Next Article in Journal
Genome Sequence Analysis of the Fungal Pathogen Fusarium graminearum Using Oxford Nanopore Technology
Next Article in Special Issue
Proline-Specific Fungal Peptidases: Genomic Analysis and Identification of Secreted DPP4 in Alkaliphilic and Alkalitolerant Fungi
Previous Article in Journal
Chronic Occupational Mold Exposure Drives Expansion of Aspergillus-Reactive Type 1 and Type 2 T-Helper Cell Responses
Previous Article in Special Issue
Enhanced Lignocellulolytic Enzyme Activities on Hardwood and Softwood during Interspecific Interactions of White- and Brown-Rot Fungi
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Trends and Applications of Omics Technologies to Functional Characterisation of Enzymes and Protein Metabolites Produced by Fungi

1
Institute for the Development of Energy for African Sustainability (IDEAS), College of Science, Engineering and Technology, University of South Africa, P.O. Box 392, UNISA, Pretoria 0001, South Africa
2
Department of Environmental Science, College of Agricultural and Environmental Science, University of South Africa, P.O. Box 392, UNISA, Pretoria 0001, South Africa
*
Author to whom correspondence should be addressed.
J. Fungi 2021, 7(9), 700; https://doi.org/10.3390/jof7090700
Submission received: 22 July 2021 / Revised: 19 August 2021 / Accepted: 23 August 2021 / Published: 27 August 2021
(This article belongs to the Special Issue Fungal Enzymes 2021)

Abstract

:
Identifying and adopting industrial applications for proteins and enzymes derived from fungi strains have been at the focal point of several studies in recent times. To facilitate such studies, it is necessary that advancements and innovation in mycological and molecular characterisation are concomitant. This review aims to provide a detailed overview of the necessary steps employed in both qualitative and quantitative research using the omics technologies that are pertinent to fungi characterisation. This stems from the understanding that data provided from the functional characterisation of fungi and their metabolites is important towards the techno-economic feasibility of large-scale production of biological products. The review further describes how the functional gaps left by genomics, internal transcribe spacer (ITS) regions are addressed by transcriptomics and the various techniques and platforms utilised, including quantitive reverse transcription polymerase chain reaction (RT-qPCR), hybridisation techniques, and RNA-seq, and the insights such data provide on the effect of environmental changes on fungal enzyme production from an expressional standpoint. The review also offers information on the many available bioinformatics tools of analysis necessary for the analysis of the overwhelming data synonymous with the omics approach to fungal characterisation.

1. Introduction

The use of fungal biological products predates this modern era; consistently, fungi have been used as a source of food, food additives, hallucinogens, pharmaceuticals (antibiotics, hormones, and immunological adjuvants), citric acid, specialty chemicals production, and derived enzymes have been used in a variety of bioprocesses including textile bleaching, alcohol fermentation, and environmental pollution remediation [1,2,3,4]. However, the isolation and identification of a desired species of fungi and its functional characterisation to determine the potential biological product that it will contribute towards domestic and commercial exploits are prerequisite stages for the commencement of any application. The precise identification of a fungal strain is crucial, especially when we consider that visual misidentification by unsuspecting individuals and resultant ingestion have often led to grave consequences, including fatalities [5,6]. Moreover, the pertinent role played by members of the fungi kingdom as major decomposers and recyclers in nature and their ubiquity in both terrestrial and aquatic environments has motivated investigations that culminate in the accepted importance of their degrading enzymes in several industries. Some fungi have been isolated and known to thrive in very extreme conditions such as arid desert conditions, low pH and high metal concentrations such as is present in some soils, mine tailings and acid mine drainages, and even in high-salt concentrations as found in marine ecosystems [7,8,9,10]. In the quest for survival by the fungi, these extreme environments have facilitated the evolution of unique enzymes that find applications in several industrial and environmental exploits. The most important enzymes studied to date are those produced by the species of Ascomycetes, Basidiomycetes, and Deuteromycetes, particularly the lignin-modifying enzymes (LMEs) such as cellulases, laccases, and peroxidases that have also proved very useful in the degradation of xenobiotic compounds due to their multi-specific activities [11,12,13].
In the past, research on fungal characterisation relied mostly on the classical mycological and microbiological approaches of bioprospecting which involved isolating wild fungal species of interest from their natural environments, cultivation under growth conditions that strived to simulate favourable conditions for proliferation using selective growth media on plates and sometimes cultivation in submerged cultures, under aseptic culture conditions [14,15,16]. However, even with advancements, not all taxa have been successfully cultured on media, and some fungi have demonstrated recalcitrance to cultivation on most current commercially available media. More recently, there have even been attempts to overcome this challenge with the use of laser-printing technology [17]. The earliest studies involving fungi used adaptations of the laser-induced forward transfer (LIFT), such as absorbing-film-assisted, laser-induced forward transfer (AFA-LIFT) for the fungus (Trichoderma). The LIFT enables the rapid but controlled transfer of minute quantities of biological samples to different substrates thinly coated to a metal surface plate with other materials such as silica. However, AFA-LIFT especially ensures this thin layer is protected from the laser irradiation by a metal film capable of absorbing the laser energy and conversion to kinetic energy; this energy actually stimulates rapid germination and/or growth of conidia as was observed in T. longibrachiatum with germination observed within a 20 h incubation period [18,19,20,21].
In general, growth and isolation are typically achieved using solid-surface culture techniques; fungal growth in liquid media is also possible, especially for the purpose of enrichment to maximise the proliferation of mycelia of a desired fungus [22]. However, submerged cultures cultivation is challenged by oxygen levels constraints for these aerobic eukaryotic cells as the mycelia will tend to grow on the liquid–air interface. Nonetheless, the latter cultivation approach has undergone several adaptations, including pseudo-submerged culture conditions, as the growth in liquid media appears to favour the collection of the extracellular enzymes produced by fungi [23].
Although the process of obtaining axenic cultures of fungi is tedious and requires several sub-culturing and strain transfers on selective media, it is unavoidable in the quest to derive biological products from a fungal strain [24]. Moreover, pure plate cultivation ensures storage and reproducibility. In the past, identification was based on observable traits, and characterisation employed morphological variations that relied on nuanced visualisation facilitated by staining techniques and microscopy, for example, light, electron, fluorescence, phase contrast, etc. [25,26]. Morphological characterisation depends greatly on comparative analysis with previously documented fungal features [27]. Even with advances in microscopy and the additional computational analyses that have improved its accuracy, morphological characterisation is still considered an imprecise method, especially when applied for differentiating members of the same genus and when, at mycelial stages, distinct features have not developed. The reliance of microscopy on an individual’s expertise further increases biases in identification [28]. Therefore, beyond phenotype-based identification, genotype studies appear to provide the best option in non-discriminatory characterisation. This review provides a synthesis of information on progress made thus far in integrating the omics approach in fungal investigation and functional characterisation of economically significant enzymes in the hope of bringing attention and encouraging routine use of these frameworks in research to boost the lagging database for fungi-derived biological products. There is a general acceptance that any progress in fungi-derived product upscaling and increased industrial product volumes can only be the result of our clear understanding and elucidation of fungal responses to environmental changes, their molecular expression patterns and mechanisms of metabolite productions and adaptations. Such knowledge will be concomitant to advances in the field of mycology and industrial applications. Moreover, it will be associated with the ease in accessibility and cheap application of enzymes and other protein metabolites employed in industrial and environmental processes. Additionally, the application of these enzymes will potentially reduce effluent production owing to its specificity in chemical reactions, and it will serve as an environmental pollution mitigation strategy. Therefore, it becomes necessary to routinely re-evaluate the status of advances made in fungal characterisation to allow researchers to identify gaps requiring improvements. The omics approach, as will be discussed, encompasses metagenomics, genomics, transcriptomics, and proteomics. Figure 1 presents an overview of the protein (enzyme) expression process and the connection to the omics approach in functional characterisation. Research thus far has been sparse, in which omics technologies have been comprehensively used in the characterisation of novel fungal enzymes. In a recent publication, Oates et al. [29] used a multi-omics approach to identify a number of potentially important CAZymes from P. putredinis NO1. A potential industrially relevant extracellular oxidase capable of cleaving β-ether linkages and releasing tricin from monocot lignin was also identified. Brenelli et al. [30] used a similar approach to identify multiple redox-active enzymes produced by the marine fungus Peniophora sp. CBMAI 1063. Using a combination of genome and secretome analyses, the authors discovered valuable extracellular CAZymes, especially lignin and polyphenols-degrading enzymes. It is hoped that this review will draw attention to this gap in research and provide insights on a systematic approach to fungal protein metabolite characterisation.

2. Relevant Metabolites Produced by Fungi

Fungi produce both primary and secondary metabolites of medical and economic significance. Pertinent to the review are three major types of metabolites (see Figure 2): the enzymes, bioactive compounds (including amino acids), and fungal ribotoxins. Although it is estimated that at least 5.1 million species of fungi exist, the Basidiomycota and Ascomycota phyla appear to have predominance and include several wood- and litter-decomposing fungi [31]. Moreover, these two phyla have thus far produced the most impactful metabolites, including enzymes that have gained industrial importance. They have shown demonstrable application in carbon recycling and therefore bioremediation, as well as usefulness in food, feed, and pharmaceutical industries [32,33,34,35,36]. The enzymes produced by fungi are unique because of their broad substrate specificity as well as their strong oxidative and hydrolytic abilities, and of utmost importance is the ease of collection since most of them are expressed extracellularly. Table 1 shows some important examples of industrially relevant fungal enzymes and the aspects of industry in which they have found application. Several authors in recent times have reviewed the applications of these enzymes [37,38,39].
There are additional enzymes that have been recently elucidated that find application in the bioremediation of recalcitrant pollutants as well as heterocyclic substrates, including aromatic hydrocarbons, halogenated biphenyl esters, and chlorinated benzenes [40]. These two newly elucidated enzymes are chloroperoxidase and the so-called aromatic or unspecific peroxygenases (APO/UPO) and are classified as either heme-thiolate peroxidases or peroxygenases with capabilities to incorporate an oxygen molecule from their hydrogen peroxide acceptor into their substrate, thereby demonstrating peroxygenase activity [41]. Moreover, there has been the identification of dye-decolourising peroxidases (DyP) in the heme peroxidase class with a preference for the oxidation of anthraquinone dyes. These enzymes show high catalytic efficiency and, in the future, will most likely play a very important role in the decolourisation of synthetic dyes [42,43].
Several secondary metabolites are produced by fungi, often referred to as bioactive compounds; most of these metabolites are associated with endophytic fungi, a group of fungi closely associated with plants. These bioactive compounds have demonstrated efficacy as antibiotics and anticancer drugs and include Penicillenols, Taxol, Clavatol, Sordaricin, Jesteron, and Javancin. Bioprospecting of these groups of fungi offers great potential for investigation and drug discovery as it is believed that only a small fraction of these compounds has been identified to date. Recently, Hoeksma et al. [44] used a combination of mycology, embryology, and chemistry to test and identify biologically active secondary metabolites from over 10,000 species of fungi using developing zebrafish embryos. The authors successfully identified 34 active compounds consisting of both therapeutically known and unexplored compounds. Molecular characterisation approaches have also recently been applied for the screening of bioactive compounds from endophytic fungi [45].
Furthermore, fungi produce a number of free amino acids that contribute nutritional value and unique tastes to food preparations. Previously, Tseng et al. [46] identified free amino acids as some of the non-volatile flavour components of the medicinal mushroom Ganoderma tsugae. A total of 17 amino acids were found to be present. Recently, Tagkouli et al. [47] cultivated fungi on agricultural by-products and used this to specifically assess the free amino acids profile of Pleurotus mushrooms. The authors identified a total of 22 free amino acids, including all the essential ones. The work by Landi et al. [48] on the Pioppino mushroom (Cyclocybe cylindracea) also illustrates the presence of these important products in basidiomycetes as their content analysis of this mushroom revealed the presence of various amino acids and their antioxidant activity.
Table 1. Economically significant enzymes produced by basidiomycetes and ascomycetes.
Table 1. Economically significant enzymes produced by basidiomycetes and ascomycetes.
Class of EnzymeSub-CategoryEnzymeIndustrial ApplicationsReferences
OxidoreductasesHeme-containing
peroxidases and peroxygenases
Lignin peroxidaseProduction of vanillin, textile dye degradation and bleaching, xenobiotic compounds degradation[37]
Manganese peroxidaseBioethanol production, dye decolourisation, degradation of rubber, distillery waste treatment[37,39]
Versatile peroxidaseNovel delignification strategies[38]
ChloroperoxidaseHalogenation and oxygenation of organic compounds, lignin cleavage[41,49]
Aromatic/Unspecific
peroxygenases
Bioremediation of recalcitrant pollutants such as polycyclic aromatic hydrocarbons, halogenated biphenyl esters, and chlorinated benzenes[40]
Dye-decolourising
peroxidases
Decolourisation of synthetic dyes[42,43]
Flavin-containing
oxidases and
dehydrogenases
Aryl-alcohol oxidaseSynthesis of high value-added chemicals, flavours, production of bio-based polymer precursors, dye decolourisation, pulp bleaching[50,51]
Copper-containing
oxidases and
monooxygenases
LaccaseDetoxifier in industries such as pulp and paper, textile, and petrochemical Heavy metal precipitation, biosensor production, food processing, drug preparation[52,53,54]
HydrolasesProteasesAlkaline, acid, and
neutral proteases
Fabrication of laundry detergents, beer haze clarification, brewing, cheese making, meat tenderisation, synthesis of bioactive peptides[33,34]
Glycoside
hydrolase
Cellulases: endoglucanases, exocellulases, and processive endoglucanasesJuice extraction and clarification, supplementation of livestock, biofuel production, laundry detergent production[35]
XylanaseHemicellulases,
endo-1,4-β-D-xylanases,
β-D-xylosidases, α-glucuronidase acetylxylan esterase,
α-L-arabinofuranosidases,
p-coumaric esterase, and
ferulic acid esterase
Juice clarification, paper bleaching, and increasing the nutritional value of animal feeds.[36,55]
Amylaseα-amylase, β-amylase, and
glucoamylase
Bread making, detergent manufacture, treatment of digestive disorders[56]
LipaseTriacylglycerolacyl hydrolasesProduction of detergents and cosmetic products, synthesis of biopolymers and biodiesel, manufacturing of food products[36]
In recent times, attention has been focused on the highly toxic extracellular ribonucleases produced by fungi because they have previously demonstrated antitumor and immunotoxin effects [57], making them useful in cancer therapy and immunotoxin synthesis. Ribotoxins exert ribonucleolytic activity on the larger molecule of RNA of the ribosome, which results in the cessation of protein synthesis and cell death through apoptosis [58]. Olombrada et al. [59] reviewed the potential biotechnological applications of these highly specific enzymes. Research completed by Li and Xia [60] showed the efficacy of ribotoxins as pest control agents either alone or conjugated with other compounds. This was used in experiments involving a recombinant form of the insecticidal ribotoxin hirsutellin A (HtA) produced from Pichia pastoris. The insecticide formulation showed dose-dependent cytotoxicity to Sf9 insect cells and insecticidal activity against Galleria mellonella larvae.
Recently, a novel family of proteins similar to ribotoxins has been identified and characterised. Landi and colleagues first characterised and purified Ageritin, the first basidiomycete-produced ribotoxin-like protein capable of inhibiting protein synthesis in vitro and releasing the α-fragment when incubated with ribosomes [61]. More recently, the same researchers identified a second ribotoxin-like protein called Ostreatin [62]. Although functionally similar to ribotoxins, these enzymes show no homology with ribotoxins from the ascomycetes family. They show more similarity to members of the ribonuclease T1 family, however. Ragucci et al. [63] recently reviewed Ageritin, focusing on its structural, biological, antipathogenic, and enzymatic characteristics. The authors also discussed some of the possible biotechnological applications of Ageritin, including its use as an immunotoxin and pest control agent, which could be achieved through gene transfer in plant cells.

3. Advances in Molecular Characterisation of Fungi

The study and application of molecular biology to fungi characterisation has rapidly gained popularity as a gold standard approach and is entrenched now as validation for most morphological characterisation [64]. Its routine adoption to the taxonomic identification of pathogenic fungi has greatly assisted in the resolution of previously unidentified fungi to species level [26,65,66]. The process begins with genomic extraction techniques, sequencing which employs specific DNA markers that allow for identification to species level; further bioinformatics analyses facilitate phylogeny establishment.
Developed in 1977, Sanger sequencing is the current standard molecular tool for identification [67]. Sanger sequencing is made possible through polymerase chain reaction (PCR), a molecular technique that enables the synthesis of a complementary DNA strand from a template using the enzyme taq polymerase [68,69]. In 1990, a ground-breaking advancement in fungal molecular characterisation was introduced with the identification of fungal nuclear ribosomal RNA (rRNA) operon primers [70]. The DNA sequences associated with the large ribosomal subunit (nrLSU-26S or 28S), the small ribosomal subunit (nrSSU-18S), and the whole internal transcribed spacer region (ITS1, 5.8S, ITS2; 650-900 bp) have since become the target region of fungal identification by the Sanger sequencing approach [64]. Different evolution rates have been observed in the ITS region, resulting in diverse levels of variation from one organism to the next. As the fastest evolving and most variable segment, the ITS region has become the golden standard for fungal identification using ITS1 and ITS4 primers [64,71,72,73]. The ITS1 and ITS4 gained popularity because they are capable of differentially identifying to significant extents members of the Ascomycetes and Basidiomycetes class of fungi. Molecular characterisation also owes its success to the rapid advancement in bioinformatics tools and the continuous addition of new species and strains to different databases that enable identification. Molecular characterisation involves the initial amplification of a conserved DNA region using a genomic DNA template and specific primers that target specific regions. Amplification is then terminated using di-deoxynucleotides. The derived sequences are matched to other previously submitted sequences in a comprehensive database. These comparisons characterise and identify through the matching of sequences to the closest relatives on the database by allocating a value range of between 0 and 100 for percentage similarities, the latter indicating an exact match [67,74]. Databases such as e-Fungi, NCBI-BLAST, FungiDB, and UNITE serve as repositories for sequences and are also crucial to molecular characterisation as it is through them that identification can be precisely achieved [75,76,77]. Nilsson et al. [78] as well as Kõljalg et al. [79] describe UNITE, a database designed for fungi molecular identification curating all public fungal sequences as well as those that have not successfully been assigned to any taxonomic lineage beyond phylum. The latter groups are assigned a unique digital object identifier (DOI) in the interim. Loeffler et al. [80] discuss the importance of reference genomes to metagenomics studies but pertinently warn of the poor integration of new sequences into the existing database as well as the cooperative interactions between different databases with the resulting lack of a credibly comprehensive database despite progress in sequence curation. The direct implication is the doubtful identification of any fungi that are analysed using only one reference database [81]. This challenge significantly impacts progress in metagenomic analysis as it hinders the pace of identification and adds the cumbersome process of further analysis on several databases to ensure that accuracy in identification is achieved. The consensus solution will be to open dialogue between database creators and seek a partnership that allows the merging of most platforms to reduce the overlap created and prevent the ‘re-invention of the wheel’, synonymous with a lack of cooperation amongst developers. This cohesion will reduce discrepancies in identification between multiple platforms and databases and accelerate progress in functional genomics.

4. Transcriptomics in Fungi Functional Characterisation

Genomic sequences provide, to a certain extent, sufficient information that facilitates the molecular characterisation of fungi. However, it is by no means adequate in giving insights into an organism’s physiology because the sequence data elucidate all genes present; however, these genes are usually not expressed at the same time or in the same manner [82]. Changes in environmental conditions significantly influence gene expression and metabolites production by causing epigenetic modifications such as DNA methylation and chromatin re-modelling [83]. This implies that to truly understand metabolite production, it is important to look in-depth at the succession of events that occur during gene expression under normal circumstances and make comparisons against varied changes in conditions. This aspect of research is referred to as transcriptomics and has also evolved over time in response to technological advances for fungal functional genomic characterisation.
The central dogma of molecular biology is premised on the transcription of genes into messenger RNA (mRNA) and the translation of the latter into peptides; it also integrates the study of the post-translational processes and the organisation of polypeptide chains into functional proteins [84,85]. This study is important in connecting genes to morphological and physiological characteristics. More recently, the focus is firmer on gene expression going from the genome to the transcriptome [86]. The field of transcriptomics bridges the gap between genes and proteins [87] and has made it possible to monitor and measure gene expression in different cells and tissues as well as the responses to different conditions at different time points. More importantly, with eukaryotic cells and their pre-disposition towards expression being directed by a complex array of genes, but also in some other instances, expression may be the consequence of single gene’s activity, while others can even be attributed to the activity of a cluster of genes [88]. Advances in our understanding of transcriptomics have enabled the identification of many previously unannotated genes, such as witnessed in the recent work of Noriega et al. [89], where transcriptomics provided insights to genes crucial to different developmental stages in the coffee berry borer Hypothenemus hampei. There are similarities in the approaches used by Noriega and colleagues to that employed by Kim et al. [90] in the identification of genes associated with a particular phenotype; in their research, the focus was on identifying genes responsible for the display of yellow colouration on leaves obtained after the γ-ray-based mutagenesis of a Cymbidium orchid. They utilised the Kyoto Encyclopedia of Genes and Genomes database to assign a total of 144,918 unigenes derived from over 25 million reads into 22 metabolic pathways. Remarkably, the RNA sequencing analysis identified 2267 differentially expressed genes between wild-type and mutant Cymbidium sp. In summary, they were able to pinpoint the alteration of chlorophyll metabolism to seven genes also observed to be involved in ion transport as well as chlorophyllase-2 production. This implied a possible link between leaf colouration in Cymbidium orchids and combined alterations in chlorophyll metabolism and ion transport.
Transcriptomics has also contributed to our understanding of white-rot fungi and the process of LMEs production. Although it is still sparsely utilised in fungi research, techniques such as quantitative reverse transcription PCR (RT-qPCR) have been found to be very useful in identifying and quantifying mRNA at any given time in biological systems [91] and can be integrated into fungal differential expression studies. It should be noted that in comparison to other PCR techniques, RT-qPCR gives accurate quantification of mRNA due to its ability to quantify in real-time. Therefore, it is very useful in investigations of gene expression resulting from different treatments. Moreover, it has other advantages, including the absence of a need for amplification, an exercise that tends to introduce bias; it also avails the user a mid-throughput analysis and, to an extent, ease of automation that allows for the processing of a relatively significant number of samples [92,93]. RT-qPCR stabilises a target mRNA through its transcription into complementary DNA (cDNA) while amplification is ongoing; there is an associated emission of fluorescence which enables real-time quantification [94]. Fernández-Fueyo et al. [95] used RT-qPCR to investigate the effect of environmental parameters of temperature and pH on gene expression and regulation in Pleurotus ostreatus. By monitoring gene expression under various conditions, the authors were able to present certain predictions on the correlation between temperature and pH in gene regulation. Previously, RT-qPCR was used to identify the peroxidase genes lip and mnp from soil fungi [96] and in studies involving peroxidase genes in ectomycorrhizal fungi [97], as well as the expressional profile of Pleurotus ostreatus laccase genes [98] and many more. More recently, Vasina et al. [99] used similar tools in the quantification of the expressional patterns of 18 peroxidase genes encoding class II peroxidases in Trametes hirsuta. Moreover, the characterisation of this multigene family enabled the design of specific primers that were used for further studies.
Although the RT-qPCR is considered fast and effective, its limitation remains its mid-throughput feature which implies that it can only process a few genes at a time. This can be a constraint, especially impactful when studying eukaryotic organisms and fungi with their tendency to have multiple gene involvement in the production of a given compound. Therefore, it limits the depth of analysis in such instances [100,101]. Moreover, a further challenge in employing RT-qPCR in fungal investigations is with the choice of a correct reference gene, which is sometimes hampered by the sharp impact that environmental/external parameters and growth conditions tend to have on gene expression [102]. This has motivated the introduction of novel, high-throughput technologies in order to ameliorate these limitations [87,103]. Hybridisation techniques such as microarrays were able to partially address these issues by increasing the amounts of genes to be analysed per run. However, prior knowledge of these genes is still required to develop complementary probes [87]. It would appear that most of the limitations highlighted previously in this review were addressed in the design and application of the RNA sequencing (RNA-seq) platform development; of most importance is the ability to analyse thousands of known or unknown genes at once without the need for a reference gene [104,105,106]. Moreover, the RNA-seq platforms have the ability to integrate analysis of post-transcriptional modifications, which is important in eukaryotic cell studies and specifically fungi, where gene fusion, alternative gene spliced transcripts and mutations/single nucleotide polymorphisms (SNPs), and changes in gene expression may occur over time. RNA sequencing platforms currently use the next-generation sequencing principle to obtain the whole transcriptome profile (type and quantity) of a cell, tissue, organ, or entire organism [107]. Although modifications have been made since its introduction, all RNA-seq platforms basically use the same principles, differing only in terms of read lengths, throughput, error, and price. Different authors have reviewed and compared these platforms for various applications [107,108,109,110]. Ravichandran et al. [111] used RNA-seq on an Illumina platform to obtain insight on the degradative ability of a white-rot fungus based on the expression of genes for these degrading enzymes. Henske et al. [112] used RNA-seq to investigate the differential expression of LMEs in the presence of different substrates. Additionally, Ma et al. [113] previously used the same tool to study fungal metabolic regulation by identifying all genes under the regulation of a given transcription factor, such as the Xyr factor presumed to be responsible for carbohydrate metabolism.
In as much as there have been some studies that have helped to deepen our understanding of the expressional patterns followed by fungi, especially with regard to LMEs, several aspects of functionality are yet to be elucidated. Most of these studies point to a need for continuation in research and in-depth studies as many more questions are raised when attempts are made that seek to probe previously held assertions [114]. It is without a doubt that these questions around expressions and translation patterns need investigations especially motivated by the potential applications that these enzymes produced by fungi represent to several industrial and environmental bioremediation pursuits. For example, Korripally et al. [115] used RNA-seq to study the regulation of LMEs gene expression. By applying whole transcriptome shotgun sequencing, they observed an increase of up to four-fold from an initial 356 genes at the inception of ligninolytic enzyme production. This link presents questions on what the possible trigger in terms of external conditions could have been that caused this sharp increase. This is further emphasised because, in the same study, they made further observations that there were at least 356 up-regulated genes and that 165 were of unknown functions. Similar thought-provoking results have been observed in other studies [116,117], which indicate the need for more transcriptomics studies in order to bring better insight to this area of research and possibly serve as a starting point for the elucidation of LMEs gene expression, as it tends to have a direct impact on the amount of enzyme produced.

Bioinformatics in Fungi Transcriptomics Studies

The evolution in transcriptomics studies and the general acceptance of RNA-seq platforms as standard for transcriptomics-based studies resulted in a change in the approach to data analysis owing to the copious amounts of data it generates. This has motivated a transition to the application of complex bioinformatics and computational tools to enable a clearer interpretation of data, accuracy, and reproducibility of results. These data analyses involve three major steps, namely: quality control of the raw reads, mapping and alignment, and quantification of reads. Similarly, further investigative research on differential expression analysis also involves each of the steps identified and several other bioinformatics tools [87,118]. This stage-gated approach in reviewing data is premised on the susceptibility of the huge amount of complex data to many variations that may arise from technical and random sources [119]. Therefore, the quality control of raw data is especially critical in ensuring the accuracy of analysis, primarily focused on tackling likely biases introduced from nucleotide composition, for example, GC-content that tends to affect differential expression analysis [120]. This problem is usually overcome with normalisation to guarantee accuracy in the inference of expression levels and other analyses, but if this is not considered for prior treatment before analysis, such an anomaly will bias final results. To date, various bioinformatics tools have been developed for RNA-seq quality control (QC). In their work, Hernández-Domínguez et al. [118] present FastQC as one of the most popular QC tools for the Illumina platform. FastQC reports on the quality of data based on reads or sequences run on its platform; it also provides the proportion of each nucleotide base in the reads. The applications and efficiency of FastQC were also reviewed by Qi et al. [121]. Other QC tools, including FASTX-Toolkit, QC-Chain, and NGS QC Toolkit, were discussed by Zhou et al. [122], who also introduced RNA-QC-chain, a novel comprehensive tool for QC which comes with the advantages of trimming, automatic rRNA detection, and contaminating species identification. Quality control not only takes place at the beginning of data analysis but also precedes each step of further data analysis.
Initial quality control is followed by mapping and alignment, where all reads are located either with respect to a reference genome or using de novo assembly. In their work, Schaarschmidt et al. [123] evaluate seven different mapping tools (bwa, CLC Genomics Workbench, HISAT2, kallisto, RSEM, salmon, and STAR) using experimental data from Arabidopsis thaliana. Interestingly, similar results are obtained, with all of them showing high reproducibility. Hernández-Domínguez et al. [118] describe three strategies that could be followed during the mapping process. When the goal is to identify new transcripts, reads are aligned with gaps to a reference genome. Tools such as STAR have been found to be best adapted for this type of mapping as they can map spliced sequences of any length [124]. When new transcripts are not the query for analysis, reads are aligned to the reference genome without gaps using tools such as RSEM. In other cases, where a reference genome is not available, reads are used for de novo assembly. Previously, Haas et al. [125] described Trinity, one of the most popular tools used for de novo assembly. Trinity is an ‘assembly-first’ tool for transcriptome reconstruction consisting of three modules, namely, Inchworm, Chrysalis, and Butterfly. Trinity is able to assemble transcriptomes by separating the data into many de Bruijn graphs that are processed separately before using parallel computing to reconstruct the transcriptome [126].
Transcript quantification is an important step in data analysis as it provides the important quantitative aspect to RNA-seq applications which allows comparison between different expression points and provides insights as to when the organisms begin to respond to externalities. Jin et al. [127] evaluated different quantification methods using tools such as TopHat, RSEM, HTSEq, and featureCounts. The authors also differentiated between alignment-based and alignment-free methods depending on the presence or absence of a reference genome. Another important step in transcript quantification is the previously mentioned normalisation of data to remove the influence of all possible biases. Normalised measures such as RPKM (reads per kilobase of exon model per million reads), FPKM (fragments per kilobase of exon model per million mapped reads), and TPM (transcripts per million) are then used to report expression values [128]. Based on the obtained data, differential expression analysis can then be carried out by comparing values from different samples. Various tools are also available for differential expression, including DESeq, Cufflinks, PoissonSeq, UpperQuartile, etc. [107,128]. Although most currently used tools have been found to be very effective for differential expression, Assefa et al. [129] reported low performance when assessing differential expression of long non-coding RNAs (lncRNAs) using 25 different pipelines. The authors correlated this with sample levels and variability, as lncRNAs are expressed at low levels and are quite variable. For fungal specific studies, Wang et al. [130] described a workflow for differential expression in fungal species using the Bioconductor package DESeq2. In recent work, Pawlik et al. [131] used RNA-seq to study differential expression in the fungus Cerrena unicolor FCL139 when grown under different lighting conditions. Mapping was completed using a reference genome, and the DESeq 2 package was also used to identify the differentially expressed genes.

5. Proteomics in Fungi Translational Characterisation

Translation of mRNA into functional proteins is the second step of the central dogma; it is considered a precursor to metabolomics studies and most cellular biochemical activities. It is also where products of economic significance are derived. Depending on the complexity and size of the organisms’ protein; the final product can be simple or complex, with the latter requiring a further post-translational modification (PTM) step of folding at the cellular or intercellular levels; without this crucial refinement, the long peptide chains that spool out during translation may not gain functionality or specificity [132]. Proteomics looks at the identification and quantification of all proteins present in a given cell, tissue, or organism [133,134]. The study of protein expression, modification, structure, and function by means of proteomics has brought progress in various fields of science and technology because proteins are usually the end-products of gene expression and the sought-after bioproducts. Amiri-Dashatan et al. [135] reviewed the application of proteomics to food technology, biomarker, and drug target identification. An in-depth review of the applications of proteomics in pharmaceuticals was also executed by Yokota [136], including the use of proteomics to study expression profiling, protein–protein interactions, and post-translational modifications. Other authors have focused their proteomics studies on fungi specifically in order to identify potential vaccines and drug targets [137,138,139]. In some of the research studies, the authors used quantitative Mass Spectrometry—Elevated Collision Energy to identify fungal proteins with no significant homology with human ones with the goal to ascertain their feasibility as vaccine candidates. Importantly, Ball et al. [140] reviewed advancements in MS-based proteomics as it relates to fungal pathogenesis and interactions between these fungi and the host. Even though the focus was on medicine and pharmacology, this work provides a benchmark for studies in industrial and environmental biotechnology.

5.1. Protein Production and Analysis

Eukaryotes, in general, have more complex machinery involved in protein production as compared to bacteria. Fungi proteins undergo post-translational modifications (PTMs) involving phosphorylation, glycosylation, acetylation, and many other processes to ensure functionality. PTMs take the form of covalent bonding modification to straight polypeptide chains, causing a change in their structural shapes, with the end result allowing these proteins to gain function. Such protein functionality imbues in fungi, for example, virulence and pathogenicity [141]. Other examples of PTMs can be observed in proteins found in the cell membrane responsible for cell-to-cell interactions or secretory functions, which is extracellularly useful [142]. PTMs may be reversible or irreversible; however, external factors may cause irreversible changes to already modified proteins, which inevitably affects their ability to function as well. It must be noted that structure and shape are very much linked to specificity and the type of biochemical interactions individual proteins can participate in.
Wang et al. [143] describe filamentous fungi as having a mature PTMs machinery, especially for glycosylation. Unlike yeasts that mainly produce glycoproteins with high mannose content, filamentous fungi use many other monosaccharides for their glycosylation processes during PTMs, making them more appropriate to produce mammalian-like proteins for pharmaceutical uses. In their work, Wang et al. [144] reviewed the evolution of glycosylation in eukaryotes. The authors differentiate between N- and O-glycosylation. It should be noted that while N-glycoproteins have glycans attached to the amide group of their asparagine residues, O-glycoproteins have this modification attached to the carboxyl group of their serine, lysine, threonine, and proline residues. Additionally, they describe glycosylation as one of the most complex PTMs, but they assert that the process is crucial in excretory proteins’ folding. This assertion is confirmed by Ramazi and Zahiri [142], who remarkably linked glycosylation, or the lack thereof, to the onset of conditions such as cancer and diabetes in humans. Previously, Goto et al. [145] had reviewed structures and functions of O-glycosylation in fungi, providing a dossier of its cellular usefulness. Most fungal secretory proteins are glycosylated through the action of O-mannosyltransferase and several other glycosyltransferases as they move from the endoplasmic reticulum to the golgi apparatus before reaching the cell exterior. These modifications have been reported to add stability and solubility to extracellular proteins [142,145,146]. PTMs therefore play an important role in enzyme production strategies, especially when the research progression strategies may involve cloning and/or artificial synthesis. In this regard, Tokmakov et al. [147] investigated the correlation between PTMs and the success of heterologous protein synthesis. Results suggested that prior identification of potential PTMs using protein sequences could predict and optimise heterologous synthesis. Proteomics technologies such as mass spectrometry (MS) are able to study these modifications with the goal of identifying their cellular location, and the resulting data are stored on different databases, including the more popularly used PhosphoGRID, PHOSIDA, PhosphoELM, and iPTMNet [133,148].
Filamentous fungi have the advantageous tendency of producing extracellular enzymes, which allow them to play their role of complex compound degradation in the environment. This has largely contributed to extracellular enzymes being desirous in industrial applications because of the minimal requirement of downstream processing for their collection [12,149,150]. Recently, Arnau et al. [151] presented strategies and challenges for the production of industrial enzymes using the extracellular machinery of fungi. The authors described the use of classical mutagenesis and screening in order to identify and/or develop mutant organisms with the ability to produce higher enzyme titres. Moreover, many other strategies, including the use of antagonistic interspecific interactions, stronger promoters, codon optimisation of gene sequences, deletion of protease-coding genes, addition of artificial transcription factors, etc., have been employed for industrial enzyme production [152,153,154].

5.2. Enzyme Production through Recombinant DNA Technology

Cellular energy conservation and utilisation efficiency often imply that enzyme syntheses have feedback mechanisms and various other factors that prevent excessive production of all metabolites, including enzymes. Moreover, some organisms that are capable of enzyme over-expression may not necessarily be amenable to manageable growth conditions in the laboratory. Sometimes, the feasible solution to these challenges is the application of recombinant technology, involving the modification of a more amenable organism by integrating DNA molecules of the desired traits to allow its over-expression in the host organism, achieved through the use of a vector [155]. Since the introduction of recombinant DNA technology in the 1970s, it has contributed greatly to the large-scale production of many important proteins, including enzymes, in fields ranging from agriculture to pharmaceuticals and cosmetics [156,157]. Over the years, different expression hosts have been developed for recombinant protein production. In their work, Tripathi and Shrivastava [158] review these expression hosts, including bacteria, mammalian cells, yeasts, insects, and transgenic plants. The authors also point out the importance of using eukaryotic hosts to produce therapeutic proteins, as these require post-translational modifications, especially glycosylation, for their efficacy. As previously discussed, PTMs play a crucial role in eukaryotic protein stability, solubility, and functionality. As such, PTMs must therefore be considered during the selection of the host for recombinant protein production. Von Schaewen et al. [159] further explain this attribute as they discuss the limitations of bacterial hosts such as E. coli for the production of recombinant eukaryotic proteins.
Besides host selection, recombinant protein production also relies greatly on the effectiveness of cloning. The expression of recombinant eukaryotic proteins involves cloning of the cDNA of interest into an appropriate expression vector, followed by its insertion into the host cell [156]. Several cloning vectors are available depending on the intended use. Various cloning methods have also been developed. Jia and Jeon [160] describe different cloning methods and their possible applications for high-throughput recombinant protein production. These include restriction enzyme-based cloning, which utilises restriction enzymes to determine and cut the beginning and the end of the inserted gene; recombination-based cloning, where a site-specific recombinase is used to make the recombinant vector without using restriction enzymes; and ligation-independent cloning, which facilitates directional cloning of any insert with no restriction enzyme, nor recombinase needed. Additionally, Nishikawa et al. [161] provided an in-depth analysis of a group of plasmid vectors used for gene manipulation in fungi collectively referred to as ‘pFungiway’. Interestingly, the core constitution of this group of vectors is the binary plasmid pCAMBIA2200, and they advised that the vector Agrobacterium tumefaciens can be used to mediate transformation. They were able to use it to successfully bring about transformation in two fungi, a basidiomycete, Flammulina velutipes, and in the pathogenic plant fungus Fusarium oxysporum.
Different vectors can be designed to introduce a recombinant gene into the selected host. Depending on the chosen host, these vectors contain elements or motifs necessary for the optimal expression of the recombinant protein. These include a promoter region, affinity tags, fluorescence tag, and many more as described by previous researchers [162,163]. Various authors have reported different vectors compatible with eukaryotic hosts. Some of the most commonly used vectors for fungi in recent times are summarised in Table 2.

5.3. Expression Analysis

Depending on the objective of a study, protein expression analysis can be completed using high-throughput methods such as MS or protein arrays technology. Alharbi [172] is of the opinion that protein array is less preferred in comparison to MS due to the huge amounts of data it generates, making its application difficult for most studies. Moreover, its dynamic range of detection is limited as a result of both background and saturation levels signals and the interferences that they pose. Its potential to analyse hundreds of antibodies–antigen interactions at once brings with it the challenge of quantifying protein concentrations of the newly introduced proteins. It follows that it would need to possess the capability not only for detection but immediate quantification of the varying concentrations. It also has the additional problem of generating false-negative results in the basic investigation of protein interactions on an array. This can be a consequence of the inevitable change in the folding of a protein structure within the microarray, thereby destroying the antibody–antigen pairing, leading to a false-negative result. The same will likely occur in investigative analysis involving enzyme–substrate interactions, as the issue is associated with space constraints. Further, these challenges are possibly linked to the strategies employed in protein microarray detection, which are either label-based or label-free. The labelling process and its use of materials such as fluorescent dyes or radioisotopes tend to alter surface characteristics of the protein and also likely introduce biases to the results. Thus, there is a favouring of the label-free approach, which uses inherent properties of the probe samples such as its mass to avoid interference to the protein molecule within the probe. For example, in the application of surface plasmon resonance (SPR) as a technique, where proteins are immobilised on the microarray using a thin layer of gold coating on the surface structure when the unlabelled protein probe is added, any changes in the angle of reflection of light are regarded to have been caused by their interaction and used for detection. However, the sensitivity and specificity in the unlabelled approach are lower as compared to the label-based approach. It is hoped that with refinements in aspects of quantitative information on protein-binding kinetics, there will be a shift towards greater use of label-free approach [173].
On the other hand, the MS application brings with it accuracy and sensitivity, the two features that have contributed to its popularity. Data obtained through analysis by LC-MS/MS, MALDI-TOF/MS, for example, enable the discovery and identification of protein biomarkers, most of which can be found on databases such as Mascot, MS-Tag, and PepProb. In certain cases, selected proteins can be analysed using low-throughput techniques such as ELISA and Western blotting, which depend on the reaction between a protein and a complementary tag [148,172]. Previously, Braitbard and colleagues [174] used ELISA to assay human proteins using specific peptides and antibodies. These techniques not only confirm the presence or absence of a given protein, but they also, to some extent, give a quantitative proportionality to the total protein content of a sample. Although not common in fungal studies, ELISA can be effective for the serological detection of fungi; in this regard, some of its first applications go back to the 1990s with works such as that of Kim et al. [175], where ELISA was used to identify white- and brown-rot fungi with their ligninolytic metabolites serving as antigens. In more recent work, Martin-Souto et al. [176] used ELISA to detect fungi from cystic fibrosis patients. Using whole protein extract from S. boydii as an antigenic extract, Scedosporium and Lomentospora fungal species were serologically detected in patients’ sera with 100% sensitivity. Similar work has also been completed using Western blotting to identify fungi, but mostly those causing diseases [177]. More research should shift towards possible modifications and adaptations of these existing methodologies as they can be employed as a benchmark for developing methods more suited to research involving enzyme elucidation targeting industrial and environmental applications.

5.4. Protein (Enzyme) Purification and Separation

As many different fungal proteins and mostly enzymes are produced extracellularly, purification is necessary to separate the enzyme(s) of interest from the rest of the secretome. Purification is a prerequisite for most analytical work on proteins. Pure proteins are crucial for the effectiveness and accuracy of subsequent studies such as structural and functional determination [178]. Moreover, pure proteins are important for applications in various industries as the presence of contaminants could result in different and unwanted reactions. In the pharmaceutical industry, for instance, protein therapeutics require high levels of purity to prevent unwanted interferences and reactions that may result in human fatalities.
Purification aims to exploit differentiating physicochemical characteristics of enzymes in order to separate them from a mixture [179]. For example, Landi and colleagues [61,62] exploited this understanding in the purification of the novel protein Ageritin by applying acid precipitation techniques followed by chromatographic separation steps. As basic proteins, the high isoelectric point of ribotoxin-like proteins (≥9.5) allows them to remain soluble under acidic conditions as most contaminants precipitate.
Various techniques have been applied for protein purification, as summarised in Table 3 with most of them using liquid chromatography. More et al. [180] reported the chromatographic purification of a laccase enzyme using anionic exchange followed by gel filtration by means of Fast Protein Liquid Chromatography (FPLC). Using this approach, purification is initially executed based on the net charge of the proteins, followed by their molecular weights. Additionally, their study implemented an ultrafiltration step that used the Amicon® system for further separation. This additional step enabled the concentration of the pure protein. In their studies, Mukhopadhyay et al. [181], Irfan et al. [182], and Carrasco et al. [183] used ammonium sulfate precipitation based on the biochemical process of salting in/out, which allows proteins to precipitate out of solution as a result of a change in ammonium sulfate ionic strength. When dealing with LMEs, careful considerations should be given to the material composition of filter panels; it is advisable to avoid reactive materials such as cellulose-based filters.
Regardless of the method used, purification is most often coupled with a separation technique for monitoring and evaluating the extent of purity. Cruz et al. [148] report that gel-based techniques remain the main separation techniques for proteins. Sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE), two-dimensional gel electrophoresis (2-DE), and two-dimensional differential gel electrophoresis (2D-DIGE) allow separation and visualisation of proteins within a mixture. Furthermore, 2-DE gives the protein profile of a sample which is of valuable use in comparative studies using databases such as the World-2DPAGE [191,192].

5.5. Structural Studies and Sequencing

After purification, especially when working with a novel protein, structural studies are critical towards function predictions [193]. X-ray crystallography and NMR are the principal techniques used to determine the 3D structure of proteins [148,194]. Although they can be used individually, Yee et al. [194] describe these techniques as being complementary. This is also shown in the work of Bryn Fenwick et al. [195], where using these techniques in synergy gives more accuracy in the structural analysis. The structures derived from these studies can be deposited in the different structure databases such as Protein Data Bank (PDB), PDBsum, or ModBase [196]. Functions can be predicted using the various approaches reviewed by Mills et al. [197]. They provided notable perspectives on functional annotations at the molecular level of uncharacterised proteins but also highlighted 3D-structure-based methods for protein function prediction, which such studies benefit from using software programs such as MolScript. The analysis and determination of structure allow for protein sequence elucidation. Miyashita et al. [198] describe protein sequencing using Edman degradation, a classical technique used to determine the amino acid sequence of proteins. Although effective in many cases, Edman degradation is limited to proteins without post-translational modifications on the N-terminal. This limitation is the reason for the focus on mass spectrometry application as the more favoured approach in protein sequencing technology, even though more novel approaches are continuously being developed [199,200]. Using the sequences derived, bioinformatics tools such as UnitProt and RefSeq can be employed to develop a clearer perspective of the protein and its attributes [201]. However, the one feature that integrates a consortium of databases makes UniProt very popular as a platform for protein sequences. Its constituent databases, UniProt Knowledgebase, UniProt Archive, UniProt Reference Clusters, and UniProt Proteomes, cater to all protein inquiries from functional information, annotations, publications, homology, and full proteome information [201]. Similarly, RefSeq gives non-redundant sequence information of proteins, including conserved regions and variations. It is important to note that as part of NCBI, RefSeq can be accessed from all NCBI tools such as BLAST [202]. McGarvey et al. [203] describe the functioning of Refseq as it relates to the mouse genome annotations. The authors present the various resources accessible through Refseq, including gene annotations, publications, nucleotide and protein records, even including new features such as the ‘Identical Proteins Report’, which reveals proteins that are identical both in length and in sequence. It is a useful study for fungal protein elucidation from a eukaryotic classification standpoint.

6. Conclusions

The introduction of omics technologies has and continues to deepen our understanding of fungi and processes involved in the production of their unique and very useful enzymes. While molecular characterisation is considered a routine and necessary procedure, cohesion and sharing of information among the different available databases would provide better identification of new fungi species and reduce redundancy in research. Moreover, innovative adaptation of existing technologies designed for medical studies may be necessary to advance enzyme and protein studies targeting the industrial and environmental sector requirements of bioproducts. The study of fungal transcriptomes has enabled us to look beyond the genome as a static concept and perceive it as a dynamic collection of genes whose expressions are dependent on several factors. Transcriptomics studies have brought to our knowledge numerous genes that impact protein production both directly and indirectly. While we continue to rely on technologies such as RNA-seq for the identification of these genes, more transcriptomics studies are necessary to further elucidate the molecular basis of protein production, especially in the case of the important LMEs. Fungal proteomics also continues to be a growing area of research with the goal of discovering enzymes and proteins that will be more impactful to industry and the environment. Finally, bioinformatics analyses have gained importance as tools that facilitate the amalgamation of functional prediction and commercial value addition in our quest for bioproducts; they are tools that will continue to carry great significance in the future of research targeting protein elucidation and applications.

Author Contributions

G.N.I. conceptualised the research; the original draft preparation was completed by G.N.I. and S.M.H. Project administration and supervision was performed by T.S.M., and funding acquisition was completed by M.T. All authors contributed to the several drafts’ corrections and editing to achieve a final draft. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the University of South Africa, the Department of Science and Technology (DST) (grant numbers DST/CON 0197/2017), and the Technology Innovation Agency (TIA) (grant number 2018/FUN/0166), both of South Africa.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors wish to acknowledge Asheal Mutungwazi for his contribution to the artwork and Weiz Nurmahomed for his critiques and contributions to the editing process.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in influencing opinions expressed in this review.

References

  1. Cerimi, K.; Akkaya, K.C.; Pohl, C.; Schmidt, B.; Neubauer, P. Fungi as source for new bio-based materials: A patent review. Fungal Biol. Biotechnol. 2019, 6, 17. [Google Scholar] [CrossRef] [Green Version]
  2. Yuvaraj, M.; Ramasamy, M. Role of Fungi in Agriculture. In Biostimulants in Plant Science; Mirmajlessi, S.M., Radhakrishnan, R., Eds.; IntechOpen: London, UK, 2020. [Google Scholar] [CrossRef]
  3. Hyde, K.D.; Xu, J.; Rapior, S.; Jeewon, R.; Lumyong, S.; Niego, A.G.T.; Abeywickrama, P.D.; Aluthmuhandiram, J.V.S.; Brahamanage, R.S.; Brooks, S.; et al. The amazing potential of fungi: 50 ways we can exploit fungi industrially. Fungal Divers 2019, 97, 1–136. [Google Scholar] [CrossRef] [Green Version]
  4. Meyer, V.; Basenko, E.Y.; Benz, J.P.; Braus, G.H.; Caddick, M.X.; Csukai, M.; De Vries, R.P.; Endy, D.; Frisvad, J.C.; Gunde-Cimerman, N.; et al. Growing a circular economy with fungal biotechnology: A white paper. Fungal Biol. Biotechnol. 2020, 7, 5. [Google Scholar] [CrossRef] [Green Version]
  5. Susca, A.; Villani, A.; Moretti, A.; Stea, G.; Logrieco, A. Identification of toxigenic fungal species associated with maize ear rot: Calmodulin as single informative gene. Int. J. Food Microbiol. 2020, 319, 108491. [Google Scholar] [CrossRef]
  6. Wennig, R.; Eyer, F.; Schaper, A.; Zilker, T.; Andresen-Streichert, H. Mushroom Poisoning. Dtsch. Arztebl. Int. 2020, 117, 701–708. [Google Scholar] [CrossRef]
  7. Kumar, A.; Chandra, R. Ligninolytic enzymes and its mechanisms for degradation of lignocellulosic waste in environment. Heliyon 2020, 6, e03170. [Google Scholar] [CrossRef]
  8. Chávez, R.; Fierro, F.; García-Rico, R.O.; Vaca, I. Filamentous fungi from extreme environments as a promising source of novel bioactive secondary metabolites. Front. Microbiol. 2015, 6, 903. [Google Scholar] [CrossRef] [Green Version]
  9. Murgia, M.; Fiamma, M.; Barac, A.; Deligios, M.; Mazzarello, V.; Paglietti, B.; Cappuccinelli, P.; Al-Qahtani, A.; Squartini, A.; Rubino, S.; et al. Biodiversity of fungi in hot desert sands. MicrobiologyOpen 2019, 8, e595. [Google Scholar] [CrossRef]
  10. Raghukumar, S.; Raghukumar, C.; Manohar, C.S. Fungi living in diverse extreme habitats of the marine environment. Kavaka 2014, 42, 145–153. [Google Scholar]
  11. Ostrem Loss, E.M.; Lee, M.-K.; Wu, M.-Y.; Martien, J. Utilization of Benzo [a] Pyrene by Aspergillus Species. Mol. Biol. Physiol. 2019, 10, 1–15. [Google Scholar]
  12. Asemoloye, M.D.; Tosi, S.; Daccò, C.; Wang, X.; Xu, S.; Marchisio, M.A.; Gao, W.; Jonathan, S.G.; Pecoraro, L. Hydrocarbon degradation and enzyme activities of Aspergillus oryzae and Mucor irregularis isolated from nigerian crude oil-polluted sites. Microorganisms 2020, 8, 1912. [Google Scholar] [CrossRef]
  13. Henn, C.; Monteiro, D.A.; Boscolo, M.; Da Silva, R.; Gomes, E. Biodegradation of atrazine and ligninolytic enzyme production by basidiomycete strains. BMC Microbiol. 2020, 20, 266. [Google Scholar] [CrossRef]
  14. Basu, S.; Bose, C.; Ojha, N.; Das, N.; Das, J.; Pal, M.; Khurana, S. Evolution of bacterial and fungal growth media. Bioinformation 2015, 11, 182–183. [Google Scholar] [CrossRef]
  15. Shi, X.X.; Qiu, H.P.; Wang, J.Y.; Zhang, Z.; Wang, Y.L.; Sun, G.C. A handy method to remove bacterial contamination from fungal cultures. PLoS ONE 2019, 14, e0224635. [Google Scholar] [CrossRef]
  16. Black, W.D. A comparison of several media types and basic techniques used to assess outdoor airborne fungi in Melbourne, Australia. bioRxiv 2020. [Google Scholar] [CrossRef]
  17. Cheptsov, V.S.; Tsypina, S.I.; Minaev, N.V.; Yusupov, V.I.; Chichkov, B.N. New microorganism isolation techniques with emphasis on laser printing. Int. J. Bioprint. 2019, 5, 156. [Google Scholar] [CrossRef]
  18. Mahmood, M.A.; Popescu, A.C. 3D Printing at Micro-Level: Laser-Induced Forward Transfer and Two-Photon Polymerization. Polymers 2021, 13, 2034. [Google Scholar] [CrossRef]
  19. Hopp, B.; Smausz, T.; Antal, Z.; Kresz, N.; Bor, Z.; Chrisey, D. Absorbing film assisted laser induced forward transfer of fungi (Trichoderma conidia). J. Appl. Phys. 2004, 96, 3478–3481. [Google Scholar] [CrossRef]
  20. Hopp, B.; Smausz, T.; Barna, N.; Vass, C.; Antal, Z.; Kredics, L.; Chrisey, D. Time-resolved study of absorbing film assisted laser induced forward transfer of Trichoderma longibrachiatum conidia. J. Phys. D Appl. Phys. 2005, 38, 833–837. [Google Scholar] [CrossRef]
  21. Hopp, B.; Smausz, T.; Nógrádi, A. Absorbing-Film Assisted Laser Induced Forward Transfer of Sensitive Biological Subjects. In Cell and Organ Printing; Ringeisen, B., Spargo, B., Wu, P., Eds.; Springer: Dordrecht, The Netherlands, 2010. [Google Scholar] [CrossRef]
  22. Nevalainen, H.; Kautto, L.; Te’o, J. Methods for Isolation and Cultivation of Filamentous Fungi. In Environmental Microbiology: Methods and Protocols; Paulsen, I.T., Holmes, A.J., Eds.; Springer Science+Business Media: Berlin, Germany, 2014; Volume 1096, pp. 2–16. ISBN 9781493990047. [Google Scholar]
  23. Ijoma, G.N.; Selvarajan, R.; Oyourou, J.N.; Sibanda, T.; Matambo, T.; Monanga, A.; Mkansi, K. Exploring the application of biostimulation strategy for bacteria in the bioremediation of industrial effluent. Ann. Microbiol. 2019, 69, 541–551. [Google Scholar] [CrossRef]
  24. De Vero, L.; Boniotti, M.B.; Budroni, M.; Buzzini, P.; Cassanelli, S.; Comunian, R.; Gullo, M.; Logrieco, A.F.; Mannazzu, I.; Musumeci, R.; et al. Preservation, characterization and exploitation of microbial biodiversity: The perspective of the italian network of culture collections. Microorganisms 2019, 7, 685. [Google Scholar] [CrossRef] [Green Version]
  25. Toledo, A.V.; Simurro, M.E.; Balatti, P.A. Morphological and molecular characterization of a fungus, Hirsutella sp., isolated from planthoppers and psocids in Argentina. J. Insect Sci. 2013, 13, 18. [Google Scholar] [CrossRef] [Green Version]
  26. Alsohaili, S.A.; Bani-Hasan, B.M. Morphological and molecular identification of fungi isolated from different environmental sources in the northern eastern Desert of Jordan. Jordan J. Biol. Sci. 2018, 11, 329–337. [Google Scholar]
  27. Rachmania, M.; Oetari, A.; Fitri, R.; Susetyo-Salim, T.; Sjamsuridzal, W. Isolation and morphological characterization of fungi from deteriorated old Chinese manuscripts from Central Library Universitas Indonesia. AIP Conf. Proc. 2018, 2023, 020144. [Google Scholar] [CrossRef]
  28. Lutzoni, F.; Kauff, F.; Cox, C.J.; McLaughlin, D.; Celio, G.; Dentinger, B.; Padamsee, M.; Hibbert, D.; James, T.Y.; Baloch, E.; et al. Assembling the fungal tree of life: Progress, classification, and evolution of subcellular traits. Am. J. Bot. 2004, 91, 1446–1480. [Google Scholar] [CrossRef]
  29. Oates, N.C.; Abood, A.; Schirmacher, A.M.; Alessi, A.M.; Bird, S.M.; Bennett, J.P.; Leadbeater, D.R.; Li, Y.; Dowle, A.A.; Liu, S.; et al. A multi-omics approach to lignocellulolytic enzyme discovery reveals a new ligninase activity from Parascedosporium putredinis NO1. Proc. Natl. Acad. Sci. USA 2021, 118, e2008888118. [Google Scholar] [CrossRef]
  30. Brenelli, L.B.; Persinoti, G.F.; Cairo, J.P.L.F.; Liberato, M.V.; Gonçalves, T.A.; Otero, I.V.R.; Mainardi, P.H.; Felby, C.; Sette, L.D.; Squina, F.M. Novel redox-active enzymes for ligninolytic applications revealed from multiomics analyses of Peniophora sp. CBMAI 1063, a laccase hyper-producer strain. Sci. Rep. 2019, 9, 17564. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Blackwell, M. The fungi: 1, 2, 3 … 5.1 million species? Am. J. Bot. 2011, 98, 426–438. [Google Scholar] [CrossRef]
  32. Srilakshmi, J.; Madhavi, J.; Lavanya, S.; Ammani, K. Commercial Potential of Fungal Protease: Past, Present and Future Prospects. J. Pharm. Chem. Biol. Sci. 2015, 2, 218–234. [Google Scholar]
  33. de Souza, P.M.; de Assis Bittencourt, M.L.; Caprara, C.C.; de Freitas, M.; de Almeida, R.P.C.; Silveira, D.; Fonseca, Y.M.; Filho, E.X.F.; Pessoa Junior, A.; Magalhães, P.O. A biotechnology perspective of fungal proteases. Braz. J. Microbiol. 2015, 46, 337–346. [Google Scholar] [CrossRef] [Green Version]
  34. da Silva, R.R. Bacterial and Fungal Proteolytic Enzymes: Production, Catalysis and Potential Applications. Appl. Biochem. Biotechnol. 2017, 183, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Ahmed, A.; Bibi, A. Fungal Cellulase; Production and Applications: Minireview. LIFE Int. J. Health Life-Sci. 2018, 4, 19–36. [Google Scholar] [CrossRef] [Green Version]
  36. McKelvey, S.M.; Murphy, R.A. Biotechnological Use of Fungal Enzymes. In Fungi: Biology and Applications; Kavanagh, K., Ed.; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2018; pp. 201–225. ISBN 9781119976950. [Google Scholar]
  37. Maciel, M.J.M.; Castro e Silva, A.; Ribeiro, H.C.T. Industrial and biotechnological applications of ligninolytic enzymes of the basidiomycota: A review. Electron. J. Biotechnol. 2010, 13, 1–13. [Google Scholar] [CrossRef]
  38. Kong, W.; Fu, X.; Wang, L.; Alhujaily, A.; Zhang, J.; Ma, F.; Zhang, X.; Yu, H. A novel and efficient fungal delignification strategy based on versatile peroxidase for lignocellulose bioconversion. Biotechnol. Biofuels 2017, 10, 218. [Google Scholar] [CrossRef] [Green Version]
  39. Chowdhary, P.; Shukla, G.; Raj, G.; Ferreira, L.F.R.; Bharagava, R.N. Microbial manganese peroxidase: A ligninolytic enzyme and its ample opportunities in research. SN Appl. Sci. 2019, 1, 45. [Google Scholar] [CrossRef] [Green Version]
  40. Karich, A.; Ullrich, R.; Scheibner, K.; Hofrichter, M. Fungal unspecific peroxygenases oxidize the majority of organic EPA priority pollutants. Front. Microbiol. 2017, 8, 1463. [Google Scholar] [CrossRef]
  41. Schmidt-Dannert, C. Biocatalytic Portfolio of Basidiomycota Claudia. Curr. Opin. Chem. Biol. 2016, 31, 40–49. [Google Scholar] [CrossRef] [Green Version]
  42. Duan, Z.; Shen, R.; Liu, B.; Yao, M.; Jia, R. Comprehensive investigation of a dye-decolorizing peroxidase and a manganese peroxidase from Irpex lacteus F17, a lignin-degrading basidiomycete. AMB Express 2018, 8, 119. [Google Scholar] [CrossRef]
  43. Qin, X.; Luo, H.; Zhang, X.; Yao, B.; Ma, F.; Su, X. Dye-decolorizing peroxidases in Irpex lacteus combining the catalytic properties of heme peroxidases and laccase play important roles in ligninolytic system. Biotechnol. Biofuels 2018, 11, 302. [Google Scholar] [CrossRef]
  44. Hoeksma, J.; Misset, T.; Wever, C.; Kemmink, J.; Kruijtzer, J.; Versluis, K.; Liskamp, R.M.J.; Boons, G.J.; Heck, A.J.R.; Boekhout, T.; et al. A new perspective on fungal metabolites: Identification of bioactive compounds from fungi using zebrafish embryogenesis as read-out. Sci. Rep. 2019, 9, 17546. [Google Scholar] [CrossRef] [Green Version]
  45. Vasundhara, M.; Kumar, A.; Reddy, M.S. Molecular approaches to screen bioactive compounds from endophytic fungi. Front. Microbiol. 2016, 7, 1774. [Google Scholar] [CrossRef] [Green Version]
  46. Tseng, Y.H.; Lee, Y.L.; Li, R.C.; Mau, J.L. Non-volatile flavour components of Ganoderma tsugae. Food Chem. 2005, 90, 409–415. [Google Scholar] [CrossRef]
  47. Tagkouli, D.; Kaliora, A.; Bekiaris, G.; Koutrotsios, G.; Christea, M.; Zervakis, G.I.; Kalogeropoulos, N. Free Amino Acids in Three Pleurotus Species Cultivated on Agricultural and Agro-Industrial By-Products. Molecules 2020, 25, 4015. [Google Scholar] [CrossRef] [PubMed]
  48. Landi, N.; Pacifico, S.; Ragucci, S.; Di Giuseppe, A.M.A.; Iannuzi, F.; Zarreli, A.; Piccolella, S.; Di Maro, A. Pioppino mushroom in the Southern Italy: An undervalued source of nutrients and bioactive compounds. J. Sci. Food Agric. 2017, 97, 5388–5397. [Google Scholar] [CrossRef] [PubMed]
  49. Ortiz-Bermúdez, P.; Srebotnik, E.; Hammel, K.E. Chlorination and cleavage of lignin structures by fungal chloroperoxidases. Appl. Environ. Microbiol. 2003, 69, 5015–5018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Hernández-Ortega, A.; Ferreira, P.; Martínez, A.T. Fungal aryl-alcohol oxidase: A peroxide-producing flavoenzyme involved in lignin degradation. Appl. Microbiol. Biotechnol. 2012, 93, 1395–1410. [Google Scholar] [CrossRef]
  51. Urlacher, V.B.; Koschorreck, K. Pecularities and applications of aryl-alcohol oxidases from fungi. Appl. Microbiol. Biotechnol. 2021, 105, 4111–4126. [Google Scholar] [CrossRef]
  52. Upadhyay, P.; Shrivastava, R.; Agrawal, P.K. Bioprospecting and biotechnological applications of fungal laccase. 3 Biotech 2016, 6, 15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Ahmadi Khozani, M.; Emtiazi, G.; Aghaei, S.S.; Ghasemi, S.M.; Zolfaghari, M.R. Application of fungal laccase for heavy metals precipitation using tannin as a natural mediator. Int. J. Environ. Sci. Technol. 2020, 18, 2335–2344. [Google Scholar] [CrossRef]
  54. Senthivelan, T.; Kanagaraj, J.; Panda, R.C. Recent trends in fungal laccase for various industrial applications: An eco-friendly approach—A review. Biotechnol. Bioprocess Eng. 2016, 21, 19–38. [Google Scholar] [CrossRef]
  55. Walia, A.; Guleria, S.; Mehta, P.; Chauhan, A.; Parkash, J. Microbial xylanases and their industrial application in pulp and paper biobleaching: A review. 3 Biotech 2017, 7, 11. [Google Scholar] [CrossRef] [Green Version]
  56. Carreras-Sangra, N.; Alvarez-Garcia, E.; Herrero-Galan, E.; Tome, J.; Lacadena, J.; Alegre-Cebollada, J.; Onaderra, M.; Gavilanes, J.; Martinez-del-Pozo, A. The Therapeutic Potential of Fungal Ribotoxins. Curr. Pharm. Biotechnol. 2008, 9, 153–160. [Google Scholar] [CrossRef]
  57. Saranraj, P.; Stella, D. Fungal Amylase-A Review. Int. J. Microbiol. Res. 2013, 4, 203–211. [Google Scholar] [CrossRef]
  58. Lacadena, J.; Álvarez-García, E.; Carreras-Sangrà, N.; Herrero-Galán, E.; Alegre-Cebollada, J.; García-Ortega, L.; Oñaderra, M.; Gavilanes, J.G.; Del Pozo, Á.M. Fungal ribotoxins: Molecular dissection of a family of natural killers. FEMS Microbiol. Rev. 2007, 31, 212–237. [Google Scholar] [CrossRef] [Green Version]
  59. Olombrada, M.; Lázaro-Gorines, R.; López-Rodríguez, J.C.; Martínez-Del-Pozo, Á.; Oñaderra, M.; Maestro-López, M.; Lacadena, J.; Gavilanes, J.G.; García-Ortega, L. Fungal ribotoxins: A review of potential biotechnological applications. Toxins 2017, 9, 71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Li, H.; Xia, Y. High cell density fed-batch production of insecticidal recombinant ribotoxin hirsutellin A from Pichia pastoris. Microb. Cell Fact. 2018, 17, 145. [Google Scholar] [CrossRef]
  61. Landi, N.; Pacifico, S.; Ragucci, S.; Iglesias, R.; Piccolella, S.; Amici, A.; Di Giuseppe, A.M.A.; Di Maro, A. Purification, characterization and cytotoxicity assessment of Ageritin: The first ribotoxin from the basidiomycete mushroom Agrocybe aegerita. Biochim. Biophys. Acta Gen. Subj. 2017, 1861, 1113–1121. [Google Scholar] [CrossRef]
  62. Landi, N.; Ragucci, S.; Russo, R.; Valletta, M.; Pizzo, E.; Ferreras, J.M.; Di Maro, A. The ribotoxin-like protein Ostreatin from Pleurotus ostreatus fruiting bodies: Confirmation of a novel ribonuclease family expressed in basidiomycetes. Int. J. Biol. Macromol. 2020, 161, 1329–1336. [Google Scholar] [CrossRef] [PubMed]
  63. Ragucci, S.; Landi, N.; Russo, R.; Valletta, M.; Pedone, P.V.; Chambery, A.; Di Maro, A. Ageritin from pioppino mushroom: The prototype of ribotoxin-like proteins, a novel family of specific ribonucleases in edible mushrooms. Toxins 2021, 13, 263. [Google Scholar] [CrossRef] [PubMed]
  64. Raja, H.A.; Miller, A.N.; Pearce, C.J.; Oberlies, N.H. Fungal Identification Using Molecular Tools: A Primer for the Natural Products Research Community. J. Nat. Prod. 2017, 80, 756–770. [Google Scholar] [CrossRef]
  65. Bechem, E.T.; Afanga, Y.A. Morphological and molecular identification of fungi associated with corm rot and blight symptoms on plantain (Musa paradisiaca) in macro-propagators. Int. J. Biol. Chem. Sci. 2017, 11, 2793. [Google Scholar] [CrossRef] [Green Version]
  66. Ab Majid, A.H.; Zahran, Z.; Abd Rahim, A.H.; Ismail, N.A.; Abdul Rahman, W.; Mohammad Zubairi, K.S.; Dieng, H.; Satho, T. Morphological and molecular characterization of fungus isolated from tropical bed bugs in Northern Peninsular Malaysia, Cimex hemipterus (Hemiptera: Cimicidae). Asian Pac. J. Trop. Biomed. 2015, 5, 707–713. [Google Scholar] [CrossRef]
  67. Slatko, B.E.; Gardner, A.F.; Ausubel, F.M. Overview of Next Generation Sequencing technologies. Curr. Protoc. Mol. Biol. 2018, 122, e59. [Google Scholar] [CrossRef] [PubMed]
  68. Crossley, B.M.; Bai, J.; Glaser, A.; Maes, R.; Porter, E.; Killian, M.L.; Clement, T.; Toohey-Kurth, K. Guidelines for Sanger sequencing and molecular assay monitoring. J. Vet. Diagn. Investig. 2020, 32, 767–775. [Google Scholar] [CrossRef]
  69. Valones, M.A.A.; Guimarães, R.L.; Brandão, L.A.C.; De Souza, P.R.E.; De Albuquerque, T.C.; Crovela, S. Principles and applications of polymerase chain reaction in medical diagnostic fields: A review. Braz. J. Microbiol. 2009, 40, 1–11. [Google Scholar] [CrossRef] [Green Version]
  70. White, T.J.; Bruns, T.; Lee, S.; Taylor, J. Amplification and Direct Sequencing of Fungal Ribosomal Rna Genes for Phylogenetics. In PCR Protocols: A Guide to Methods and Applications; Innis, M.A., Gelfand, D.H., Sninsky, J.J., White, T.J., Eds.; Academic Press, Inc.: New York, NY, USA, 1990; pp. 315–322. [Google Scholar]
  71. Hoang, M.T.V.; Irinyi, L.; Chen, S.C.A.; Sorrell, T.C.; Meyer, W.; Arabatzis, M.; Arthur, I.; Cano-Lira, J.F.; Cardinali, G.; Castañón, L.R.; et al. Dual DNA barcoding for the molecular identification of the agents of invasive fungal infections. Front. Microbiol. 2019, 10, 1647. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Cadez, N.; Raspor, P.; De Cock, A.W.A.M.; Boekhout, T.; Smith, M.T. Molecular identification and genetic diversity within species of the genera Hanseniaspora and Kloeckera. FEMS Yeast Res. 2002, 1, 279–289. [Google Scholar] [CrossRef] [PubMed]
  73. Schoch, C.L.; Seifert, K.A.; Huhndorf, S.; Robert, V.; Spouge, J.L.; Levesque, C.A.; Chen, W.; Consortium, F.B. Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi. Proc. Natl. Acad. Sci. USA 2012, 109, 6241–6246. [Google Scholar] [CrossRef] [Green Version]
  74. Totomoch-Serra, A.; Marquez, M.F.; Cervantes-Barragán, D.E. Sanger sequencing as a first-line approach for molecular diagnosis of Andersen-Tawil syndrome. F1000Research 2017, 6, 1016. [Google Scholar] [CrossRef] [Green Version]
  75. Hedeler, C.; Wong, H.M.; Cornell, M.J.; Alam, I.; Soanes, D.M.; Rattray, M.; Hubbard, S.J.; Talbot, N.J.; Oliver, S.G.; Paton, N.W. e-Fungi: A data resource for comparative analysis of fungal genomes. BMC Genom. 2007, 8, 426. [Google Scholar] [CrossRef]
  76. Federhen, S. The NCBI Taxonomy database. Nucleic Acids Res. 2012, 40, D136–D143. [Google Scholar] [CrossRef] [Green Version]
  77. Meena, B.; Anburajan, L.; Vinithkumar, N.V.; Kirubagaran, R.; Dharani, G. Molecular characterization, phylogenetic and in silico sequence analysis data of trehalose biosynthesis genes; otsA and otsB from the deep sea halophilic actinobacteria, Streptomyces qinglanensis NIOT-DSA03. Data Brief 2021, 35, 106727. [Google Scholar] [CrossRef]
  78. Nilsson, R.H.; Larsson, K.H.; Taylor, A.F.S.; Bengtsson-Palme, J.; Jeppesen, T.S.; Schigel, D.; Kennedy, P.; Picard, K.; Glöckner, F.O.; Tedersoo, L.; et al. The UNITE database for molecular identification of fungi: Handling dark taxa and parallel taxonomic classifications. Nucleic Acids Res. 2019, 47, D259–D264. [Google Scholar] [CrossRef] [PubMed]
  79. Kõljalg, U.; Abarenkov, K.; Nilsson, R.H.; Larsson, K.-H.; Taylor, A.F.S. The UNITE Database for Molecular Identification and for Communicating Fungal Species. Biodivers. Inf. Sci. Stand. 2019, 3, e37402. [Google Scholar] [CrossRef]
  80. Loeffler, C.; Karlsberg, A.; Martin, L.S.; Eskin, E.; Koslicki, D.; Mangul, S. Improving the usability and comprehensiveness of microbial databases. BMC Biol. 2020, 18, 37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Lücking, R.; Aime, M.C.; Robbertse, B.; Miller, A.N.; Ariyawansa, H.A.; Aoki, T.; Cardinali, G.; Crous, P.W.; Druzhinina, I.S.; Geiser, D.M.; et al. Unambiguous identification of fungi: Where do we stand and how accurate and precise is fungal DNA barcoding? IMA Fungus 2020, 11, 14. [Google Scholar] [CrossRef] [PubMed]
  82. Finotello, F.; Di Camillo, B. Measuring differential gene expression with RNA-seq: Challenges and strategies for data analysis. Brief. Funct. Genom. 2015, 14, 130–142. [Google Scholar] [CrossRef]
  83. Choi, J.K.; Kim, S.C. Environmental effects on gene expression phenotype have regional biases in the human genome. Genetics 2007, 175, 1607–1613. [Google Scholar] [CrossRef] [Green Version]
  84. Guo, J. Transcription: The epicenter of gene expression. J. Zhejiang Univ. B Biomed. Biotechnol. 2014, 15, 409–411. [Google Scholar] [CrossRef] [Green Version]
  85. Koonin, E.V. Why the Central Dogma: On the nature of the great biological exclusion principle. Biol. Direct 2015, 10, 52. [Google Scholar] [CrossRef] [Green Version]
  86. Manzoni, C.; Kia, D.A.; Vandrovcova, J.; Hardy, J.; Wood, N.W.; Lewis, P.A.; Ferrari, R. Genome, transcriptome and proteome: The rise of omics data and their integration in biomedical sciences. Brief. Bioinform. 2018, 19, 286–302. [Google Scholar] [CrossRef] [PubMed]
  87. Lowe, R.; Shirley, N.; Bleackley, M.; Dolan, S.; Shafee, T. Transcriptomics technologies. PLoS Comput. Biol. 2017, 13, e1005457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Elizondo, L.; Jafar-Nejad, P.; Clewing, J.; Boerkoel, C. Gene Clusters, Molecular Evolution and Disease: A Speculation. Curr. Genom. 2009, 10, 64–75. [Google Scholar] [CrossRef] [PubMed]
  89. Noriega, D.D.; Arias, P.L.; Barbosa, H.R.; Arraes, F.B.M.; Ossa, G.A.; Villegas, B.; Coelho, R.R.; Albuquerque, E.V.S.; Togawa, R.C.; Grynberg, P.; et al. Transcriptome and gene expression analysis of three developmental stages of the coffee berry borer, Hypothenemus hampei. Sci. Rep. 2019, 9, 12804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Kim, S.H.; Kim, S.W.; Lim, G.H.; Lyu, J.I.; Choi, H.I.; Jo, Y.D.; Kang, S.Y.; Kang, B.C.; Kim, J.B. Transcriptome analysis to identify candidate genes associated with the yellow-leaf phenotype of a Cymbidium mutant generated by γ-irradiation. PLoS ONE 2020, 15, e0228078. [Google Scholar] [CrossRef] [PubMed]
  91. Pombo, M.A.; Zheng, Y.; Fei, Z.; Martin, G.B.; Rosli, H.G. Use of RNA-seq data to identify and validate RT-qPCR reference genes for studying the tomato-Pseudomonas pathosystem. Sci. Rep. 2017, 7, 44905. [Google Scholar] [CrossRef] [PubMed]
  92. Adams, G. A beginner’s guide to RT-PCR, qPCR and RT-qPCR. Biochemist 2020, 42, 48–53. [Google Scholar] [CrossRef]
  93. Bleve, G.; Rizzotti, L.; Dellaglio, F.; Torriani, S. Development of reverse transcription (RT)-PCR and real-time RT-PCR assays for rapid detection and quantification of viable yeasts and molds contaminating yogurts and pasteurized food products. Appl. Environ. Microbiol. 2003, 69, 4116–4122. [Google Scholar] [CrossRef] [Green Version]
  94. Arya, M.; Shergill, I.S.; Williamson, M.; Gommersall, L.; Arya, N.; Patel, H.R. Basic principles of real-time quantitative PCR. Expert Rev. Mol. Diagn. 2005, 5, 209–219. [Google Scholar] [CrossRef]
  95. Fernández-Fueyo, E.; Castanera, R.; Ruiz-dueñas, F.J.; López-lucendo, M.F.; Ramírez, L.; Pisabarro, A.G.; Martínez, A.T. Ligninolytic peroxidase gene expression by Pleurotus ostreatus: Differential regulation in lignocellulose medium and effect of temperature and pH. Fungal Genet. Biol. 2014, 72, 150–161. [Google Scholar] [CrossRef] [Green Version]
  96. Stuardo, M.; Vásquez, M.; Vicuña, R.; González, B. Molecular approach for analysis of model fungal genes encoding ligninolytic peroxidases in wood-decaying soil systems. Lett. Appl. Microbiol. 2004, 38, 43–49. [Google Scholar] [CrossRef] [Green Version]
  97. Chen, D.M.; Taylor, A.F.S.; Burke, R.M.; Cairney, J.W.G. Identification of genes for lignin peroxidases and manganese peroxidases in ectomycorrhizal fungi. New Phytol. 2001, 152, 151–158. [Google Scholar] [CrossRef]
  98. Castanera, R.; Pérez, G.; Omarini, A.; Alfaro, M.; Pisabarro, A.G.; Faraco, V.; Amore, A.; Ramírez, L. Transcriptional and enzymatic profiling of Pleurotus ostreatus laccase genes in submerged and solid-state fermentation cultures. Appl. Environ. Microbiol. 2012, 78, 4037–4045. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Vasina, D.V.; Moiseenko, K.V.; Fedorova, T.V.; Tyazhelova, T.V. Lignin-degrading peroxidases in white-rot fungus Trametes hirsuta 072. Absolute expression quantification of full multigene family. PLoS ONE 2017, 12, e0173813. [Google Scholar] [CrossRef] [PubMed]
  100. Smith, C.J.; Osborn, A.M. Advantages and limitations of quantitative PCR (Q-PCR)-based approaches in microbial ecology. FEMS Microbiol. Ecol. 2009, 67, 6–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Nonis, A.; De Nardi, B.; Nonis, A. Choosing between RT-qPCR and RNA-seq: A back-of-the-envelope estimate towards the definition of the break-even-point. Anal. Bioanal. Chem. 2014, 406, 3533–3536. [Google Scholar] [CrossRef]
  102. Castanera, R.; López-Varas, L.; Pisabarro, A.G.; Ramíre, L. Validation of reference genes for transcriptional analyses in Pleurotus ostreatus by using reverse transcription-quantitative PCR. Appl. Environ. Microbiol. 2015, 81, 4120–4129. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Moyerbrailean, G.A.; Davis, G.O.; Harvey, C.T.; Watza, D.; Wen, X.; Pique-Regi, R.; Luca, F. A high-throughput RNA-seq approach to profile transcriptional responses. Sci. Rep. 2015, 5, 14976. [Google Scholar] [CrossRef] [Green Version]
  104. Costa-Silva, J.; Domingues, D.; Lopes, F.M. RNA-Seq differential expression analysis: An extended review and a software tool. PLoS ONE 2017, 12, e0190152. [Google Scholar] [CrossRef] [Green Version]
  105. Kukurba, K.R.; Montgomery, S.B. DNA sequencing and analysis. Cold Spring Harb. Protoc. 2015, 2015, 951–969. [Google Scholar] [CrossRef] [Green Version]
  106. Wang, Z.; Gerstein, M.; Snyder, M. RNA-Seq: A revolutionary tool for transcriptomics. Nat. Rev. Genet. 2009, 10, 57–63. [Google Scholar] [CrossRef]
  107. Jazayeri, S.M.; Melgarejo Muñoz, L.M.; Romero, H.M. RNA-seq: A glance at technologies and methodologies. Acta Biol. Colomb. 2015, 20, 23–35. [Google Scholar] [CrossRef]
  108. Loman, N.J.; Misra, R.V.; Dallman, T.J.; Constantinidou, C.; Gharbia, S.E.; Wain, J.; Pallen, M.J. Performance comparison of benchtop high-throughput sequencing platforms. Nat. Biotechnol. 2012, 30, 434–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Quail, M.A.; Smith, M.; Coupland, P.; Otto, T.D.; Harris, S.R.; Connor, T.R.; Bertoni, A.; Swerdlow, H.P.; Gu, Y. A tale of three next generation sequencing platforms: Comparison of Ion Torrent, Pacific Biosciences and Illumina MiSeq sequencers. BMC Genom. 2012, 13, 341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Wang, Y.J.; Schug, J.; Lin, J.; Wang, Z.; Kossenkov, A.; Kaestner, K.H. Comparative analysis of commercially available single-cell RNA sequencing platforms for their performance in complex human tissues. bioRxiv 2019, 541433. [Google Scholar] [CrossRef] [Green Version]
  111. Ravichandran, A.; Kolte, A.P.; Dhali, A.; Gopinath, S.M.; Sridhar, M. Transcriptomic analysis of Lentinus squarrosulus provide insights into its biodegradation ability. bioRxiv 2020. [Google Scholar] [CrossRef]
  112. Henske, J.K.; Springer, S.D.; O’Malley, M.A.; Butler, A. Substrate-based differential expression analysis reveals control of biomass degrading enzymes in Pycnoporus cinnabarinus. Biochem. Eng. J. 2018, 130, 83–89. [Google Scholar] [CrossRef]
  113. Ma, L.; Chen, L.; Zhang, L.; Zou, G.; Liu, R.; Jiang, Y.; Zhou, Z. RNA Sequencing Reveals Xyr1 as a Transcription Factor Regulating Gene Expression beyond Carbohydrate Metabolism. Biomed. Res. Int. 2016, 2016, 4841756. [Google Scholar] [CrossRef] [PubMed]
  114. Xu, H.; Guo, M.Y.; Gao, Y.H.; Bai, X.H.; Zhou, X.W. Expression and characteristics of manganese peroxidase from Ganoderma lucidum in Pichia pastoris and its application in the degradation of four dyes and phenol. BMC Biotechnol. 2017, 17, 19. [Google Scholar] [CrossRef] [Green Version]
  115. Korripally, P.; Hunt, C.G.; Houtman, C.J.; Jones, D.C.; Kitin, P.J.; Cullen, D.; Hammel, K.E. Regulation of gene expression during the onset of ligninolytic oxidation by Phanerochaete chrysosporium on spruce wood. Appl. Environ. Microbiol. 2015, 81, 7802–7812. [Google Scholar] [CrossRef] [Green Version]
  116. Liu, Z.H.; Le, R.K.; Kosa, M.; Yang, B.; Yuan, J.; Ragauskas, A.J. Identifying and creating pathways to improve biological lignin valorization. Renew. Sustain. Energy Rev. 2019, 105, 349–362. [Google Scholar] [CrossRef]
  117. Minami, M.; Suzuki, K.; Shimizu, A.; Hongo, T.; Sakamoto, T.; Ohyama, N.; Kitaura, H.; Kusaka, A.; Iwama, K.; Irie, T. Changes in the gene expression of the white rot fungus Phanerochaete chrysosporium due to the addition of atropine. Biosci. Biotechnol. Biochem. 2009, 73, 1722–1731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Hernández-Domínguez, E.M.; Castillo-ortega, L.S.; García-esquivel, Y.; Mandujano-gonzález, V.; Díaz-godínez, G.; Álvarez-cervantes, J. Bioinformatics as a Tool for the Structural and Evolutionary Analysis of Proteins. In Computational Biology and Chemistry; Behzadi, P., Bernabò, N., Eds.; IntechOpen: London, UK, 2019. [Google Scholar] [CrossRef] [Green Version]
  119. Merino, G.A.; Fresno, C.; Netto, F.; Netto, E.D.; Pratto, L.; Fernández, E.A. The impact of quality control in RNA-seq experiments. J. Phys. Conf. Ser. 2016, 705, 012003. [Google Scholar] [CrossRef] [Green Version]
  120. Wang, L.; Wang, S.; Li, W. RSeQC: Quality control of RNA-seq experiments. Bioinformatics 2012, 28, 2184–2185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Qi, W.; Schlapbach, R.; Rehrauer, H. RNA-Seq Data Analysis: From Raw Data Quality Control to Differential Expression Analysis. Methods Mol. Biol. 2017, 1669, 295–307. [Google Scholar] [CrossRef]
  122. Zhou, Q.; Su, X.; Jing, G.; Chen, S.; Ning, K. RNA-QC-chain: Comprehensive and fast quality control for RNA-Seq data. BMC Genom. 2018, 19, 144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Schaarschmidt, S.; Fischer, A.; Zuther, E.; Hincha, D.K. Evaluation of Seven Di ff erent RNA-Seq Alignment Tools Based on Experimental Data from the Model Plant Arabidopsis thaliana. Int. J. Mol. Sci. 2020, 21, 1720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Dobin, A.; Gingeras, T.R. Optimizing RNA-Seq Mapping with STAR. In Data Mining Techniques for the Life Sciences; Carugo, O., Eisenhaber, F., Eds.; Springer Science+Business Media: New York, NY, USA, 2016; Volume 1415, pp. 245–262. ISBN 9781493935727. [Google Scholar]
  125. Haas, B.J.; Papanicolaou, A.; Yassour, M.; Grabherr, M.; Philip, D.; Bowden, J.; Couger, M.B.; Eccles, D.; Li, B.; Macmanes, M.D.; et al. Reference generation and analysis with Trinity. Nat. Protoc. 2014, 8, 1494–1512. [Google Scholar] [CrossRef]
  126. Grabherr, M.G.; Haas, B.J.; Yassour, M.; Levin, J.Z.; Thompson, D.A.; Amit, I.; Adiconis, X.; Fan, L.; Raychowdhury, R.; Zeng, Q.; et al. Full-length transcriptome assembly from RNA-Seq data without a reference genome. Nat. Biotechnol. 2011, 29, 644–652. [Google Scholar] [CrossRef] [Green Version]
  127. Jin, H.; Wan, Y.; Liu, Z. Comprehensive evaluation of RNA-seq quantification methods for linearity. BMC Bioinform. 2017, 18, 51–59. [Google Scholar] [CrossRef] [Green Version]
  128. Conesa, A.; Madrigal, P.; Tarazona, S.; Gomez-cabrero, D.; Cervera, A.; Mcpherson, A.; Szcze, W.; Gaffney, D.J.; Elo, L.L.; Zhang, X.; et al. A survey of best practices for RNA-seq data analysis. Genome Biol. 2016, 17, 13. [Google Scholar] [CrossRef] [Green Version]
  129. Assefa, A.T.; De Paepe, K.; Everaert, C.; Mestdagh, P.; Thas, O.; Vandesompele, J. Differential gene expression analysis tools exhibit substandard performance for long non-coding RNA-sequencing data. Genome Biol. 2018, 19, 96. [Google Scholar] [CrossRef] [Green Version]
  130. Wang, C.; Schröder, M.S.; Hammel, S.; Butler, G. Using RNA-seq for Analysis of Differential Gene Expression in Fungal Species. In Methods in Molecular Biology; Devaux, F., Ed.; Springer Science+Business Media: New York, NY, USA, 2016; Volume 1361, pp. 1–40. ISBN 9781493930791. [Google Scholar]
  131. Pawlik, A.; Mazur, A.; Wielbo, J.; Koper, P.; Zebracki, K.; Kubik-Komar, A.; Janusz, G. RNA Sequencing Reveals Differential Gene Expression of Cerrena Unicolor in Response to Variable Lighting Conditions. Int. J. Mol. Sci. 2019, 20, 290. [Google Scholar] [CrossRef] [Green Version]
  132. Wang, Y.C.; Peterson, S.E.; Loring, J.F. Protein post-translational modifications and regulation of pluripotency in human stem cells. Cell Res. 2014, 24, 143–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Aslam, B.; Basit, M.; Nisar, M.A.; Khurshid, M.; Rasool, M.H. Proteomics: Technologies and their applications. J. Chromatogr. Sci. 2017, 55, 182–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Graves, P.R.; Haystead, T.A.J. Molecular Biologist’s Guide to Proteomics. Microbiol. Mol. Biol. Rev. 2002, 66, 39–63. [Google Scholar] [CrossRef] [Green Version]
  135. Amiri-dashatan, N.; Koushki, M.; Abbaszadeh, H.; Rostami-nejad, M.; Rezaei-Tavirani, M. Proteomics Applications in Health: Biomarker and Drug Discovery and Food Industry. Iran. J. Pharm. Res. 2018, 17, 1523–1536. [Google Scholar] [PubMed]
  136. Yokota, H. Applications of proteomics in pharmaceutical research and development. Biochim. Biophys. Acta Proteins Proteom. 2019, 1867, 17–21. [Google Scholar] [CrossRef]
  137. Champer, J.; Ito, J.I.; Clemons, K.V.; Stevens, D.A.; Kalkum, M. Proteomic Analysis of Pathogenic Fungi Reveals Highly Expressed Conserved Cell Wall Proteins. J. Fungi 2016, 2, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Quecine, M.C.; Leite, T.F.; Bini, A.P.; Regiani, T.; Franceschini, L.M.; Budzinski, I.G.F.; Marques, F.G.; Labate, M.T.V.; Guidetti-Gonzalez, S.; Moon, D.H.; et al. Label-free quantitative proteomic analysis of Puccinia psidii uredospores reveals differences of fungal populations infecting eucalyptus and guava. PLoS ONE 2016, 11, e0145343. [Google Scholar] [CrossRef] [Green Version]
  139. Bianco, L.; Perrotta, G. Methodologies and perspectives of proteomics applied to filamentous fungi: From sample preparation to secretome analysis. Int. J. Mol. Sci. 2015, 16, 5803–5829. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Ball, B.; Bermas, A.; Carruthers-lay, D.; Geddes-McAlister, J. Mass Spectrometry-Based Proteomics of Fungal Pathogenesis, Host—Fungal Interactions, and Antifungal Development. J. Fungi 2019, 5, 52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Leach, M.D.; Brown, A.J.P. Posttranslational modifications of proteins in the pathobiology of medically relevant fungi. Eukaryot. Cell 2012, 11, 98–108. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Ramazi, S.; Zahiri, J. Posttranslational modifications in proteins: Resources, tools and prediction methods. Database 2021, 2021, baab012. [Google Scholar] [CrossRef] [PubMed]
  143. Wang, Q.; Zhong, C.; Xiao, H. Genetic Engineering of Filamentous Fungi for Efficient Protein Expression and Secretion. Front. Bioeng. Biotechnol. 2020, 8, 293. [Google Scholar] [CrossRef]
  144. Wang, P.; Wang, H.; Gai, J.; Tian, X.; Zhang, X.; Lv, Y.; Jian, Y. Evolution of protein N-glycosylation process in Golgi apparatus which shapes diversity of protein N-glycan structures in plants, animals and fungi. Sci. Rep. 2017, 7, 40301. [Google Scholar] [CrossRef] [Green Version]
  145. Goto, M. Protein O-glycosylation in fungi: Diverse structures and multiple functions. Biosci. Biotechnol. Biochem. 2007, 71, 1415–1427. [Google Scholar] [CrossRef] [Green Version]
  146. Deshpande, N.; Wilkins, M.R.; Packer, N.; Nevalainen, H. Protein glycosylation pathways in filamentous fungi. Glycobiology 2008, 18, 626–637. [Google Scholar] [CrossRef]
  147. Tokmakov, A.A.; Kurotani, A.; Takagi, T.; Toyama, M.; Shirouzu, M.; Fukami, Y.; Yokoyama, S. Multiple post-translational modifications affect heterologous protein synthesis. J. Biol. Chem. 2012, 287, 27106–27116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Cruz, E.R.; Nguyen, H.; Nguyen, T.; Wallace, I.S. Functional analysis tools for post-translational modification: A post-translational modification database for analysis of proteins and metabolic pathways. Plant J. 2019, 99, 1003–1013. [Google Scholar] [CrossRef]
  149. Rugabber, T.P.; Talley, J.W. Enhancing bioremediation with enzymatic processes: A review. Pract. Period. Hazard. Toxic Radioact. Waste Manag. 2006, 10, 73–85. [Google Scholar] [CrossRef]
  150. Ellouze, M.; Sayadi, S. White-Rot Fungi and their Enzymes as a Biotechnological Tool for Xenobiotic Bioremediation. In Management of Hazardous Wastes; IntechOpen: London, UK, 2016; pp. 103–120. [Google Scholar]
  151. Arnau, J.; Yaver, D.; Hjort, C.M. Strategies and Challenges for the Development of Industrial Enzymes Using Fungal Cell Factories. In Grand Challenges in Biology and Biotechnology; Nevalainen, H., Ed.; Springer Nature: Basel, Switzerland, 2020; pp. 179–210. ISBN 9783030295417. [Google Scholar]
  152. Duzenli, O.F.; Okay, S. Promoter engineering for the recombinant protein production in prokaryotic systems. AIMS Bioeng. 2020, 7, 62–81. [Google Scholar] [CrossRef]
  153. Sevillano, L.; Vijgenboom, E.; van Wezel, G.P.; Díaz, M.; Santamaría, R.I. New approaches to achieve high level enzyme production in Streptomyces lividans. Microb. Cell Fact. 2016, 15, 28. [Google Scholar] [CrossRef] [Green Version]
  154. Che, Z.; Cao, X.; Chen, G.; Liang, Z. An effective combination of codon optimization, gene dosage, and process optimization for high-level production of fibrinolytic enzyme in Komagataella phaffii (Pichia pastoris). BMC Biotechnol. 2020, 20, 63. [Google Scholar] [CrossRef]
  155. Khan, S.; Ullah, M.W.; Siddique, R.; Nabi, G.; Manan, S.; Yousaf, M.; Hou, H. Role of recombinant DNA technology to improve life. Int. J. Genom. 2016, 2016, 2405954. [Google Scholar] [CrossRef]
  156. Gupta, V.; Sengupta, M.; Prakash, J.; Tripathy, B.C. Production of Recombinant Pharmaceutical Proteins. In Basic and Applied Aspects of Biotechnology; Springer Science+Business Media: Singapore, 2017; pp. 77–100. ISBN 9789811008757. [Google Scholar]
  157. Gifre, L.; Arís, A.; Bach, À.; Garcia-Fruitós, E. Trends in recombinant protein use in animal production. Microb. Cell Fact. 2017, 16, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Tripathi, N.K.; Shrivastava, A. Recent Developments in Bioprocessing of Recombinant Proteins: Expression Hosts and Process Development. Front. Bioeng. Biotechnol. 2019, 7, 420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. von Schaewen, A.; Jeong, I.S.; Rips, S.; Fukudome, A.; Tolley, J.; Nagashima, Y.; Fischer, K.; Kaulfuerst-Soboll, H.; Koiwa, H. Improved recombinant protein production in Arabidopsis thaliana. Plant Signal. Behav. 2018, 13, e1486149. [Google Scholar] [CrossRef] [Green Version]
  160. Jia, B.; Jeon, C.O. High-throughput recombinant protein expression in Escherichia coli: Current status and future perspectives. Open Biol. 2016, 6, 160196. [Google Scholar] [CrossRef] [Green Version]
  161. Nishikawa, R.; Yoshida, M.; Noda, T.; Okuhara, T.; Taguchi, G.; Inatomi, S.; Shimosaka, M. pFungiway: A series of plasmid vectors used for gene manipulation in fungi. Ann. Microbiol. 2016, 66, 825–832. [Google Scholar] [CrossRef]
  162. Hartley, J.L. Cloning technologies for protein expression and purification. Curr. Opin. Biotechnol. 2006, 17, 359–366. [Google Scholar] [CrossRef] [PubMed]
  163. Rocco, C.J.; Dennison, K.L.; Klenchin, V.A.; Rayment, I.; Escalante-Semerena, J.C. Construction and use of new cloning vectors for the rapid isolation of recombinant proteins from Escherichia coli. Plasmid 2008, 59, 231–237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Hohnholz, R.; Pohlmann, K.J.; Achstetter, T. A set of isomeric episomal plasmids for systematic examination of mitotic stability in Saccharomyces cerevisiae. Yeast 2017, 34, 267–275. [Google Scholar] [CrossRef] [Green Version]
  165. Siddiqui, M.S.; Choksi, A.; Smolke, C.D. A system for multi-locus chromosomal integration and transformation-free selection marker rescue. FEMS Yeast Res. 2014, 14, 1171–1185. [Google Scholar] [CrossRef] [Green Version]
  166. Amen, T.; Kaganovich, D. Integrative modules for efficient genome engineering in yeast. Microb. Cell 2017, 4, 182–190. [Google Scholar] [CrossRef] [Green Version]
  167. Falcon, A.A.; Rios, N.; Aris, J.P. 2-micron circle plasmids do not reduce yeast life span. FEMS Microbiol. Lett. 2005, 250, 245–251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. Żymańczyk-Duda, E.; Brzezińska-Rodak, M.; Klimek-Ochab, M.; Duda, M.; Zerka, A. Yeast as a Versatile Tool in Biotechnology. In Yeast-Industrial Applications; Morata, A., Loira, I., Eds.; IntechOpen: London, UK, 2017; pp. 3–40. [Google Scholar]
  169. Gnügge, R.; Rudolf, F. Saccharomyces cerevisiae Shuttle vectors. Yeast 2017, 34, 205–221. [Google Scholar] [CrossRef] [Green Version]
  170. Larionov, V.; Kouprina, N.; Graves, J.; Resnick, M.A. Highly selective isolation of human DNAs from rodent-human hybrid cells as circular yeast artificial chromosomes by transformation-associated recombination cloning. Proc. Natl. Acad. Sci. USA 1996, 93, 13925–13930. [Google Scholar] [CrossRef] [Green Version]
  171. Jung, H.M.; Kim, Y.H. Simultaneous overexpression of integrated genes by copy number amplification of a mini-yeast artificial chromosome. J. Microbiol. Biotechnol. 2018, 28, 821–825. [Google Scholar] [CrossRef]
  172. Alharbi, R.A. Proteomics approach and techniques in identification of reliable biomarkers for diseases. Saudi J. Biol. Sci. 2020, 27, 968–974. [Google Scholar] [CrossRef] [PubMed]
  173. Syahir, A.; Usui, K.; Tomizaki, K.; Kajikawa, K.; Mihara, H. Label and Label-Free Detection Techniques for Protein Microarrays. Microarrays 2015, 4, 228–244. [Google Scholar] [CrossRef] [Green Version]
  174. Braitbard, O.; Bishara-Shieban, J.; Glickstein, H.; Kott-Gutkowski, M.; Pace, U.; Rund, D.G.; Stein, W.D. An ELISA-based procedure for assaying proteins in digests of human leukocytes and cell lines, using specifically selected peptides and appropriate antibodies. Proteome Sci. 2006, 4, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Kim, Y.S.; Jellison, J.; Goodell, B.; Tracy, V.; Chandhoke, V. The Use of ELISA for the Detection of White- and Brown-Rot Fungi. Holzforschung 1991, 45, 403–406. [Google Scholar] [CrossRef]
  176. Martin-Souto, L.; Buldain, I.; Areitio, M.; Aparicio-Fernandez, L.; Antoran, A.; Bouchara, J.P.; Martin-Gomez, M.T.; Rementeria, A.; Hernando, F.L.; Ramirez-Garcia, A. ELISA Test for the Serological Detection of Scedosporium/Lomentospora in Cystic Fibrosis Patients. Front. Cell. Infect. Microbiol. 2020, 10, 602089. [Google Scholar] [CrossRef]
  177. Santana, A.E.; Taborda, C.P.; Severo, J.S.; Gomes Rittner, G.M.; Muñoz, J.E.; Larsson, C.E.; Larsson, C.E. Development of enzyme immunoassays (ELISA and Western blot) for the serological diagnosis of dermatophytosis in symptomatic and asymptomatic cats. Med. Mycol. 2018, 56, 95–102. [Google Scholar] [CrossRef] [Green Version]
  178. Kim, Y.; Bigelow, L.; Borovilos, M.; Dementieva, I.; Duggan, E.; Hatzos, C.; Joachimiak, G.; Li, H.; Maltseva, N.; Mulligan, R.; et al. High-Throughput Protein Purification for X-Ray Crystallography and NMR. Adv. Protein Chem. Struct. Biol. 2008, 75, 85–105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Thiemann, J.; Jankowski, J.; Rykl, J.; Kurzawski, S.; Pohl, T.; Wittmann-Liebold, B.; Schlüter, H. Principle and applications of the protein-purification-parameter screening system. J. Chromatogr. A 2004, 1043, 73–80. [Google Scholar] [CrossRef] [PubMed]
  180. More, S.S.; Renuka, P.S.; Pruthvi, K.; Swetha, M.; Malini, S.; Veena, S.M. Isolation, purification, and characterization of fungal laccase from Pleurotus sp. Enzym. Res. 2011, 2011, 248735. [Google Scholar] [CrossRef] [Green Version]
  181. Mukhopadhyay, M.; Banerjee, R. Purification and biochemical characterization of a newly produced yellow laccase from Lentinus squarrosulus MR13. Biotechniques 2015, 5, 227–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Irfan, M.; Mehmood, S.; Irshad, M.; Anwar, Z. Optimized production, purification and molecular characterization of fungal laccase through Alternaria alternata. Turk. J. Biochem. 2018, 43, 613–622. [Google Scholar] [CrossRef]
  183. Carrasco, M.; Alcaíno, J.; Cifuentes, V.; Baeza, M. Purification and characterization of a novel cold adapted fungal glucoamylase. Microb. Cell Fact. 2017, 16, 75. [Google Scholar] [CrossRef] [PubMed]
  184. Duong-Ly, K.C.; Gabelli, S.B. Salting out of proteins using ammonium sulfate precipitation. In Methods in Enzymology; Elsevier Inc.: Baltimore, MD, USA, 2014; Volume 541, pp. 85–94. ISBN 9780124201194. [Google Scholar]
  185. Bhat, E.; Abdalla, M.; Rather, I. Key Factors for Successful Protein Purification and Crystallization. Glob. J. Biotechnol. Biomater. Sci. 2018, 4, 001–007. [Google Scholar] [CrossRef] [Green Version]
  186. Ó’Fágáin, C.; Cummins, P.M.; O’Connor, B.F. Gel-Filtration Chromatography. In Protein Chromatography; Walls, D., Loughran, S.T., Eds.; Springer Science+Business Media: New York, NY, USA, 2017; Volume 1485, pp. 15–25. ISBN 9781493964109. [Google Scholar]
  187. Acikara, Ö.B. Ion-Exchange Chromatography and Its Applications. In Column Chromatography; IntechOpen: London, UK, 2013; pp. 31–58. [Google Scholar]
  188. Kosanović, M.; Milutinović, B.; Goč, S.; Mitić, N.; Janković, M. Ion-Exchange Chromatography Purification of Extracellular Vesicles. Biotechniques 2017, 63, 65–71. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Rodriguez, E.L.; Poddar, S.; Iftekhar, S.; Suh, K.; Woolfork, A.G.; Ovbude, S.; Pekarek, A.; Walters, M.; Lott, S.; Hage, D.S. Affinity chromatography: A review of trends and developments over the past 50 years. J. Chromatogr. B 2020, 122332. [Google Scholar] [CrossRef] [PubMed]
  190. Mant, C.T.; Chen, Y.; Yan, Z.; Popa, T.V.; Kovacs, J.M.; Mills, J.B.; Tripet, B.P.; Hodges, R.S. HPLC analysis and purification of peptides. In Methods in Molecular Biology; Fields, G., Ed.; Humana Press Inc.: Totowa, NJ, USA, 2007; Volume 386, pp. 3–55. [Google Scholar]
  191. Encarnación, S.; Hernández, M.; Martínez-Batallar, G.; Contreras, S.; del Carmen Vargas, M.; Mora, J. Comparative proteomics using 2-D gel electrophoresis and mass spectrometry as tools to dissect stimulons and regulons in bacteria with sequenced or partially sequenced genomes. Biol. Proced. Online 2005, 7, 117–135. [Google Scholar] [CrossRef] [PubMed]
  192. Dias, L.L.C.; Balbuena, T.S.; Silveira, V.; Santa-Catarina, C.; Schevchenko, A.; Floh, E.I.S. Two-dimensional gel electrophoretic protein profile analysis during seed development of Ocotea catharinensis: A recalcitrant seed species. Braz. J. Plant Physiol. 2010, 22, 23–33. [Google Scholar] [CrossRef]
  193. Najmanovich, R.J.; Torrance, J.W.; Thornton, J.M. Prediction of protein function from structure: Insights from methods for the detection of local structural similarities. Biotechniques 2005, 38, 847–851. [Google Scholar] [CrossRef]
  194. Yee, A.A.; Savchenko, A.; Ignachenko, A.; Lukin, J.; Xu, X.; Skarina, T.; Evdokimova, E.; Liu, C.S.; Semesi, A.; Guido, V.; et al. NMR and X-ray crystallography, complementary tools in structural proteomics of small proteins. J. Am. Chem. Soc. 2005, 127, 16512–16517. [Google Scholar] [CrossRef]
  195. Bryn Fenwick, R.; Van Den Bedem, H.; Fraser, J.S.; Wright, P.E. Integrated description of protein dynamics from room-temperature X-ray crystallography and NMR. Proc. Natl. Acad. Sci. USA 2014, 111, E445–E454. [Google Scholar] [CrossRef] [Green Version]
  196. Chen, C.; Huang, H.; Wu, C.H. Protein Bioinformatics Databases and Resources. Methods Mol. Biol. 2017, 1558, 3–39. [Google Scholar] [CrossRef]
  197. Mills, C.L.; Beuning, P.J.; Ondrechen, M.J. Biochemical functional predictions for protein structures of unknown or uncertain function. Comput. Struct. Biotechnol. J. 2015, 13, 182–191. [Google Scholar] [CrossRef] [Green Version]
  198. Miyashita, M.; Presley, J.M.; Buchholz, B.A.; Lam, K.S.; Lee, Y.M.; Vogel, J.S.; Hammock, B.D. Attomole level protein sequencing by Edman degradation coupled with accelerator mass spectrometry. Proc. Natl. Acad. Sci. USA 2001, 98, 4403–4408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Steen, H.; Mann, M. The ABC’s (and XYZ’s) of peptide sequencing. Nat. Rev. Mol. Cell Biol. 2004, 5, 699–711. [Google Scholar] [CrossRef] [PubMed]
  200. Restrepo-Pérez, L.; Joo, C.; Dekker, C. Paving the way to single-molecule protein sequencing. Nat. Nanotechnol. 2018, 13, 786–796. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Pundir, S.; Magrane, M.; Martin, M.J.; O’Donovan, C.; Consortium, T.U. Searching and navigating UniProt databases. Curr. Protoc. Bioinform. 2015, 50, 1.27.1–1.27.10. [Google Scholar] [CrossRef] [Green Version]
  202. Pruitt, K.D.; Tatusova, T.; Maglott, D.R. NCBI reference sequences (RefSeq): A curated non-redundant sequence database of genomes, transcripts and proteins. Nucleic Acids Res. 2007, 35, 61–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. McGarvey, K.M.; Goldfarb, T.; Cox, E.; Farrell, C.M.; Gupta, T.; Joardar, V.S.; Kodali, V.K.; Murphy, M.R.; O’Leary, N.A.; Pujar, S.; et al. Mouse genome annotation by the RefSeq project. Mamm. Genome 2015, 26, 379–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Fungal protein (enzyme) expression process.
Figure 1. Fungal protein (enzyme) expression process.
Jof 07 00700 g001
Figure 2. Relevant fungal metabolites.
Figure 2. Relevant fungal metabolites.
Jof 07 00700 g002
Table 2. Examples of Eukaryotic expression vectors compatible with fungi.
Table 2. Examples of Eukaryotic expression vectors compatible with fungi.
HostVectorReferences
Yeast
  • Yeast Episomal Shuttle Vectors (YEp type)
[164]
  • Yeast Integrative Plasmids (YIps)
[165,166]
  • Yeast Replicative Plasmids (YRps)
[167,168]
  • Yeast Centromeric Plasmids (YCps)
[169]
  • Yeast Artificial Chromosome (YAC)
[168,170,171]
Agrobacterium tumefaciens
  • pFungiway vectors 1
[161]
1 Several vector types within this group.
Table 3. Protein purification techniques.
Table 3. Protein purification techniques.
TechniquePrincipleReferences
Salting in/outExploiting protein solubility by increasing salt concentration in the solution[184]
DialysisUsing size exclusion to separate proteins from small molecules and ions that pass through a semi-permeable membrane[185]
Gel filtration chromatographySeparation of proteins based on size[186]
Ion-exchange chromatographySeparation of proteins based on their net charge[187,188]
Affinity chromatographyExploiting the affinity of proteins for given chemical groups to separate them[189]
High-Pressure Liquid Chromatography (HPLC)Can use different principles of column chromatography to separate proteins using high pressure to give better resolution[190]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ijoma, G.N.; Heri, S.M.; Matambo, T.S.; Tekere, M. Trends and Applications of Omics Technologies to Functional Characterisation of Enzymes and Protein Metabolites Produced by Fungi. J. Fungi 2021, 7, 700. https://doi.org/10.3390/jof7090700

AMA Style

Ijoma GN, Heri SM, Matambo TS, Tekere M. Trends and Applications of Omics Technologies to Functional Characterisation of Enzymes and Protein Metabolites Produced by Fungi. Journal of Fungi. 2021; 7(9):700. https://doi.org/10.3390/jof7090700

Chicago/Turabian Style

Ijoma, Grace N., Sylvie M. Heri, Tonderayi S. Matambo, and Memory Tekere. 2021. "Trends and Applications of Omics Technologies to Functional Characterisation of Enzymes and Protein Metabolites Produced by Fungi" Journal of Fungi 7, no. 9: 700. https://doi.org/10.3390/jof7090700

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