the prediction of BGCs in other genomes (2.).
Pharmaceutics 14 01837 g001
Figure 2. Molecular biology methods for BGC activation. Silent genes are depicted as grey or black arrows while active genes are green. Yellow thunder bolts represent genetic manipulation. Genetic manipulation that results in overexpression or deletion of a gene is abbreviated as “OE” or “KO” respectively. (a) The exchange of one (i.e., pathway-specific TF) or several promoter regions of the BGC of interest leads to gene activation. (b) Overexpression of a global regulator can lead to the activation of several BGCs all over the genome. (c) Heterologous expression of a BGC of interest in an expression host followed by cultivation, extraction and structure elucidation of the compound encoded within the BGC. (d) Genetic manipulation of a chromatin remodeler can lead to the de-repression or silencing of a BGC that is regulated by the respective remodeler. Silencing and expression-promoting histone modifications are depicted as grey and green pins respectively.
Figure 2. Molecular biology methods for BGC activation. Silent genes are depicted as grey or black arrows while active genes are green. Yellow thunder bolts represent genetic manipulation. Genetic manipulation that results in overexpression or deletion of a gene is abbreviated as “OE” or “KO” respectively. (a) The exchange of one (i.e., pathway-specific TF) or several promoter regions of the BGC of interest leads to gene activation. (b) Overexpression of a global regulator can lead to the activation of several BGCs all over the genome. (c) Heterologous expression of a BGC of interest in an expression host followed by cultivation, extraction and structure elucidation of the compound encoded within the BGC. (d) Genetic manipulation of a chromatin remodeler can lead to the de-repression or silencing of a BGC that is regulated by the respective remodeler. Silencing and expression-promoting histone modifications are depicted as grey and green pins respectively.
Pharmaceutics 14 01837 g002
Figure 3. Maturation of guide RNAs and Cas mediated Gene editing. (a) Methods for release of mature sgRNAs with Cas9. Top: the ~100 bp long sgRNA is flanked by the 37 bp Hammerhead (HH) and the 68 bp HDV ribozyme. Upon transcription by RNA polymerase II (RNA POL II), the sgRNA is released by the formation of secondary structure and autocleavage of the ribozymes. Bottom: Cas9 sgRNAs can be matured and released by flanking it by ~72 bp long tRNAs (i.e., glycine). Upon transcription by RNA polymerase III (RNA POL III), the tRNAs are cleaved by the endogenous RNAse P and RNAse Z and the mature sgRNA is released which then can form a functional complex with Cas9. Arrays of sgRNA and tRNAs can be cloned and expressed in vivo for multiplexing purposes. (b) Cas12a can process the CRISPR transcript itself. The expression fragment contains an only ~19 bp direct repeat followed by the approximately 25 bp long protospacer sequence. Due to its small size, it is easy to clone several guide RNAs in sequence for multiplexing purposes. After expression by RNA polymerase III (RNA POL III), the transcript is processed by Cas12a and the mature sgRNAs are released which can form a functional complex with Cas12a. (c) two options for gene editing with CRISPR/Cas technology. After the successful introduction of the double-strand break, the DNA can either be repaired via the non-homologous end joining pathway (NHEJ) or, if an oligo or gene with complementary regions (orange) is provided (i.e., HDR Substrate), the homology directed repair mechanism can be activated (HDR) and the provided genetic material is introduced into the target locus.
Figure 3. Maturation of guide RNAs and Cas mediated Gene editing. (a) Methods for release of mature sgRNAs with Cas9. Top: the ~100 bp long sgRNA is flanked by the 37 bp Hammerhead (HH) and the 68 bp HDV ribozyme. Upon transcription by RNA polymerase II (RNA POL II), the sgRNA is released by the formation of secondary structure and autocleavage of the ribozymes. Bottom: Cas9 sgRNAs can be matured and released by flanking it by ~72 bp long tRNAs (i.e., glycine). Upon transcription by RNA polymerase III (RNA POL III), the tRNAs are cleaved by the endogenous RNAse P and RNAse Z and the mature sgRNA is released which then can form a functional complex with Cas9. Arrays of sgRNA and tRNAs can be cloned and expressed in vivo for multiplexing purposes. (b) Cas12a can process the CRISPR transcript itself. The expression fragment contains an only ~19 bp direct repeat followed by the approximately 25 bp long protospacer sequence. Due to its small size, it is easy to clone several guide RNAs in sequence for multiplexing purposes. After expression by RNA polymerase III (RNA POL III), the transcript is processed by Cas12a and the mature sgRNAs are released which can form a functional complex with Cas12a. (c) two options for gene editing with CRISPR/Cas technology. After the successful introduction of the double-strand break, the DNA can either be repaired via the non-homologous end joining pathway (NHEJ) or, if an oligo or gene with complementary regions (orange) is provided (i.e., HDR Substrate), the homology directed repair mechanism can be activated (HDR) and the provided genetic material is introduced into the target locus.
Pharmaceutics 14 01837 g003
Figure 4. Targeted gene activation approaches via CRISPR/Cas technology. (a) targeted chromatin remodeling via a chromatin active domain (CAD). DNA is depicted in dotted black lines, nucleosomes are cyan, heterochromatic histone modifications are grey pins while euchromatic histone modifications are green pins. A fusion protein of dCas and a CAD is targeted into a nucleosome-free region (NFR) in a heterochromatic area where it modifies the local histones. This leads to an increase in DNA accessibility and promotes transcription. (b) Targeted gene activation by dCas fused to a trans-activating domain (TAD; e.g., VPR). DNA is depicted as a black bar, genes are either grey (inactive) or green (active) arrows. The colored bars in front of the genes are target sites for dCas-TAD (i.e., protospacer sequence). dCas-TAD is sent into accessible promoter regions of target genes. The TAD recruits components of the transcriptional machinery which leads to the transcription of the target gene. By using multiple guide RNAs, several genes/promoters can be targeted simultaneously.
Figure 4. Targeted gene activation approaches via CRISPR/Cas technology. (a) targeted chromatin remodeling via a chromatin active domain (CAD). DNA is depicted in dotted black lines, nucleosomes are cyan, heterochromatic histone modifications are grey pins while euchromatic histone modifications are green pins. A fusion protein of dCas and a CAD is targeted into a nucleosome-free region (NFR) in a heterochromatic area where it modifies the local histones. This leads to an increase in DNA accessibility and promotes transcription. (b) Targeted gene activation by dCas fused to a trans-activating domain (TAD; e.g., VPR). DNA is depicted as a black bar, genes are either grey (inactive) or green (active) arrows. The colored bars in front of the genes are target sites for dCas-TAD (i.e., protospacer sequence). dCas-TAD is sent into accessible promoter regions of target genes. The TAD recruits components of the transcriptional machinery which leads to the transcription of the target gene. By using multiple guide RNAs, several genes/promoters can be targeted simultaneously.
Pharmaceutics 14 01837 g004
Table 1. List of platforms and databases used for BGC prediction. Ordered by the most recent update. Status March 2022.
Table 1. List of platforms and databases used for BGC prediction. Ordered by the most recent update. Status March 2022.
Tools and Database Used for BGC Predictions
Tool/
Database
TypePurposesLast UpdateLink
(Accessed on 15 March 2022)
Ref.
EMBL-EBI/ENADBuncurated Database (large repository: DNA, RNA, genome annotations)regular updatesebi.ac.uk[87,88]
GenBankDBuncurated Database (large repository: DNA, RNA, genome annotations)every 2 monthncbi.nlm.nih.gov/genbank[89]
GNPS DB“Global Natural Products Social Molecular Networking”
community-curated mass spectrometry database
2022gnps.ucsd.edu[90]
fungiSMASHPTGold standard for BGC prediction. Allrounder with an extensive variety of integrated tools and algorithms, web-application or stand-alone; Rule-based and rule-independent algorithms. Versions for fungi, bacteria and plants available. Fungal version has additional BGC border detection based on TF-binding sites 2021https://fungismash.secondarymetabolites.org/#!/start[78]
NP-atlas 2.0DBcurated community database of microbially-derived SMs
compounds from marine macro algae and diatoms are excluded
2021npatlas.org/[91]
TIGRFAMsDBcurated database of protein families (mainly prokaryotic)2021ncbi.nlm.nih.gov/genome/annotation_prok/tigrfams/[92]
PFAM v35DBpartly curated database protein domains and families2021pfam.xfam.org/[93]
FunOrderPTdetection of essential genes within a BGC by a co-evolutionary approach2021github.com/gvignolle/FunOrder[72]
ARTS 2.0PTAntibiotics cluster prediction by detection of resistance genes; web application2020arts.ziemertlab.com[83]
HMMERPTDNA based Homology search via pHMM;
as source code or as web application
2020hmmer.org/[94]
CO-OCCUR PTgenome-wide co-occurence of BGC-related gene pairs; source code available2020github.com/egluckthaler/co-occur[37]
TOUCAN PTBGC prediction via phylogeny, function, composition at amino acid level (k-mers, Pfam protein domains, and GO terms);
source code (python/perl) available
2020github.com/bioinfoUQAM/TOUCAN[95]
antiSMASH Database v3DBhigh-quality genomes and BGCs of bacteria, archaea, and fungi2020antismash-db.secondarymetabolites.org/[96]
NORINEDBmanually curated DB for NRPS2020bioinfo.lifl.fr/norine/[97]
FRIGGPTAntibiotics BGC prediction based on presence of resistance gene; Pipeline that combines different tools; Python and R-based script available2019zenodo.org/record/2560245#.YjlMHepBw5s[85]
BIG-SCAPE + CORASONOutput refinementtakes output of antismash and generates phylogenic BGC family networks; software application2019git.wageningenur.nl/medema-group/BiG-SCAPE[98]
MiBIg 2DBcurated database (community effort and institute curation) of known BGCs and their products2019mibig.secondarymetabolites.org/[99]
IMG-ABC v5.0DBpredicted BGCs with environmental metadata (prediction via antiSMASH v5)2019img.jgi.doe.gov/abc-public[100]
Deep-BGCPTDeep learning strategy to determine BGC product classes and chemical activity2019github.com/Merck/deepbgc[101]
RiPPMinerPTRiPPs; web-application2017nii.ac.in/rippminer.html[102]
SBSPKSv2 PTPKS/NRPS (structure/ substrate/ product)2017nii.ac.in/sbspks2.html[103,104]
SEMPI 2PTgenome-based PKS I modular product prediction; web-application2017pharmaceutical-bioinformatics.de/SeMPI/[105]
CASSIS and SMIPSPTPromoter base- BGC prediction focused around SM-key enzymes, online or as application2016sbi.hki-jena.de/cassis[106]
FunGene-ClusterSPTTranscriptome and genome-assisted BGC prediction; standalone and web application2016fungiminions.shinyapps.io/FunGeneClusterS[107,108]
2metDB/SecmetDBPTPKS/NRPS-prediction tool; software application2015sourceforge.net/projects/secmetdb/[109]
GNPPTStructure prediction and LC-MS data peak identification; web-application2015magarveylab.ca/gnp/[110]
Smiles2Mono-mers (s2m)PTpredict monomers from polymeric structure; webserver and standalone2015bioinfo.lifl.fr/norine/smiles2monomers.jsp[111]
ClusterFinderPTBGC detection via Pfam domain occurrence in genome and metagenomic data;
software application and integrated into
antiSMASH and IMG-ABC
2014github.com/petercim/ClusterFinder[82]
MIPS-CGPTClusters of BGC-related Pfam-domains outside of syntenic blocks; web-application2014fung-metb.net/
(currently offline)
[47]
MIDDAS-MPTMotif-Independent transcriptome and genome-assisted BGC prediction2013133.242.13.217/MIDDAS-M
(currently offline)
[86]
ClusterMine-360DBcrowd-source, semi and auto-curated database of microbial NRPS and PKS2013clustermine360.ca/
(currently offline)
[112]
NRPS-predictor2PTpredict bacterial and fungal NRPS adenylation domain substrate specificity; web-application2011nrps.informatik.uni-tuebingen.de/
(currently offline)
[113]
SMURF PTRule-based-backbone gene search via HMMs in genome data; web-application2010jcvi.org/smurf/[79]
CLUSEANPTPKS/NRPS-prediction tool; software application2009bitbucket.org/tilmweber/clusean[114]
Abbreviations: Database (DB), Prediction Tool (PT).
Table 2. Successful application of BGC induction by mimicking environmental conditions or by genetic manipulation of global regulators.
Table 2. Successful application of BGC induction by mimicking environmental conditions or by genetic manipulation of global regulators.
BGC Activation by Natural Triggers and Manipulation of Their Global Regulators
Section 1: Aspergillus and Penicillium
TriggerOrganismObservationRef.
ApyapA∆Aspergillus parasiticusApyapA∆: increased aflatoxin production[173]
csnE∆Aspergillus nidulanscsnE∆: production of orsellinic acid-related metabolites: orcinol, diorcinol,
cordyol C, violaceol I and violaceol II; activation of DHMBA cluster
[181,182]
mcRA∆ (TF)Aspergillus nidulans
Aspergillus terreus
Penicillium canescens
mcRA∆: Upregulation of secondary metabolites in all three organisms
A. nidulans dereplication strain: two new metabolites from the cichorine pathway, production of felinone A
new compounds of other organisms not identified
[175]
sumO∆Aspergillus nidulanssumO∆: production of Asperthecin, decrease of austinol/dehydroaustinol and sterigmatocystin[183]
Addditives
(supernatant extract)
Aspergillus nidulans
Streptomyces rapamycinicus
a novel a guanidine containing macrolide named polaramycin B, produced by S. rapamycinicus, is responsible for the derepression of the ors- cluster and its derivates in A. nidulans[184]
AdditivesPenicillium citrinumCultivation in presence of rare earth metals (i.e., scandium chloride) and isolation of three new peptide derivatives[185]
C-source
∆/OE of creA
Aspergillus flavuscreA∆ (Carbon catabolite repressor)
Growth defects, impaired conidia production, increased amount of sclerotia, near abolishment of aflatoxin production, impaired virulence
OE of creA: similar to wild type
[122]
C-sourceAspergillus nidulansCarbon source dependent BGC expression[186]
C-source
laeA∆
Penicillium expansumsucrose decreases both patulin and laeA levels and increases creA expression
LaeA regulates patulin BGC, laeA results in less virulence
[144]
Cultivation (chemostat)
N starvation
Aspergillus nidulansProduction of antiproliferative compounds sanghaspirodins A and B
Convergence of two BGCs pathways (anthraquinone and orsellinic acid-derived)
[128]
Light
Temperature
veA∆
laeA∆
Aspergillus fumigatusConnection between temperature and light regulation of 11 BGCs
Involvement of VeA at 37 °C and LaeA at 30 °C and 37 °C in BGC regulation
[157]
Light
velB∆
veA∆
Aspergillus nidulansvelvet complex VelB/VeA/LaeA regulates developmental regulation and secondary metabolism in a light-dependent manner.
velB∆ or veA∆: defects in sexual fruiting-body formation and secondary metabolite production
[138]
Light
laeA∆, veA∆, velB∆
Aspergillus ochraceuslaeA∆/veA∆/velB∆: differential regulation of 66% of all BGC backbone genes (majority downregulated)
drastic reduction of Ochratoxin A production
[136]
Light
Velvet:
∆Apc.LaeA∆ Apc.VeA∆
Aspergillus pachycristatusApc.LaeA∆ or Apc.VeA reduce production of echinocandin B and sterigmatocystin
Impact of deletions on aerial hyphae, pigmentation, development of conidiophores, conidial germination rate, and ascospore maturation
[134]
Media
Mimicking of plant and rhizosphere environment
Nitrogen, Metals, Amino acids
Aspergillus terreus(independent) signals for terrain production: methionine, nitrogen limitation, iron starvation (=host rhizosphere)
Global regulators AreA and AtfA essential for terrain production
[119]
Media composition
Carbon source
Amino acids
CreA
Aspergillus terreusdiscovery of isoflavipucine and dihydroisoflavipucine
production only in the presence of certain amino acids, alkaline pH, and strictly repressed in the presence of glucose (CreA).
[167]
Media
C-source
Penicillium citrinumIncreased production of citrinin on glucose compared to sucrose[187]
Media
HDACi
DNMTi
Aspergillus awamoriMedia: MEA, rice, nutrient, triptone soya agar
HDACi: valproic acid, nicotinamide, trichostatin A
DNMTi: 5-azacytidine
Different metabolic profiles between media and epigenetic modifiers
Nicotinamide was found to be the best epigenetic inducer
Rice grains were the best medium for SM induction
[188]
Metals and Trace elementsAspergillus fumigatusXanthocillin BGC-derived isocyanides help accumulate copper and exhibit antimicrobial activity[166]
Metals and Trace elementsAspergillus fumigatusHapX and SreA, which oppositely regulate iron homeostasis also control an iron-dependent secondary metabolite network comprising the virulence factor hexadehydroastechrome (tryptophan-derived iron (III)-complex) which causes an iron starvation phenotype upon overexpression[165]
Metals and Trace elements
Oxidative stress
sidC∆
Aspergillus nidulansNRPS SidC involved in iron regulation (excess and starvation)
Upregulated by oxidative stress
sidC∆: bad iron utilization (reduced growth under iron-starvation and higher iron demand), delayed germination under iron-depleted conditions, higher sensitivity of conidia to oxidative stress, no cleistothecia formation
[169]
Metals and Trace elementsPenicillium brasilianumCuSO4 and MnSO4 alter the SM-profile
production of a series of cyclodepsipeptides
repression of verruculogen biosynthesis
[163]
Metals and Trace elementsPenicillium urticaeeight metal ions tested
manganese dependent conversion of 6-methylsylicylic acid to patulin
[164]
OE laeAAspergillus nidulansoverexpression of laeA: activation of the Terrequinone (tdi) cluster and production of terrequinone A[189]
Osmotic &
Saline stress
Aspergillus aculeatus
Three ex-type strains from three different collections
Osmotic (glycerol) and saline stress had different impacts on SM profile of fungal strains. Also strain-source specific chemotypes.
Same morphology across all treatments
[159]
Oxidative stress
Carbon source
Aspergillus flavusStrains that produce less aflatoxin are more prone to oxidative stress and have more differentially regulated genes (also SM genes) under oxidative stress conditions
carbon source affects BGC expression and production
[190]
Oxidative stress
ApyapA∆
Aspergillus ochraceusApyapA∆ shows missing redox balance which triggers aflatoxin synthesis and an involvement of oxidative stress in ochratoxin A regulation[174]
Oxidative stress
veA∆
Aspergillus parasiticustranscription factor AtfB regulates conidial tolerance towards oxidative stress
AtfB binds to promoters with CRE sites in the aflatoxin cluster under inducing conditions. veA∆ abolishes these interactions
[171]
pHA. nidulans
A. parasiticus
Sterigmatocystin/aflatoxin expressed at higher levels at acidic pH than on neutral or alkalic pH
Constitutive active PacC in A. nidulans reduced sterigmatocystin production
[150]
pH
Acidic stress
Acid-Tolerant Fungi Penicillium
Aspergillus
Talaromyces
Cladosporium
Allophoma
Alternaria
Trichoderma
Supernatant extracts: ¾ shows cytotoxic activity,
<¼ show antimicrobial or anti-H1N1 activity
pH-related chemical diversity of P. oxalicum
[149]
pH
AcpacC∆
Aspergillus carbonariusPacC involved in pathogenicity and Ochratoxin A production
AcpacC∆: impaired fungal growth at neutral/alkaline pH, diminished gluconic and citric acid production, reduced virulence toward grapes and nectarine fruits
[148]
pH
Light
Temperature
Aspergillus flavusAflatoxin production:
5x increase of in darkness
pH 4 media produces 30 times more than pH 7.4
strongly reduced production at 18 or 40 °C
[191]
pH
pacC
Aspergillus nidulansPenicillin highest at alkaline pH and in mutated pacC (mimics alkaline conditions)
Production was lowest at acid pH and in strains that mimic acid pH
[151]
Solid-state vs. submerged fermentation
Media
Aspergillus oryzaeSolid-state: Increase of coumarins and oxylipins, higher antimicrobial activities
submerged: Terpenoids abundant
[178]
Solid-state vs. submerged fermentationPenicillium expansumSubmerged fermentation: strongly increased production of polyketide metabolites (agonodepside B, rotiorin, verrucosidin, and ochrephilone)
Solid-state fermentation: exclusive production of meroterpenoid compounds (andrastin A and C)
[177]
TemperatureAspergillus fumigatusTemperature affects stress tolerance, pigmentation, and trypacidin accumulation in conidia[192]
TemperatureAspergillus fumigatusEndocrocin biosynthesis in spores is temperature-dependent and endocrine is connected to pathogenicity[193]
TemperaturePenicillium oxalicum
P. citrinum
Unique metabolite profile under temperature stress[194]
Temperature Oxidative stress
Culture conditions
Aspergillus flavusFloat culture method
Stress conditions:
aflatoxin is downregulated under stress
secondary metabolites with antioxidant properties (e.g., kojic acid and imizoquins) were upregulated under stress
[118]
Velvet:
ΔlaeA
laeA-OE
Aspergillus spp.laeA∆: abolishment of sterigmatocystin, penicillin, and lovastatin
laeA-OE: increased penicillin and lovastatin production
laeA expression is negatively regulated by AflR
[142]
mpkA∆
(kinase knockout library)
Aspergillus nidulansmpkA∆: activation of the pkf-cluster and production of aspernidine A[195]
Section 2: Fusarium
TriggerOrganismObservationRef.
∆/OE lae1Fusarium fujikuroiΔlae1: downregulation of unknown NRPS4 cluster
OE-lae1 induced beauvericin cluster and STC7-cluster (product not identified)
generally LaeA has a substantial impact on many SMs
[147]
∆areA, ∆areBFusarium fujikuroideletions have an impact on >50% of BGCs
∆areB: upregulation of cryptic clusters PKS09, NRPS04, NRPS21
[127]
∆areA, areBFusarium fujikuroiAreA and AreB are required for gibberellic acid production[196]
∆pacCFusarium fujikuroipacC deletion de-represses bikaverin cluster-complementation with active pacC variant represses bikaverin cluster[153]
Global regulator of development
Δcsm1 (NsdD)
Fusarium fujikuroiΔcsm1: elevated conidiation in wt and also in non-sporulating ΔveA strain. Induction of BGC for bikaverin and fusarubins
Impact on 19 of 47 BGCs
[176]
Light
N-source
∆veA, ∆laeA
Fusarium oxysporumvelvet complex and nitrate metabolism intertwined
∆veA or ∆laeA: growth impaired on nitrate/nitrite
AREA involved in chromatin accessibility of two velvet-regulated BGCs (beauvericin and siderophore ferricrocin)
[131]
Light
N- source
pH
∆areA, ∆areB
∆vel1, ∆lae1
∆pacC
Fusarium fujikuroi∆areA: reduces fusaric acid production
∆areB: abolished fusaric acid production
∆vel1 and ∆lae1: significant reduction of fusaric acid
∆pacC: abolishment of fusaric acid cluster expression
[197]
Light
∆fgwc-1
∆fgwc-2
Fusarium graminearum∆fgwc-1/∆fgwc-2:
impaired carotenogenesis, photoreactivation and maturity of perithecia
mutants produced more aurofusarin and trichothecenes independent from light and have derepressed conidiation during constant light
[198]
Light
∆vel1, ∆vel2, ∆lae1
Fusarium fujikuroi∆ffvel1 and ∆fflae1: velvet positively regulates gibberellic acid, fumonisins and fusarin C, velvet negatively regulates bikaverin[141]
N-source
pH
Host infection
Fusarium fujikuroi
F. circinatum
F. mangiferae
F. oxysporum
PKS19 cluster is induced on rice but not maize
Gibberelic acid cluster only expressed in F. fujikuroi
Different conditions for different metabolites necessary: fusarins: high nitrogen, bikaverin and fusarubins: acidic low nitrogen or alkaline low nitrogen, Fumonisin B1: acidic low nitrogen
[125]
N-source
∆areA
Fusarium graminearumasparagine was found to be a preferential nitrogen source.
∆areA led to poor growth on NaNO3
Utilization of aspartic acid, histidine, isoleucine, leucine, threonine, tyrosine, and valine as nitrogen sources was shown to depend of a functional AREA.
AREA is required for the production of deoxynivalenol (DON), zearalenone, and fusarielin H regardless of the medium
[129]
N-source
∆areA
Gibberella fujikuroi∆areA: significantly reduced gibberellin production[199],
[200]
N-source/starvation
∆ of N-pathway regulators
Fusarium FujikuroiEffect of N-source (glutamine, nitrate) and starvation in different mutants
ΔniaD, ΔnrtA, ΔnirA and ΔareA
expression of backbone genes of gibberellic acid and bikaverin BGC after 72 h in starvation in wt and all mutants except for gibberellic acid in ΔareA
[124]
pH
PAC1
Fusarium graminearumPac1 negatively regulates trichothecene BGC; induction under acidic pH
Δfgpac1: reduced development under neutral and alkaline pH, increased sensitivity to H2O2 and early onset of Tri expression and product synthesis
[155]
TemperatureFusarium langsethiae
F. sporotrichioides
Impact of temperature and time on spore amount and T-2, HT-2 toxin production[201]
Section 3: Other Species
TriggerOrganismObservationRef.
AdditivesColletotrichum gloeosporioidesTriggering silent BGCs by natural dietry components: Grape skin and turmeric extracts resulted in 20 and 14 additional compounds[202]
C-sourceSpicaria elegansstarch vs. glucose
novel spicochalasin A and five new aspochalasins (M–Q)
[203]
Co-cultivation
Additives
Isaria felinapotassium bromide and cocultivation with Aspergillus sulphureus:
discovery of Isariketide B and Oxirapentyn L
[204]
Cultivation optimization for SM productionPhomopsis sp. Hant25Optimization of different media and cultivation parameters for overproduction of mycoepoxydiene
Parameters: combination of stationary and agitated, solid support (cellulose paper disc), different medias, cultivation time
[179]
Light
∆lreA
(=WC-1)
Alternaria alternataOpposite light regulation of mycotoxin and spore production
spores decreased under light and
production of alternariol increased two to three times
altertoxin production was strictly dependent on light.
∆ white-collar 1 (WC-1) gene (lreA):
derepression of spore formation in dark and in light.
altertoxin formation strongly induced in the dark
alternariol formation partially dependent on LreA
still able to partially respond to blue light
[205]
Light
Velvet
∆velB
∆veA
∆laeA
Cuetootiopsis microsporaVelvet Complex involved in the development, stress tolerance, and secondary metabolism
∆velB: hypersensitive to osmotic stress and congo red
pestalotiollide B production requires velB and laeA.
veA inhibits pestalotiollide B biosynthesis
[132]
Light
Velvet
Iron starvation
Neurospora crassaVelvet controls development, secondary metabolism, and light-dependent carotenoid production
Heterotrimeric velvet complex (VE-1/VE-2/LAE-1) suppresses the production of the siderophore coprogen under iron starvation
[133]
Light
C-source
∆CRE1
Trichoderma reeseiIntertwined regulation of light and carbon pathways
CRE1- regulated cluster is responsible for light-dependent production of dihydrotrichotetronin
Biosynthesis of the antibiotic paracelsin was influenced
[135]
Light
∆SUB1
Trichoderma reeseiSUB1 is light-regulated
SUB1 affects growth and is required for female fertility
SM is regulated by SUB1 in a light- and nutrient-dependent manner
[206]
Media
Salts
Cultivation type
Nigrospora sp. MA75Discovery of 2,3-didehydro-19a-hydroxy-14-epicochlioquinone B, 6-O-desmethyldechlorogriseofulvin 6′-hydroxygriseofulvin
Media: NaCl vs. NaI, static/solid vs. agitated/liquid, rice-based vs. yeast extract
[207]
Media
HDACi
Drechslera sp.,HDACi: SAHA, valproate, octanoylhydroxamic acid
2 new metabolites: chromanone and prenylhydroxybenzoic acid
different metabolome between MEA and minimal media
biotransformation of the inhibitors occurred
[208]
Media
HDACi
DNMTi
Cladosporium resinaeMedia. Czapek-DOX, YES, PDA, starch, liquid vs. solid
HDACi: SAHA
DNMTi: 5-AZA
Czapek yeast extract broth: more total secondary metabolites
DNMTi and HDACi: expressions of silent genes
[209]
N-source
∆AcareA
Acremonium chrysogenum AcareA is required for nitrogen metabolism and cephalosporin production [130]
OE-laeAChaetomium globosumoverexpression of laeA: activation of the chaetoglobosin gene cluster and production of chaetoglobosin Z[145]
pH
ΔpacC
pacCc
Trichoderma virensΔpacC: decreased competitiveness against Rhizoctonia solani and Sclerotium rolfsii
Constitutive PacC: wt like overgrowth of R. solani
[152]
Plant HormonesArthrinium saccharidiscovery of the polyketides kinetin, 6-benzylaminopurine, and forchlorfenuron[210]
Salinity47 marine fungi isolatesNaCl and KCl had growth-promoting effect on most marine fungi and 15% of isolates showed increased antifungal activity against Candida albicans,
NaCl had impact on the metabolite profile
[160]
SalinitySpicaria elegansDifferent secondary metabolites on different saline conditions (3% and 10%)
Four metabolites, only at 10% salinity: (2E,2’Z)-3,3’-(6,6’-dihydroxybiphenyl-3,3’-diyl) (new compound), diacrylic acid, aspulvinone E, aspochalasin E and trichodermamide B
[162]
Solid vs. liquid fermentation
Solid growth support
Cylindrocarpon sp.
Acremonium sp.
Penicillium sp.
Addition of inert absorbent polymeric attachment supports in solid state and submerged fermentation leading to distinct metabolite profiles
Impact on culture morphology and relative metabolite yields
Solid agar with moist polyester-cellulose paper:
Cylindrocarpon sp. (pyrrocidines A and B), Acremonium sp. (acremonidins A–E)
Agitated submerged cultures:
Complex HPLC metabolome, decrease in target compounds
Liquid fermentation with various inert growth supports (polypropylene, polypropylene cellulose, polyester-cellulose, or polyurethane): production of
rugulosin, skyrin, flavomannin, and the new compound WF159-A (bisanthracene)
[180]
Temperature40 polar fungal isolates45% of fungal strains showed antimicrobial activity
Different fungal metabolite profiles dependent on temperature (4, 10, 15, 28 °C)
[158]
ΔvmlaeAValsa mali ΔvmLaeA : reduced virulence and differential regulated secondary metabolism profile (31/60). [143]
The table is separated into three sections: Section 1 contains studies that were conducted on fungi of the genus Aspergillus or Penicillium, Section 2 comprises fungi from the genera Fusarium and Section 3 presents studies with filamentous fungi of other genera. The sections themselves are ordered alphabetically by the trigger that was applied. Signs and Abbreviations: Δ: deletion of a gene; OE: overexpression of a gene; TF: transcription factor.
Table 3. Discovery of new secondary metabolites by co-cultivation. Ordered by Set-up type.
Table 3. Discovery of new secondary metabolites by co-cultivation. Ordered by Set-up type.
BGC Activation by Co-Cultivation
SetupMethodOrganismObservationRef.
1Co-cultivation of two developmental stages (morphs) of a marine alga-derived fungus
Pre-grow both morphs on agar plates (2–3 weeks) → small liquid pre-cultures → combine mono-cultures after 14 days → cultivate for 15–20 days
Aspergillus alliaceus and its teleomorph Petromyces alliaceusProduction of the cytotoxic chlorinated bianthrone allianthrone A (only liquid)
Morphs develop new phenotype in co-culture under liquid but not solid fermentation
Restreaks after co-culture keep phenotype and can produce bianthrones on liquid and solid media
[218]
1Submerse co-culture and membrane-separated culture
5 mL of pre-culture (A. flavipes and Streptomyces sp. in a ratio of 1:4 (v/v)) were added to 200 mL medium 8 days of cultivation
Aspergillus flavipes marine actinomycete Streptomyces sp.Production of five aspochalasins and rosellichalasin by fungus with cytotoxic effects against Streptomyces. Physical interaction necessary[219]
13 media (ISP2, GYE, F) inoculated with 3d old pre-cultures. 4L culture, fungal preculture was incubated 2 d earlier, 6 h before harvest 50 g/L Diaion HP-20 resin was addedAspergillus fumigatus MR2012 (marine)
Streptomyces leeuwenhoekii (C34 and C58) (desert)
Fungus: production of luteoride D (luteoride derivative), pseurotin G (pseurotin derivative), terezine D, 11-O-methylpseurotin A
Bacteria C34: Chaxapeptin C58: Chaxapeptin doubled, pentalenic acid
[220]
116 h old mycelia of A. fumigatus were washed and placed in fresh AMM with 1/20 vol of the streptomyces culture or with 1/20 filtered supernatant. Harvest after 12 hAspergillus fumigatus
Streptomyces rapamycinicus
induction of fungal PKS FgnA, the discovery of bacteria-specific germination inhibitor fumigermin[221]
12 day 30 °C liquid agitated co-culture (93 mL London + 32 mL GYM medium). simultaneous inoculation of both MOs (equal amounts of spores)Aspergillus nidulans
Streptomyces mobaraensis
bacterial glycopeptide (antibiotic) triggers antibacterial and iron-chelating tropolonesanhydrosepedonin and antibiotic C (A. nidulans), polyketide tripyrnidone (novel structure)[222]
1Solid rice medium Aspergillus versicolor
Bacillus subtilis
two new tetralones; aspvanicin A and aspvanicin B
several other cryptic SMs; cytotoxic activity against mouse lymphoma cells
[223]
1Confrontation assay on ISP2 agar
MALDI-TOF straight from agar
Inoculation for 48 h at 30 °C
Aspergillus fumigatus
Aspergillus niger
Bacillus
amyloliquefaciens (GA40)
The coral-derived bacteria GA40 produced an antifungal iturin[224]
1Precultures on rice media 1 week then co-culture for 2 weeks. Axenic cultivation with potassium bromideAspergillus sulphureus KMM 4640
Isaria feline
discovery of Isariketide B (rice media with KBr) and Oxirapentyn L (Co-cultivation)[204]
1Fungal bacterial stepwise liquid co-culture. Addition of bacterial broth (1/1000) to fungal culture on day 3. Further cultivation for 2 daysEmericella sp.
(CNL-878; marine fungus)
Salinispora arenicola (CNH-665; ascomycete)
Emericellamides A and B synthesis by Emericella sp.[225]
1Solid agar confrontation assay (5 days)
Submerged fermentation (2 mL of each pre-culture and cultivation for 5 days in broth)
Co-cultivation of MF028 with five other strains
time-course
Chaunopycnis sp. (CMB-MF028)
Trichoderma hamatum (CMB-MF030)
Co-isolated from mollusc Siphonaria sp.
Chaunopyran A production
Cultivation on sterilized mycelia showed no induction
[226]
1Extraction of confrontation zone (Agar)
Incubation of several weeks
Dereplication via monocultures
Method development:
~600 cultures by confrontation assay on agar
UHPLC–TOF-MS fingerprints of mono and cocultures have been prepared to filter out new peaks
Fusarium isolates (clinical, soil
or plant-derived)
Aspergillus clavatus, Trichophyton rubrum, Acremonium, Cladosporium sp. Hohenbuehelia reniformis
Bionectria ochroleuca Eutypa lata
Chemical novel MS spectra have been recorded
Results indicate that a large portion of fungi produce new metabolites in co-culture
[213]
1Solid state confrontation cultivation
17 basidiomycetes, 136 co-cultures
Molecular network analysis
Liquid co-cultivation for 5–30 days
Ganoderma
applanatum
Trametes versicolor
discovery of novel xylosides
3-phenyllactic acid and orsellinic acid
in G. applanatum
[227]
1Addition of bacterial broth (1/1000) to fungal culture on day 3. Harvest after 24 h. Screen of 50 fungal strainsLibertella sp.
(CNL-523)
α-proteobacterium (CNJ-328)
Discovery of Libertellenones A–D, produced by the addition of marine α-proteobacterium to a 3-day Libertella sp. culture.[228]
1Submerged fermentation
Different pre-grow times before co-cultivation. Optimal: addition of fungal culture to 2-day bacterial culture and further incubation for 7 days. Other setups showed metabolome of either bacterial or fungal axenic culture
Penicillium sp.
(WC-29-5) Streptomyces fradiae (007)
Production of five phenolic polyketides[229]
1Solid rice media, Cultivation parameters
Optimal: inoculation of Penicillium sp mycelium plug. and seed medium of Bacillus sp. simultaneously at the center of petri dish 15d culture
Penicillium sp.
(DT-F29)
Bacillus sp. (B31)
Ten bioactive 2,5-diketopiperazines by Penicillium sp.[230]
1Co-cultivation on solid rice mediumPenicillium strains (IO1 and IO2) isolated from spongeAccumulation of norlichexanthone and monocerin, biotransformation of the antifungal pyridoxatin to methyl-pyridoxatin[231]
1Liquid co-culture with seawater-based marine
nutrient medium. After 24 h, bacterial culture is added to fungal cultivation following further 6d cultivation
Pestalotia
(marine fungus)
CNJ-328
(marine bacterium)
Discovery of pestaolne (antibiotic)[216]
1Submerged fermentation and sea water agar, inoculation of flasks with spores of both fungi, cultivation for 9 daysTalaromyces aculeatus
Penicillium variabile
Four new polyketides (penitalarins A/ B/C and nafuredin B)
cytotoxicity against some human cancer cell lines
[232]
2Mycelia from A nidulans o/n preculture was combined with 1/20 (v/v) Sreptomyces preculture in fresh AMM medium. Further cultivation at 37 °C for 24 hAspergillus nidulans with a collection of 58 soil-dwelling actinomycetesPhysical interaction between the fungus and Streptomyces hygroscopicus. Induction of orsellinic acid and derivatives (lecanoric acid, yellow polyketides)[215]
2Co-culture in Warkingsman’s agar (Confrontation assay) and liquid medium (5 days) Fusarium graminearum
vs.
12.000 bacteria
reduced virulence of F. graminearum due to Pseudomonas piscium
(FgGcn5 inhibitor phenazine-1-carboxamide)
[214]
3treatment of Aspergillus with
extracts of Streptomyces cultures
Aspergillus nidulans
Streptomyces rapamycinicus
a guanidine containing macrolide, polaramycin B that is constitutively produced by the bacteria triggers the production of the ORS BGC[184]
3analytical scale microbial cultivations in a 24-well plate
cultivation in presence of metabolites from other cultivation
Aspergillus sp.
Streptomyces sp.
bacteriostatic cyclo-(L-Phe-trans-4-hydroxy-L-Pro) from Aspergillus stimulates nitric oxide production by Streptomyces which triggers silent BGC of Aspergillus: fungistatic heronapyrrole B[233]
Table 4. Successful application of overexpression methods for induction of BGCs. Entries are ordered alphabetically by organism.
Table 4. Successful application of overexpression methods for induction of BGCs. Entries are ordered alphabetically by organism.
Targeted BGC Activation by Overexpression
MethodTarget/ParameterOrganismObservationRef.
TFfsqA (TF)Aspergillus fumigatusOE of fsqA (fsq gene cluster)
production of pyrido [1,2-b]isoquinolines fumisoquin A and fumisoquin B
[249]
TFxanCAspergillus fumigatusOE of xanC: Activation of xan cluster and production of xanthocillin derivatives, the sulfated xanthocillin derivative BU-4704 and xanthocillin X monomethyl ether[250]
CRISPRaAN8506 AN8507 mdpEAspergillus nidulansMethod establishment: upregulation of putative BGC genes (AN8506, AN8507) and the silent mdP cluster via nucleosome map assisted targeting of inducible VPR-dCas9 into accessible chromatin in promoter [251]
CRISPRabreF, fuml, fwnAAspergillus nidulansMethod establishment: fusion of different chromatin active enzymes/domains to dCas9, activation of breF and fuml via p300-dCas9 (HAT),
repression of breF by HAT GcnE-dCas9 and HDAC HosA-dCas9 and RpdA-dCas9
activation of fwnA by HosA-dCa9
[252]
CRISPRa (multiplex)mic- BGCAspergillus nidulansmethod establishment: activation of mic cluster by dCas12a-VPR (multigene
CRISPRa): dehydromicroperfuranone production
[253]
Hybrid Activator(+)-Asperlin clusterAspergillus nidulansfusion of the DNA binding domain of AlnR (pathway-specific TF) and the trans-activation domain of AfoA lead to the activation of the (+)-Asperlin cluster[246]
TFaspyridone cluster (asp)Aspergillus nidulansoverexpression of the pathway-specific TF apdR: activation of the silent BGC and production of aspyridone[237]
TFctnRAspergillus nidulansoverexpression of the TF ctnR: Cluster activation (7 genes) and asperfuranone biosynthesis[238]
TFBB geneseveral BGCsAspergillus nidulansChicorine cluster activation
pkg cluster activation: alternariol, citreoisocoumarin, analogs of citreoisocoumarin
pki cluster: 6-Hydroxy-7-methyl-3-nonylisoquinoline-5,8-dione
pkhB /pkhA: 2,4-dihydroxy-6-[(3E,5E,7E)-2-oxonona-3,5,7-trienyl]benzaldehyde
Pkd: 2-ethyl-4,6-dihydroxy-3,5-dimethylbenzaldehyde
[243]
TFpbcRAspergillus nidulansoverexpression of pbcR resulted in upregulation of a 7-gene cluster and the production of ent-pimara-8(14),15-diene[254]
BB gene
Het. Expr.
micro-perfuranone (micA, BGC)Aspergillus nidulanspromoter exchange or heterologous expression (A. niger) of backbone gene micA was sufficient to produce microperfuranone [255]
BB gene
Het. Expr.
asqKAspergillus nidulansOE of BB gene and heterologous expression for stepwise pathway elucidation:
Products: 4′-methoxylated 6,7-benzodiazepinediones and 4′-methoxyviridicatin
[256]
Multiple
genes
inpE clusterAspergillus nidulanspromoter replacement of seven genes (inpC-inpB):
production of protease inhibitor fellutamide B
[248]
Multiple
genes
ivo clusterAspergillus nidulansOE of ivoA, B and C (and combinations): production of a dark pigment with the precursors N-acetyltryptophan and N-acetyl-6-hydroxytryptophan[247]
Multiple
genes
sgdA, D, C, FAspergillus nidulanspromoter replacements (4 genes): production of aspernidgulene A1 and A2[257]
TFazaR (TF)Aspergillus nigeroverexpression of azaR: Activation of aza cluster and production of azanigerones A–F[258]
TFaflRAspergillus parasiticusoverexpression of aflR lead to the production of aflatoxin under repressing conditions[259]
TFTF of pgm BGCAspergillus terreusOE of pgm cluster TF (ATEG_06205) led to the production of naphthoquinones[260]
Resistance geneneomycin resistanceAspergillus versicolorThe introduction of neomycin resistance leads to production of the new anti-tumor bioactive compounds[261]
TFNRPS31 BGC (APS-like) PKS19- BGCFusarium fujikuroioverexpression of pathway-specific TF (APS2) and complementation of enzyme (F. semitectum APS8) resulted in production of ApicidinF
OE of TF and PKS19 resulted in production of four novel compounds (Fujikurin A–D)
[125,262,263]
TFAPF2Fusarium fujikuroioverexpression of APF2 increases apicidin F synthesis[262]
Whole BGCBIK2 or BIK3Fusarium fujikuroiSeparate OE of BIK2 and BIK3: revelation of precursor oxo-pre-bikaverin[264]
BB geneDMATS1, FFUJ_09179Fusarium fujikuroiOE of DMATS1 yielded a reversely N-prenylated tryptophan[265]
TFTF22
(PKS-NRPS1)
Fusarium fujikuroiTet-on activation of TF22 (FFUJ_02222):
cluster activation and trichosetin production
[239]
TFFSL7 (TF)Fusarium graminearumoverexpression of the TF FSL7: activation of the FSL cluster (PKS9) and production of fusarielins, F, G, and H[266]
BB genePKS8Fusarium graminearumactivation of Gibepyrones and Prolipyrone B production[33]
TFFGM4 (fg3_54 BGC; NRPS)Fusarium graminearumOE triggers the production of fusaoctaxin A. Deletion compromises plant infection[267]
BB geneFSDSFusarium heterosporumOE of backbone gene resulted in novel Tyrosine-Derived 2,4-Pyrrolidinedione[268]
TFPEXANC (TF)Penicillium expansumOE of PEXANC resulted in the activation of the Citrinin cluster via the TF CtnA and production of Citrinin (evolutionary TF repurposing)
The theoretical cluster responsible for xanthocillin remained silent
[269]
Resistance geneneomycin resistancePenicillium purpurogenumIntroduction of neomycin resistance results in the production of the new cyclic dipeptide penicimutide[270]
CRISPRa (TF)macR (TF)Penicillium rubensinduction of macrophorin pathway-specific TF macR: production of macrophorin A, D and 4′-oxomacrophorin D[271]
Abbreviations: Heterologous expression (Het. Expr.), backbone gene (BB gene), pathway-specific Transcription factor (TF).
Table 5. Successful examples of heterologous expression of single genes and BGCs. Ordered by organism.
Table 5. Successful examples of heterologous expression of single genes and BGCs. Ordered by organism.
BGC Activation by Heterologous Expression
Target/ParameterOrganismsObservationRef.
high throughput
BGC activation
22 Species (Donors)
S. cerevisiae (Host)
method establishment: Hex
Expression of 22/41 fungal BGCs resulted in detectable compounds
[285]
156 FACs with random genome fragmentsAspergillus aculeatus (Donor)
Aspergillus terreus (Donor)
Aspergillus wentii (Donor)
Aspergillus nidulans (Host)
method establishment: FAC-MS, Identification of 15 new substances
Clusters/metabolites: A. terreus: ATEG_07067, ATEG_07500, sesterterpenoid, valactamide A, benzomalvin A/D
A. aculeatus: AACU_51108, AACU_59515,
A. wentii: ASPWE_027400, ASPWE_034272, ASPWE_042595, ASPWE_044725, ASPWE_085322, ASPWE_151732, ASPWE_163793
[284]
diketomorpholine BGCAspergilllus aculeatus (Donor)
Aspergillus nidulans (Host)
Elucidation of the biosynthetic pathway of acu-dioxomorpholine, featuring a new type of condensation domain, by FAC-MS[287]
ftmA (NRPS)Aspergillus fumigatus (Donor/Host)
Aspergillus nidulans (Host)
OE of the NRPS ftmA: Production of Brevianamide F (Product of NRPS, product of cluster is putatively fumitremorgin) [288]
pyr ClusterAspergillus fumigatus (Donor)
Aspergillus. oryzae (Host)
successive expression of the pyr cluster: Characterization of pyr cluster and isolation of the meroterpenoid pyripyropene[289]
micro-perfuranone (micA, BGC)Aspergillus nidulans (Donor)
Aspergillus niger (Host)
promoter exchange or heterologous expression of backbone gene micA was sufficient to produce microperfuranone [255]
asqJ
asqE-L
Aspergillus nidulans (Donor)
Escherichia coli (Host)
OE of asqE-L in host and elucidation of AsqJ function by expression in E. coli
Final products: 4′-methoxylated 6,7-benzodiazepinediones and 4′-methoxyviridicatin
[256]
wA (NWA)Aspergillus nidulans (Donor)
Aspergillus oryzae (Host)
expression of the WA-PKS: identification of the YWA1 compound[290]
terreazepine BGCAspergillus terreus (Donor)
Aspergillus nidulans (Host)
production of terreazepine[291]
39 BGCs from various species Donors:
Colletotrichum gloeosporioides
Alternaria alternata
Fusarium graminearum
Trichoderma viride
Aspergillus flavipes
Host: Aspergillus oryzae
method establishment: genome mining assisted automated and high-throughput (auto-HTP) biofoundry workflow
refactoring of 39 BGCs and production of 185 distinct terpenoids
[292]
STC5, STC3Fusarium fujikuroi (Donor)
E. coli (Host)
Use of homologues
STC3: (+)-eremophilene
STC5:(¢)-guaia-6,10(14)-diene
[55]
carotenoid BGC of Fusarium fujikuroiFusarium fujikuroi (Donor)
Aspergillus nidulans (Host)
method establishment: aflR scaffold as heterologous expression matrix; successful production of β-carotene[286]
STC1Fusarium fujikuroi (Donor)
E. coli (Host)
Heterologous Expression of STC1 identification of (–)-germacrene D[176]
FgJ03939 BGC Fusarium graminearum (Donor)
Saccharomyces cerevisiae (Host)
stepwise synthesis of several new sesquiterpene synthase-derived metabolites such as fusariumdiene, fusagramineol[293]
part of the fusarubin gene clusterFusarium solani (Donor)
S. cerevisiae (Host)
Expression of FVPPT, fsr1, fsr2, and fsr3 in S. cerevisiae (galactose inducible promoter) and biosynthesis of bostrycoidin[294]
pvhA-E (BGC)Penicillium variabile (Donor)
Aspergillus nidulans (Host)
Heterologous expression of genes pvhA-e: production of varicidin A and B[295]
psilocybin BGC Psilocybe carpophores (Donor)
Aspergillus nidulans (Host)
method establishment: Successful Expression of whole BGC as polycistronic RNA piconavirus assisted expression of the whole basidiomycete derived psilocybin BGC [296]
TESG_6702 to 6705Trichophyton tonsurans (Donor)
Aspergillus nidulans (Host)
Expression of the whole cluster resulted in the production of neosartoricins[297]
Table 6. Untargeted BGC activation by interference with global chromatin regulatory mechanisms. Ordered alphabetically by organism.
Table 6. Untargeted BGC activation by interference with global chromatin regulatory mechanisms. Ordered alphabetically by organism.
BGC Activation by Interference of Chromatin-Regulation
MethodOrganismObservationRef.
HDACi
DNMTi
162 mangrove endophytic FungiHDACi: sodium butyrate, DNMTi: 5-azacytidine
Activity against ESKAPE MOs and Mycobacterium tuberculosis and Leishmania donovani:
254/ 1608 extracts showed bioactivity, HDACi or DNMTi: 72 fungi produced active extracts, 40 fungi were active only in the control
[320]
Media
HDACi
DNMTi
Aspergillus awamoriMedia: MEA, rice, nutrient, triptone soya agar
HDACi: valproic acid, nicotinamide, trichostatin A, DNMTi: 5-azacytidine
Different metabolic profiles between media and epigenetic modifiers
Nicotinamide was found to be the best epigenetic inducer Rice grains was the best medium for SM induction
[188]
HDACi
DNMTi
Aspergillus calidoustus,
Aspergillus westerdijkiae
Aspergillus bombycis Aspergillus carbonarius
Aspergillus fischeri
HDACi: SAHA, DNMTi: 5-azacytidine
significant changes in secondary metabolite profile with examples of both induction and repression.
[321]
HDACi
DNMTi
Aspergillus clavatusHDACi: valproic acid, trichostatin A, butyrate
DNMTi: 5-azacytidine. N-acetyl-d-glucosamine
Source of peptone strongly influences the SM profile, epigenetic modifier effects SM profile which is metabolite-specific and dependent on growth media and incubation time
[322]
DeletionAspergillus nidulanscclA∆ (Bre2; H3K4me):
activation of mdp cluster: production of monodictyphenone, emodin, and emodin derivatives
activation of anti-osteoporosis polyketides cluster: F9775A and F9775B
[323]
DeletionAspergillus nidulansHepA occupancy decreases during sterigmatocystin activation
hepA∆/clrD∆ (H3K9me): upregulation of sterigmatocystin, isopenicillin A, terraquinone A
LaeA involved in reversing heterochromatic marks at ST BGC,
[305]
Deletion Aspergillus nidulanshosA∆ (HDAC): impact on H4ac, SM, carbohydrate metabolism, virulence, defense[324]
KnockdownAspergillus nidulansknockdown of rpdA (HDAC): production of aspercryptin A1 and A2 and fellutamides[325,326]
HDACiAspergillus nigerHDACi: SAHA, discovery of nygerone A[327]
Deletion Aspergillus nigergcnE∆ (SAGA/Ada; acetylation): activation of 12 silent clusters (1 novel compound): nigerpyrone, carbonarone A, pestalamide A, funalenone, aurasperone E, aurasperone A[328]
HDACi
DNMTi
Aspergillus sp.HDACi: suberohydroxamic acid, DNMTi: 5-azacytidine
discovery of diphenylether-O-glycoside (diorcinol 3-O-α-D-ribofuranoside)
[329]
HDACiBotryosphaeria mamaneHDACi: SAHA, Valporate, production of five unknown compounds[330]
Deletion Calcarisporium arbusculahdaA (HDAC): upregulation of 75% of BGC genes: isolation of 10 new compounds[331]
HDACi
DNMTi
Chalara sp. 6661HDACi: sodium butyrate, suberobishydroxamic acid, vorinostat; DNMTi: 5-azacytidine
vorinostat: four new modified xanthones, only one out of four inhibitors showed effect
Biotransformation of vorinostat
[332]
Media
HDACi
DNMTi
Cladosporium resinaeMedia: Czapek-DOX, YES, PDA, starch, liquid vs. solid; HDACi: SAHA; DNMTi: 5-AZA
Czapek, yeast extract broth: more total secondary metabolites
DNMTi and HDACi: expressions of silent genes
[209]
HDACi
DNMTi
Cyathus stercoreusnine previously undescribed sesquiterpenes[333]
Deletion Dothistroma septosporumΔDsKmt1 (H3K9me) or ΔDsKmt6 (H3K27me): Substantial decrease of H3K9me3 or H3K27me3. Increase of production of dothistromin in both cases[334]
Media
HDACi
Drechslera sp.,HDACi: SAHA, valproate, octanoylhydroxamic acid; 2 new metabolites: chromanone and prenylhydroxybenzoic acid; different metabolome between MEA and minimal media
biotransformation of the inhibitors
[208]
Deletion Epichloe festucaeΔclrD (H3K9me) and/or ΔezhB (H3K27me): reduction of H3K9me3 and/or H3K27me3
activation of bioprotective lolitrems (ltm) and ergot alkaloids (eas)
[307]
Deletion Fusarium fujikuroiΔhda1(hdaA; HDAC): production of beauvericin (BEA-cluster; NRPS22)[335]
Deletion Fusarium fujikuroiΔset2 and Δash1 (H3K36me): upregulation of PKS-NRPS9 and NRPS4[336]
Deletion Fusarium fujikuroiset1 (H3K4me): activation of PKSNRPS9, PKS type III, NRPS4, NRPS23, and DMATS3[337]
DeletionFusarium graminearumΔkmt6 (PRC2, H3K27me): Differential regulation of many SM genes,
upregulation of: NRPS9, NRPS14, NRPS16, NRPS19, PKS2, PKS3, PKS13, PKS22, PKS28, PKS29, STC5
[312]
DeletionFusarium graminearum (dereplication strain)kmt6 (H3K27me): discovery of Protofusarin and N-ethyl anthranilic acid, N-phenethylacetamide, N-acetyltryptamine
∆kmt6 and ∆fus1: tricinolone, tricinolonic acid
[338]
Deletion Fusarium mangiferaefmkmt1 (H3K9me): increased beauvericin biosynthesis, nearly abolished fusapyrone production; revelation of fusapyrone BGC (FmPKS40)[308]
Deletion Fusarium verticillioidesΔfvdim5 (H3K9me): increase of fumonisin B1 and melanin biosynthesis, reduced virulence[339]
Deletion Metarhizium robertsiiΔhat1: activation of 11 orphan BGCs
Meromuside A–I, Meromutides A and B
[340]
HDACi
DNMTi
Muscodor yucatanensisHDACi: SAHA, DNMTi: 5-azacytidine
different morphology, cultural characteristics, metabolites
[341]
HDACi
DNMTi
Nigrospora sphaericaHDACi: SAHA, sodium butyrate, valproic acid, quercetin
DNMTi: 5-azacytidine, hydralazine hydrochloride
induction of cryptic metabolites by all modifiers used
[342]
HDACiPenicillium brevicompactumHDACi: nicotinamide, sodium butyrate
total phenolic compounds increased drastically upon HDACi treatment
Nicotinamide: p-anisic acid, p-anisic acid methyl ester, benzyl anisate, syringic acid, sinapic acid, acetosyringone, phenyl acetic acid, gentisaldehyde and p-hydroxy benzaldehyde
Sodium butyrate enhanced the productivity of anthranilic acid and ergosterol peroxide
[343]
Deletion Penicillium chrysogenumhdaA (HDAC): new compound. Maybe crosstalk with sorbicillinoids cluster[344]
HDACiPhoma sp.SAHA triggers the production of (10’S)-verruculide B, vermistatin and dihydrovermistatin
Absence of SAHA: discovery of (S,Z)-5-(3′,4′-dihydroxybutyldiene)-3-propylfuran-2(5H)-one
[345]
HDACiTrichoderma harzianumHDACi: sodium butyrate; diterpenoids changed into cyclonerane sesquiterpenoids.
3 new compounds: cleistanthane diterpenoid, harzianolic acid A, harziane diterpenoid, harzianone E, cyclonerane sesquiterpenoid, 3,7,11-trihydroxy-cycloneran
[346]
KnockdownFusarium fujikuroiKnockdown of KMT6 (H3K27me; PRC2): differential regulation of SM-clusters,
upregulated SM genes: NRPS22, STC4, tetraterpene cyclase-encoding gene TETC1,
PKS-NRPS1, STC5, STC8, DMATS3, NRPS4, PKS2)
detection of six yet unidentified cluster-products: STC5: guaia-6,10(14)-diene, STC3: (+)-eremophilene
[55,315]
Abbreviations: HDACi (inhibition of histone deacetylases by chemical inhibitors), DNMTi (inhibition of DNA methyltransferases by chemical inhibitors), SAHA (vorinostat), 5-AZA (azacitidine).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Schüller, A.; Studt-Reinhold, L.; Strauss, J. How to Completely Squeeze a Fungus—Advanced Genome Mining Tools for Novel Bioactive Substances. Pharmaceutics 2022, 14, 1837. https://doi.org/10.3390/pharmaceutics14091837

AMA Style

Schüller A, Studt-Reinhold L, Strauss J. How to Completely Squeeze a Fungus—Advanced Genome Mining Tools for Novel Bioactive Substances. Pharmaceutics. 2022; 14(9):1837. https://doi.org/10.3390/pharmaceutics14091837

Chicago/Turabian Style

Schüller, Andreas, Lena Studt-Reinhold, and Joseph Strauss. 2022. "How to Completely Squeeze a Fungus—Advanced Genome Mining Tools for Novel Bioactive Substances" Pharmaceutics 14, no. 9: 1837. https://doi.org/10.3390/pharmaceutics14091837

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