Next Article in Journal
Detection of Toxocara cati Larvae from Ostrich and Wild Boar Meat Intended for Human Consumption
Next Article in Special Issue
Expression Patterns of miR398, miR167, and miR159 in the Interaction between Bread Wheat (Triticum aestivum L.) and Pathogenic Fusarium culmorum and Beneficial Trichoderma Fungi
Previous Article in Journal
Persistence of Antibody Responses to the SARS-CoV-2 in Dialysis Patients and Renal Transplant Recipients Recovered from COVID-19
Previous Article in Special Issue
Culturable Seed Microbiota of Populus trichocarpa
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Fungi Inhabiting the Wheat Endosphere

by
Lidia Błaszczyk
*,
Sylwia Salamon
and
Katarzyna Mikołajczak
Department of Plant Microbiomics, Institute of Plant Genetics, Polish Academy of Sciences, 34 Strzeszyńska Street, 60-479 Poznań, Poland
*
Author to whom correspondence should be addressed.
Pathogens 2021, 10(10), 1288; https://doi.org/10.3390/pathogens10101288
Submission received: 31 August 2021 / Revised: 1 October 2021 / Accepted: 4 October 2021 / Published: 7 October 2021
(This article belongs to the Special Issue Endophytes in Plant Health and Disease)

Abstract

:
Wheat production is influenced by changing environmental conditions, including climatic conditions, which results in the changing composition of microorganisms interacting with this cereal. The group of these microorganisms includes not only endophytic fungi associated with the wheat endosphere, both pathogenic and symbiotic, but also those with yet unrecognized functions and consequences for wheat. This paper reviews the literature in the context of the general characteristics of endophytic fungi inhabiting the internal tissues of wheat. In addition, the importance of epigenetic regulation in wheat–fungus interactions is recognized and the current state of knowledge is demonstrated. The possibilities of using symbiotic endophytic fungi in modern agronomy and wheat cultivation are also proposed. The fact that the current understanding of fungal endophytes in wheat is based on a rather small set of experimental conditions, including wheat genotypes, plant organs, plant tissues, plant development stage, or environmental conditions, is recognized. In addition, most of the research to date has been based on culture-dependent methods that exclude biotrophic and slow-growing species and favor the detection of fast-growing fungi. Additionally, only a few reports of studies on the entire wheat microbiome using high-throughput sequencing techniques exist. Conducting comprehensive research on the mycobiome of the endosphere of wheat, mainly in the context of the possibility of using this knowledge to improve the methods of wheat management, mainly the productivity and health of this cereal, is needed.

1. Introduction

Fungi play an essential role in natural ecosystems and in modern agriculture because of their nutritional versatility, miscellaneous lifestyle, and multifarious interactions with plants. Fungi are important decomposers and recyclers of organic materials [1]. They interact with plant roots in the rhizosphere or with aboveground plant components; while living in close association with plants, they are located either outside or within plant tissues [1]. Fungi that periodically or constantly colonize the internal parts of plant tissues without disease manifestation in their host are defined as fungal endophytes [2,3,4]. Fungal endophytes requiring plant tissues to complete their life cycle are classified as “obligate” endophytes. Well-documented examples of obligate endophytes are found among mycorrhizal fungi and members of the fungal genera Balansia, Epichloë, and Neotyphodium from the family Clavicipitaceae (Ascomycota) [5,6]. However, fungal endophytes that mainly thrive outside plant tissues and sporadically enter the plant endosphere are called “opportunistic” endophytes [7,8]. Between these two groups is an intermediate group, which includes the vast majority of endophytic fungi, the so-called “facultative” endophytes [9,10].
Fungi that remain endophytic throughout the entire life cycle of the host are categorized as clavicipitaceus endophytes (C-endophytes) and represent class I fungal endophytes [8,11]. Species of clavicipitaceus endophytes, including Balansia spp., Epichloë spp., and Claviceps spp., establish symbioses almost exclusively with grass, rush, and sledge hosts [9,12], in which they may colonize the entire host plant systemically. Members of this class proliferate in the plant shoot meristem, colonizing intercellular spaces of the newly forming shoots, and can be transmitted vertically via seeds [13]. Some Epichloë and Neotyphodium species may also be transmitted horizontally via leaf fragments falling onto the soil [14].
Fungi that do not remain endophytic throughout the entire life cycle of the host, and furthermore may not be present during the entire life cycle of the host, are categorized as non-clavicipitaceus endophytes (NC-endophytes) and represent the three functional classes: class 2, defined as containing the fungi colonizing above- and below-ground plant tissues, i.e., the rhizosphere, endorhiza, and aerial tissues [15], and being horizontally and/or vertically transmitted [16]; class 3, defined as containing members of the Dikaryomycota (Ascomycota or Basidiomycota) that are mostly confined to the air tissues of various hosts, especially trees, but also other plant taxa [17,18] and are transmitted horizontally [19]; class 4, which comprise dark, septate endophytes, which, similar to mycorrhizal fungi, are restricted to roots, where they reside inter- and/or intracellularly in the cortical cell layers [20]. Detailed information of fungal endophytes classification has been compiled in the review by Rodriguez et al. [11].
Endophytic fungal were first isolated from Lolium temulentum seeds in 1898 [3]. Currently, no plants have been found without these microorganisms. Endophytes are present in both large trees [21] and lichens [22]. Together with the plants in which they exist, fungal endophytes can occur in various environments: in agricultural and natural, terrestrial [23] and aquatic [24], tropical [25], and high-mountain [25,26]. In addition, endophytes can colonize various plant tissues (Table 1) intercellularly or intracellularly and display various interactions within their hosts. Relationships between the plant and the endophytic fungus can range from beneficial (mutualism or commensalism) to those that are pathogenic to the host plant [8]. Nonetheless, the functions of many of endogenous fungi are still poorly understood. The presence of fungi in the endosphere of plants, regardless of their lifestyle and way of nutrition, is not without significance. It is well known that sterile plants have a reduced vigor [27,28], while introducing endophytic fungi into plant tissue can provide them with many benefits, such as improving growth rates or enhancing defense and immune responses to biotic [29,30] and abiotic stresses [31,32,33]. Therefore, research on the analysis of microbiomes of cultivated plants, including cereals, in order to find beneficial endophytic fungi and formulating supplements for plants based on them is common.
Cereals are critical to global food production and global food security due to their use as an important food for humans and livestock. One of the main cereals is wheat—statistics shows that in the crop year 2018/2019 over 254 million tonnes were produced in Europe, which accounts for 33.9% of the world’s wheat production [34]. However, wheat production in Europe is mainly affected by the occurrence of drought, late spring frosts, and severe winter frosts associated with inadequate snow cover [35]. Recent crop grain breeding programs have made steady improvements in yield quantity and quality, along with biotic and abiotic stress toleration, which have changed the scope and efficiency of wheat breeding strategies [36]. Usually, high productivity of crop grains is accompanied by extensive utilization of agrochemicals for improvement of soil fertility and control of plant diseases, causing drastic effects on the environment and public health. In order to reduce the negative effects of toxic chemicals, there is a continuous global emphasis on sustainable and less chemically dependent organic agriculture. This opens the way for the use of microbial biological control agents, including endophytic fungi. As mentioned above, studying the ecological role of endophytes and understanding the complex interaction between endophytes and host wheat could lead to the identification of symbiotic fungi and, consequently, the design of plant growth biostimulants or a new generation of biological control agents to improve tolerance to the biotic and abiotic stresses of this economically important cereal.
In this paper, we review the literature in the context of the general characterization of fungal endophytes, with a particular focus on studies of wheat endophytes in different varieties and geographic regions. Despite numerous reviews of grass endophytes [5], trees [37], and tropical plants [29], no comprehensive study summarizing the knowledge regarding fungal endophytes and commonly grown cereal plants found in this study exists. We further notice the importance of epigenetic regulation in wheat–fungi interactions and demonstrate the current state of knowledge. Moreover, the possibilities of using symbiotic endophytic fungi in modern agronomy and wheat cultivation are proposed here.

2. Isolation of Fungi from the Wheat Endosphere

To study endophytic fungi, culture-based methods [38,39,40,41] and cultivation-independent techniques [42,43,44] are generally used.
Cultivation-dependent techniques are applied to extract fungi growing in plant tissues [38,39,40,41]. Their isolation is mainly based on the fragmentation of plant organs into small fragments, surface sterilization, and then their placement on microbiological agar media [41,45]. The method of plant fragmentation is very popular and is widely used, due to its simplicity and the variety of fungi obtained. On the other hand, limitations of this method also exist, which should be taken into account at the planning stage of the experiments [3,39,43]. Interestingly, Gamboa et al. [45] noticed that endophyte species diversity is negatively correlated with the size of the tissue fragment used for isolation. Thus, more fungal strains can be obtained using smaller pieces of plant tissues. In addition to the size of the plant tissues, factors such as the surface sterilization of plant fragments, the growth media, the incubation conditions for fungal cultures, and the ability of the fungi to sporulate also have impacts on the result of the experiment. A high risk of some fungi, especially those that are less competitive or that grow slowly and can therefore be easily replaced (displaced) by rapidly growing species, being overlooked exists.
In research on wheat endophytes, before surface sterilization, plant samples are usually washed under running water [46,47], and soil residues are removed from the roots, for example by brushing or scraping [48,49]. Ethyl alcohol and sodium hypochlorite are usually used to remove microorganisms from the surface of plant fragments. After sterilization, the tissues are rinsed with distilled water to remove the residues of the reagents used. However, a number of protocols are used in the surface sterilization of plant fragments. Larran et al. [46,47] sterilized wheat leaves, stems, glumes, and seeds via immersion in 96% ethanol for 1 min, sodium hypochlorite (2% available chlorine v/v) for 3 min, and again in 96% ethanol for 30 s, and finally by rinsing them twice in sterile distilled water. Comby et al. [38] sterilized wheat roots, stems, and leaves by immersing them in 70% ethyl alcohol for 2 min, in 0.5% sodium hypochlorite for 2 min, and in 70% ethyl alcohol for 1 min and finally by briefly rinsing them in sterile distilled water. In the study by Hubbard et al. [32], the wheat seeds were sterilized in 95% ethanol for 10 s, rinsed in sterile distilled water for 10 s, immersed in 5% sodium hypochlorite for 3 min, and then rinsed three times in sterile distilled water. In the work of Bouzouin et al. [41], water-washed wheat root samples were surface sterilized in 75% ethanol for 1 min, and then, after rinsing them in sterile distilled water for 1 min, the roots were immersed in 5% sodium hypochlorite for 3 min and rinsed again in sterile distilled water. Rojas et al. [50] isolated endophytes from wheat kernels. Separated for this purpose, the glumes, lemmas, paleas, and kernels were surface sterilized according to Comby et al. [38]; however, they sterilized flower tissues in 96% ethanol for 1 min, 2% sodium hypochlorite for 3 min, and 96% ethanol for 30 s, and finally rinsed the tissues twice with sterile MilliQ water. In contrast, Cłapa et al. [48] and Salamon et al. [49] developed a protocol for wheat in which plant tissue fragments were rinsed in 70% ethyl alcohol for 30 s and then in 0.5% sodium hypochlorite for 2 min. In order to remove the reagents, the wheat tissues were rinsed several times with distilled water. The sterilization efficiency is usually tested by placing a volume of the last-rinse water on agar plates.
After surface sterilization, the plant material was aseptically cut into smaller pieces 5–10 mm in length [48,49], re-sterilized, and then placed on the agar medium. Many different media can be used, but the most common ones are potato dextrose agar (PDA), malt extract agar (MEA), corn meal agar (CMA), yeast extract peptone dextrose agar, as well as minimal plant tissues or extract media [51,52,53,54]. Typically, the culture medium is supplemented with antibiotics: chloramphenicol, ampicillin, streptomycin sulphate and chlorotetracyclin hydrochloride, and amoxiallin. As in the case of sterilization, many variations of the incubation of plant tissue fragments on a solid medium exist. The temperature range used was 22 C [49]–27 C [47]. Cultures were maintained for a period of 5 days [47] to 28 days [41] or until the mycelium became visible [49], both in the dark [41,50] and in the photoperiod [47,49]. The grown mycelium was then passaged repeatedly onto agar media. However, a laborious and experiential step to obtain pure, monospore cultures via multiple dilutions and the use of micromanipulation techniques was performed. This stage is extremely important for further identification of the fungal isolate as well as for further characterization and use.

3. Identification of Endophytic Fungi

Traditionally, morphological features have been used as an approach to identify isolated fungal species, such as observations of mycelium growth in agar media, which enables the assessment of physiological properties such as colony color and growth rate or microscopic observations of spores and spore-producing structures resulting from asexual or sexual reproduction [3]. The techniques light microscopy and scanning electron microscopy make it possible to evaluate the observations of spores in terms of color, shape, and surface type as well as to identify various spore-bearing structures. This approach is still used but requires experienced mycologists. Furthermore, identification based solely on morphological features [55] is not always sufficient, especially when performing identifications at a lower level of classification, e.g., to a species [56,57]. This is the case, for example, due to the high morphological variability of isolates within one species caused by hybridization [58], due to cryptic speciation and evolutionary convergence [59]. In addition to physiology, morphology, or ultrastructure, the tissue biochemistry, ecological features, and chemotaxonomic features of fungi are not always correctly classified using traditional taxonomic methods [59].
The use of molecular techniques such as sequencing methods and the introduction of a DNA barcoding system have overcome the obstacles in traditional identification methods. The DNA barcoding system uses a short and standardized DNA fragment to identify species of microorganisms [60]. Identification is simple when the nucleotide sequence is constant within each species and unique to one species [61]. The most commonly used region for the differentiation of fungi at the genus and species levels is the internal transcribed spacer 1 (ITS1) and 2 (ITS2) flanking the 5.8S rRNA gene. The ITS region is considered very stable, has many copies, and is usually conserved within the species [60,61,62,63,64,65]. The ITS region as a DNA barcode was used in the identification of many fungi important for agriculture, such as Colletotrichum, Fusarium, Alternaria, Puccinia, and Rhizoctonia [66]. The application of the ITS region as a DNA barcode has many advantages, such as successful amplification among all lineages of fungi using universal primers; suitable fragment lengths; and numerous curated molecular databases in NCBI, UNITE, and EzTaxon.
However, the ITS barcode has several shortcomings. Various interspecific and intraspecific distances exist between groups of fungi [67], and determining the ITS divergence threshold to distinguish between fungal species is often difficult [68]. In higher taxonomic classifications, the large subunit (LSU, 28S) of rRNA has been shown to have better discriminatory power than ITS as the 28S gene is more variable and is used in classification on the genus to phylum levels [65]. Moreover, using other DNA regions as targets in the methods for identifying fungi is recommended, such as fragments of genes encoding universal proteins: β-tubulin, translational elongation factor 1α, RNA polymerase II, ATP synthase, γ-actin, and calmodulin [56,69,70,71,72,73]. Currently, species identification is built on the basis of a multiloci DNA barcode rather than a single locus. A useful tool for this is multilocus sequence typing (MLST) [74,75,76].
In wheat endophyte research, ITS was the most commonly used DNA barcode in molecular identification, showing some potential in diversity studies or in the search for endophytic strains beneficial to wheat, despite existing limitations in the species discrimination [68,77]. This region of DNA was used only by Larran et al. [47] to identify endophytic fungi isolated from wheat leaves, stalks, chaff, and grains; by Comby et al. [38] for the classification of endoffites from aerial roots and organs, including leaves, stems, anthers, chaff, sediments, and nuclei; and by Bouzouina et al. [41] to determine the species of endophytic fungi isolated from wheat roots. Using high-resolution melting (HRM) techniques and the differences in the melting points of ITS sequences, distinguishing fungi isolated from the inner tissues of wheat plants at the genus level has become possible [48]. However, the multilocus DNA barcode was used by Llorens et al. [78] for correct classification of 2 isolates from the ancestor of wheat, namely Aegilops sharonensis; by Salamon et al. [49] for identifying 54 isolates from the root endosphere of common wheat and spelt wheat; and by Rojas et al. [50] to determine the species for 163 fungal isolates from healthy wheat spikes.
As mentioned at the beginning of this section, the identification of some groups of endophytic fungi can be very difficult, especially those that are closely related to plant tissues and do not grow on standard media or those that, due to their weak substrate competition, may be overlooked in cultures. Culture-independent methods have therefore gained a lot of attention. Among them, meta-barcoding approaches, especially ITS (ITS2) amplicon sequencing, are an important tool that has also been adopted in research on the endophytic fungi of wheat. Recently, this approach has been used by Sun et al. [44], who characterized and compared the communities of fungal endophytes (FEC) from common wheat, wild emmer wheat (Triticum dicoccoides Koern.), and sharon goatgrass (Aegilops sharonensis Eig). In contrast, Latz et al. [79] analyzed mycobiomes of wheat endospheres to discover the influence of host genotype; abiotic environment (temperature, humidity, and rainfall); and fungi present in the seed material, air, and soil on the formation of endophytic fungal communities in the tissues of wheat plants, along with its growth and development.
However, ITS sequence-based meta-barcoding has severe limitations in identifying most of the unknown taxa at the species level as many fungi have not been sequenced. Difficulties in correctly identifying taxonomies is also present due to ITS sequence annotations being falsely deposited in GenBank. In addition, for some groups of fungi, the ITS sequences show high inter- and intra-species variability, so the taxonomic assignments with the generally accepted 97% similarity threshold are not consistent for identification at the species level. Taxonomic fungi identification based on high-throughput sequencing can therefore only be justified at the genus or higher levels such as family or order [66,80]. Recently, comparisons of the results of studies on other crops have shown that endophytic fungi discovered by culture-dependent methods differ from those detected by cultivation-independent methods, most puzzlingly, with some isolated strains never having been found by culturing-independent methods [81,82]. This can be explained not only by the variability in the ITS sequence in relation to the sequences of other marker genes or the scarcity of databases, but also by the lack of convergence among taxonomic results, which may be affected by prosaic technical aspects such as the surface of the organ used for analyses; the effectiveness of its surface sterilization; and in the case of high-throughput methods, the effectiveness of tissue maceration and DNA isolation, the amplification reaction; and subsequent stages. Therefore, in research on wheat fungal endophytes, the need to use both approaches to link high-throughput data sets with the results of isolated fungi that are morphologically and phylogenetically identified is worth considering. Among other aspects not discussed here, it at least provides a complete picture of the structure of endophytic fungi in the individual analyzed. For a full insight into the complexity and dynamics of the wheat endosphere mycobiome, the influence of the wheat genotype, the type and age of the organ, the type of tissues, biotic and abiotic environmental factors, and the influence of the remaining microbiome of the studied individual should also be taken into account.

4. Assortment and Role of Fungal Endophytes in Wheat

Most fungal endophytes are commensal, have no or an unrecognized effect on the host plant, or show a mutualistic (positive) effect. Such categories of cooperation are known as symbiosis. Interestingly, the type of interaction may be temporary and change under the influence of external factors (e.g., stressful conditions); therefore, endophytes are also latent pathogens and dormant saprobes [8,82,83]. Symbiotic endophytes demonstrate a beneficial impact on their host plant; for example, they can oppose pathogen development [84] by inducing defense mechanisms in their host [85] or by producing antibiotics that inhibit the growth of other microorganisms, including pathogens [86,87]. Moreover, space and resource competition between endophyte and pathogen, or the existence of endophytes acting in a similar way to parasites of plant pathogens within plants were observed [88]. Because several fungi can combine different lifestyles (saprophytic, pathogenic, or symbiotic), their boundaries are often not clear-cut [1]. Many species that are pathogenic for some hosts may be asymptomatic for others [83]. In addition, many fungal endophytes may switch between pathogenic and commensal or mutualistic lifestyles depending on environmental conditions and on the host [37,83]. Based on several investigations, growing evidence suggests that the functions of fungal endophytes and, accordingly, the type of interactions with plants are affected by various abiotic and biotic factors, including environmental conditions, plant genotypes, plant tissue type, the fungal taxon, and strain type, as well as the dynamic network of interactions within the plant microbiome [88]. Nevertheless, the ecological role of endophytic fungi in plants, including wheat, is still poorly understood.
Research into the distribution and ecological role of fungal endophytes in wheat has been especially intensive in the family Clavicipitaceae, where the asexual genus Neotyphodium and closely related species of the sexual genus Epichloë have provided model systems [89]. In contrast with the well-known Epichloë and Neotyphodium associations with wheat, a lacuna exists in our knowledge of the diversity; the life cycles; and, accordingly, the ecological role of most nonclavicipitaceus endophytic species and the effects of their presence in their wheat host. Nevertheless, the occurrence of endophytic fungi in wheat (Triticaceae) has been demonstrated. The characteristics of Triticaceae endophytic fungi are summarized in Table 1 and visualized by species and organs of wheat (Triticum aestivum L., Triticum durum) in Figure 1 and Figure 2.
The above literature review (Table 1) revealed the possibility of the functioning of various roles of endophytic fungi in relation to wheat, from symbiotic, through saprophytic, to minor or latent pathogenic. Among the species that are considered pathogenic are mainly those of the genera Fusarium, Botritis, Cladosporium, Septoria, Sclerotinia, Rhizoctonia, Pyrenophora, Penicillium, Microdochium, or Epicoccum. On the other hand, fungi that show symbiotic interactions with wheat or have a beneficial effect on wheat fitness and yield, or are characterized by antagonistic activities towards its pathogens, are Trichoderma harzianum, Trichoderma hamatum, Trichoderma longibrachiatum, Rhodotorula rubra, Clonostachys rosea, or Chaetomium globosum species. Nevertheless, the role of most fungal endophytes is still poorly understood. It is supposed that species commonly known as pathogens or saprophytes, or those showing mutualistic interactions with plants, may play a completely different role by living in wheat tissue. The function of these species will not be known until the mechanisms of both the unidirectional interactions of these microorganisms with wheat and the complex network of interactions of the entire plant holobiont are known. Therefore, until the appropriate role of the fungus in the phase of its inhabiting the wheat endosphere is known, it should not be classified on the functional level. It is known that fungi have the ability to “switch” their lifestyle and mode of nutrition. Therefore, it is necessary to penetrate the mechanisms of interaction of wheat with endophytic fungi in order to understand the reason for this plant’s “agreement” to inhabit its endosphere.
The current understanding of fungal endophytes in wheat is built on a rather small set of experimental conditions, including wheat genotypes, plant organs, plant tissues, plant stage development, or environmental conditions [38,53,93,94,99,105,106,107,108,109,110,111,112,113,114]. Furthermore, all of these investigations have been based on culture-dependent methods that exclude biotrophic and slow-growing species and favor the detection of rapidly growing fungi [99,109]. Meanwhile, developments in high-throughput technologies, such as next-generation sequencing (NGS), have opened up new perspectives in fungal endophyte biodiversity research. The pioneers were Nicolaisen et al. [115], who adapted NGS to analyze the mycobiome of 90 wheat grain samples collected from Denmark. Subsequently, Karlsson et al. [106,116] investigated the effect of fungicide use and various crop management practices on the microbiome of the wheat phylosphere in Sweden. Molecular analysis of the ITS region in 220 wheat leaves from 22 fields in Sweden showed an average level of operational taxonomic units (OTU) of 54 and 40 in organic and conventional fields, respectively [116]. Meanwhile, Sapkota et al. [117] conducted a similar analysis on four wheat varieties at two locations in Denmark and identified a total of 212 OTUs. The authors observed that both the geographic location and location of the leaf, genotype, and plant growth stage had an impact on the architecture of the wheat mycobiome. The research of Hertz et al. [118] showed changes in the structure of the mycobiom of wheat ears occurring with their development, and with the time of exposure of plants to biotic and abiotic environmental conditions. Using high-throughput ITS1 sequencing, Yashiro et al. [119] characterized wheat mycobiomes at different stages of cereal processing, comparing domestic environments in rural and urban areas. At a similar time, Vujanovic et al. [108] demonstrated the transgenerational transmission of endophytic seed fungi through three consecutive generations of wheat under control conditions and drought stress, Shiro et al. [120] observed the occurrence of spatial variation in the microbiome of the phyllosphere of commercial wheat crops growing in the same field, and Knorr et al. [121] discovered the effect of fungicide treatments using different dosages, terms, and products on the mycobiom of the wheat phyllosphere. Recently, Latz et al. [79] studied the effects of the host genotype, temperature, humidity and rainfall, and the presence of fungi in the initial seed, air, and soil on the structure of the fungal community inhabiting the wheat endosphere. The studies have shown that the structure of the wheat mycobiome is complex and depends on various elements. A solid evaluation of the factors determining the influence on the wheat microbiome was performed by Kavamura et al. [122]. These factors include: (a) host genotype, growth stage, leaf positions, niche, organs, tissues, hormones; (b) exogenous compounds, namely fungicides, glyphosates, insecticides, phosphine fumigation of stored wheat grains, plastic mulch film residues; (c) fertilization; (d) inoculation of biocontrol agent; (e) land use; (f) management type; g) verhead irrigation; (h) rotation tillage; (i) soil history, type, physicochemical characteristics, and depth; (j) abiotic and biotic stresses; (k) geographical location; (l) growing season. As mentioned, details and literature references on these factors can be found in the review by Kavamura et al. [122].
Due to the recognition of so many factors influencing the structure of the micro- and thus mycobiome of the wheat endosphere, it prompted the search for indigenous species associated with this cereal and forming the so-called “core microbiome” that is constantly associated with a given host genotype. Simonin et al. [123] observed that among 177, 41 fungal taxa were consistently detected in the wheat rhizosphere of African and European soils, constituting a core microbiome. The most frequently detected genera were: Morteriella, Fusarium, Exophiala, and Chaetomium [123]. Schlatter et al. [124] described Nectriaceae, Ulocladium, Alternaria, Mortierella, and Microdochium as core fungal taxa in the rhizosphere of dryland wheat in the Inland Pacific Northwest. Rossmann et al. [125] identified 13 taxa of fungi, namely Fusarium, Fusicolla, Purpureocillium, Acremonium, Bionectria, Trichoderma, Penicillium, Kendrickiella, Exophiala, Chaetomium, Magnaporthiopsis, and Staphylotrichum, corresponding to the core microbiome of wheat cultivated in Brazil. In general, the fungi that typically constitute the core microbiota in wheat are pathogens, mainly of the genus Fusarium. Such a generalization, however, would be highly error-prone. In most studies, the core microbiota is defined on the basis of DNA sequence, where, for high-throughput analyzes, the lowest taxonomic unit is at the genus level. Moreover, the core microbiome can be defined in various ways, for example as a component of the microbiome that is constant for the host species over time, or one that determines the functioning of the host species and affects its health and maintenance of homeostasis [126]. Taking into account the functional context on the one hand, and the fact that pathogenic species are listed as components of the core of the microbiome on the other hand, it is worth considering the standardization of its definitions and determination methods. Although no complete understanding of the interactions exists between plants and fungi inhabiting their endosphere, some processes, mainly those involving symbiotic reactions, have been recognized.

5. Molecular Interaction between Endophytic Fungi and Wheat

To create and maintain symbiosis, constant communication between the mycobiome and the host plant is required. Sending signals can alter the gene expression and can modulate secreted proteins or metabolites, which have a positive impact on the host [127]. Multi-level interactions are present in the transcriptomes, proteomes, and metabolomes in symbiotic partners. Recent investigation involving Fourier transform infrared (FTIR) spectroscopy and hyperspectral imaging examined the interaction between the endophytic fungal isolate Penicillium sp. SMCD 2206 and kernel in durum wheat under drought stress conditions [128]. An altered chemical structure of coleorhizae inoculated with endophyte was observed, which resulted in improved tolerance to drought stress. O–H stretching, acyl lipid chains, proteins, polysaccharide carbohydrates, hemicelluloses, and possibly mannan and glucan may contribute to the chemical differences observed in coleorhizae. Epigenetic mechanisms, including DNA methylation, posttranslational histone modifications, and the activity of small RNAs (sRNAs) and long noncoding RNAs (lncRNAs), alter the chromatin structure and influence the accessibility of genetic information [129]. Such modifications are crucial for blocking the expression of non-genic sequences, such as transposons, repetitive sequences, or pseudogenes in plants [130,131]. Interestingly, epigenetic regulations are induced by environmental signals and can modulate the host–plant interaction with microorganisms and can also control the expression of stress-responsive genes in plants under stress [129,132,133]. Moreover, stress-induced epigenetic changes (epimutations and epialleles) are transient, while others can be stable, maintained, memorized, and transmitted to the next generations [133,134]. Despite the importance of epigenetic control in plant–fungi interactions, the available knowledge concerning wheat plants and endogenous fungi is scarce. However, the ever-increasing availability of high-throughput next-generation sequencing techniques, such as whole-genome bisulfite sequencing (WGBS), small RNA-seq, degradome-seq, etc., as well as recent advances made in wheat genome research [135,136,137] allow the epigenetic control of wheat response to biotic factors to be studied. To draw more attention to this unexplored issue and despite the small number of studies concerning epigenetic control of wheat–fungal endophytes interaction, we performed an extended literature review on the current state of knowledge concerning epigenetic regulation in non-model plant–fungi communication, described below.

5.1. Epigenetic Control of Wheat–Fungi Interaction

5.1.1. DNA Methylation

Growing evidence indicates that DNA methylation influences the expression of genes participating in plant response to abiotic and biotic factors. Recent studies suggest that the establishment of endosymbiotic relations is controlled by DNA methylation [132,133]. Beneficial arbuscular mycorrhizal fungus Funneliformis mosseae induced changes in the DNA methylation profile in Geranium robertianum [134], and conversely, DNA adenine methylation was altered in symbiotic Mesorhizobium loti as a result of beneficial relations with their host plants [135]. In wheat, knowledge regarding biotic factor-induced changes in the DNA methylation profile is mainly limited to fungal pathogens. Saripalli et al. [138] observed alterations in the cytosine methylation profiles of susceptible and resistant transgenic wheat lines 96 h after inoculation with biotrophic fungi Puccinia triticina, which is the causative agent of leaf rust in wheat. The wheat diploid progenitor Aegilops tauschii was used to evaluate the DNA methylation profile during infection with biotrophic fungi, namely Blumeria graminis f. sp. tritici (Bgt), causing powdery mildew [139]. The authors identified the cytosine methylated in the CHH context as the main loci regulated during the studied interaction, while an expression analysis carried out for certain genes confirmed these findings. However, knowledge concerning the alteration in cytosine methylation caused by endophytic plant symbionts is unexplored. Using methyl-sensitive amplified polymorphism (MSAP), Hubbard et al. (2014) [32] described different DNA methylation patterns in inoculated and uninoculated wheat seedlings in the analyzed conditions. The authors assessed a fungal endophyte referred to as SMCD 2206, which was isolated from surface-sterilized roots of Triticum turgidum L., and its role in improving drought and heat tolerance in wheat seeds [31]. Four groups of plants were analyzed: non-stress seedlings without and with SMCD 2206 inoculation, as well as inoculated and non-inoculated plants under drought stress. The DNA methylation pattern in inoculated seedlings under drought was similar to the profile demonstrated by non-stressed samples. Possibly, by changing DNA methylation status in wheat, endosymbiont SMCD2206 differentially expressed crucial genes. This research suggests that endophytes can change the DNA methylation of wheat plants and that the observed changes enhanced wheat resistance against abiotic stresses. Nevertheless, additional studies are needed to confirm these findings as well as to identify the differentially methylated genes that participated in the studied interaction. The question of how much wheat methylome differs in relation to fungi demonstrating different lifestyles or from diverse species still persists. Whether the induced changes are maintained in the next wheat generations and what the role of DNA methylation is in establishing endosymbiotic interactions also remain to be answered. This gap in our knowledge has to be filled in the future.

5.1.2. Small RNAs

Endogenous small RNAs (sRNAs) are essential components of the regulatory network of genes participating in host–microorganism interactions. These 20–25 nucleotide-long non-coding RNA molecules repress target gene expression at the transcriptional level via cleavage of the target transcript or at the posttranscriptional level via inhibition of translation [140]. Two types of sRNA molecules can be distinguished: microRNA (miRNA) and short interfering RNA (siRNA). miRNAs are single RNA molecules with stem loop secondary structures, which are encoded by MIR genes located in plant genomes. siRNAs are double-stranded RNA encoded by transposons, viruses, or heterochromatin [141]. In wheat, so far, the involvement of miRNAs during pathogenic fungi infections has been demonstrated [142,143,144,145]. Although the biogenesis pathways of plant miRNAs have been examined in detail and their contribution to the communication between host plants and pathogens has been exposed, knowledge of miRNA participation, regulation, and function in symbiotic plant–fungi interactions is still scarce and limited to the Medicago truncatula [146,147], Solanum lycopersicum [148], and Oryza sativa [149] plant species and to arbuscular mycorrhizal fungi (AMF). However, our very recent study suggests the role of miRNAs in establishing and/or maintaining the wheat–endogenous beneficial fungi interaction (data unpublished). Diverse expression patterns in the roots and leaves of three wheat miRNAs, viz., miR398, miR167, and 159, between Trichoderma inoculated (beneficial interaction) or F. culmorum inoculated (deleterious interaction) plants and control wheat as well as between Trichoderma inoculated and F. culmorum inoculated plants were noted. Interestingly, recent studies have discovered that miRNA molecules are transported between plants and microbes and triggered gene silencing as trans-regulators in interacting organisms [150]. The transport of host miRNAs into interacting fungal pathogens has also been observed in wheat–F. graminearum interactions, where wheat miR1023 suppressed the invasion of F. graminearum by targeting and silencing FGSG_03101, which codes an alpha/beta hydrolase gene in F. graminearum [150].

5.1.3. Long Non-Coding RNA (lncRNA)

Transcriptional regulation during wheat–fungi interaction can also be mediated by lncRNAs [151,152,153,154]. This group of non-coding RNAs exceed 200 nt in length and do not contain the significant open reading frame (ORF). Studies on maize indicated that lncRNAs participate in plant–beneficial fungal interactions. Sixty-three differentially expressed lncRNAs were identified in maize under beneficial interaction with the arbuscular mycorrhizal fungi Rhizophagus irregularis [154]. In wheat plants, the 254 and 52 lincRNAs (long intergenic ncRNA) responded to pathogenic B. graminis f. sp. tritici and P. striiformis f. sp. tritici infections, respectively [151]. The aforementioned report implies that not only pathogenic but also beneficial interaction with fungal symbionts may be controlled by lncRNAs, but more detailed studies are required.
The studies concerning wheat–fungi interaction should enter an epigenetic era to understand the role of epigenetic regulation in establishing and maintaining the beneficial, endosymbiotic interactions in non-model wheat. Considering the agronomical importance of wheat as well as the lack of detailed knowledge, explorations of the issue presented are needed.

6. Application of Endophytic Fungi in Modern Agronomy

Agriculture today faces the challenge of ensuring food security for the world population, which is estimated to grow from the current level of around 7 billion to 9 billion by 2050 [155]. However, contrary to all opinions, the use of ever greater doses of artificial fertilizers does not increase the yield; on the contrary, this causes a gradual reduction in soil fertility, reduces the quality of cultivated products, and increases environmental pollution. All these aspects have prompted scientists to look for not only an environmentally friendly alternative but also one that would meet the constantly growing demand for agricultural productivity. The interactions of endophytic fungi with crops is of benefit in this regard. These fungi support plant growth and increase tolerance to biotic and abiotic stresses. In sustainable agriculture, endophytes can be used primarily as protection for the host plant against pathogens or pests. They also increase the plant’s resistance to biotic and abiotic stress and affect plant growth and development [29,30,31,32,33,156]. Importantly, they can also support host plants using the metabolites excreted to accelerate the process of nutrient uptake from the environment [157].
The available literature has shown that, among the endophytes identified in wheat, several of them exhibit symbiotic cooperation with this plant. Previous studies by Dingle and McGee [85] on endophytic fungi showed that the Chaetomium sp. strain, which was obtained from healthy wheat leaves, contributed to the reduction in the number and development of rust pustules P. recondite f. sp. tritici. The endophytic strain of T. hamatum has been identified as a potential biocontrol agent against Pyrenophora tritici-repentis (Died.) Drechsler, the causal agent of tan spots of wheat [156]. Comby et al. [53] found endophytic fungi in wheat that could be used as a biological control agent against F. graminearum, the cause of Fusarium head blight (FHB). The identified strains belonged to the following species: S. kiliense, A. proteae, C. rosea, and M. bolleyi. Similarly, S. strictum, A. floculossa, and P. olsonii were documented as potential biocontrol agents of Fusarium head blight (FHB) caused by F. graminearum in wheat [50]. Disease severity and pathogen biomass inside the analyzed wheat spikes were reduced (70–80%) when the endophytic strains were inoculated at least two days before contact with the pathogen. Interestingly, the endophytic strains used did not present an antagonistic effect on F. graminearum during the in vitro dual culture experiment [50]. Furthermore, endophytic P. olsonii and A. alternatum were identified as biocontrol agents against Zymoseptoria tritici causing Septoria tritici blotch (STB) in wheat [79]. Additional inoculation with wheat endophytic fungi also alleviates a wheat plant’s tolerance to salt stress [41]. Under the conditions of moderate salinity, C. coarctatum and A. chlamydospora intensified the growth of wheat, while under conditions of strong salinity, only A. chlamydospora showed this effect. Moreover, A. chlamydospora and F. equiseti demonstrated the ability to enhance root growth under salt stress [41]. Three endophytes isolated from the roots of T. turgidum (referred to as 2206, 2210, and 2215 from the Saskatchewan Microbial Collection Database—SMCD) demonstrated improved tolerance for heat and drought in both parental and second generation durum wheat seeds [31,32]. The authors termed this type of cooperation mycovitality due to the protective fungal effect on seeds, maintaining their vitality and causing successful germination.
Worth noting is that wheat endophytes are also the substantial source of beneficial metabolite. Pipecolisporin was recently identified in Nigrospora oryzae cultures, isolated from Triticum sp. roots. This novel compound presents antimalarial and antitrypanosomal activities by exhibiting activity in the molecular range against tropical parasites: Plasmodium falciparum and Trypanosoma cruzi, respectively [157].
The ability to infect cultivated wheat with fungal endophytes that originated in other plant species has also been demonstrated, and various positive effects have been reported. For example, endophytes reduced wheat susceptibility to insects and pathogens [158,159,160], improved heat and drought tolerance [31,32], and promoted plant growth [161]. Serfling et al. [162] have documented the ability of the endophytic species Piriformospora indica to reduce common leaf, root, and stem disease symptoms in wheat caused by pathogens such as Pseudocercosporella herpotrichoides, B. graminis f. sp. tritici, and F. culmorum. Meanwhile, Malik et al. [163] showed that inoculation with endophytic fungus Trametes hirsuta, isolated from the Chenopodium album L. plant, may improve the survival of wheat plants in metal-contaminated soils and may additionally assist in the phytoextraction of heavy metals (Pb). Similar properties in relation to wheat plant were revealed for P. ruqueforti isolated from the endosphere of Solanum surattense [164]. Studies have shown that treating wheat plants grown in soils contaminated with heavy metals Ni, Cd, Cu, Zn, and Pb with P. ruqueforti increases their tolerance to stress and nutrient uptake.
The literature review presented above indicates that endophytic fungi isolated from wheat or other plants have much potential to be used in biological control or as plant growth stimulants. However, in order for these microorganisms to be used as bio-pesticides, bio-fungicides, or growth bio-stimulants, they must meet several requirements, such as not being harmful to plants, humans, and animals; effectiveness in controlling their target; the ability to survive in various conditions; and compatibility with the other biologically active substances used in the cultivation of wheat. In addition, their large-scale production should be economically viable [165]. Taking into account the above restrictions as well as the entire commercialization process, which includes the isolation of endophytic fungi; an evaluation of the bioagent’s effectiveness in in vitro, greenhouse, and field conditions; formulation and mass production development; delivery; compatibility; registration; and release [166,167], with large-scale wheat production, bringing such bio-products into the market is a very demanding endeavor. The available literature shows that few such products for wheat have been commercialized so far. Only the following products are documented: AQ10 (Ecogen, Inc, USA) based on Ampelomyces quisqualis [167]; Sporodex (Ecogen, Inc, USA) based on Pseudozyma flocculosa [167] for protection against mildew powdery; Biomal (Canada) based on Colletotrichum gloeosporioides f. sp. malvae [168] antagonistic to Malva pusilla (round-leaved mallow); Trichodex (Bio works, USA) based on Trichoderma harzianum T-39 [169]; Canna based on Trichoderma afroharzianum [170]; Trichosan (America) based on the CBS 134709 strain [168] antagonistic to Botrytis spp; and Promot WP (USA Canna International BV, NL-Breda, Vitalin Pflanzengesundheit GmbH, D-Ober-Ramstadt JH Biotech Inc., Ventura, CA, USA) based on Trichoderma simmonsii, CBS 134706 strain [168], and Trichoderma guizhouense, CBS 134707 strain [168], antagonistic to Fusarium sp., Phytophthora infestans, and Botrytis spp.

7. New Perspectives and Research Needs

Significant and continuous technological advances have contributed to the implementation of high-throughput methods over the last ten years for studying the microbiome of various crop species, including wheat. These technological solutions, more precisely NGS, were first used in wheat seed mycobiome research by Nicolaisen et al. [115]. Progress in understanding the complexity of the structure, dynamics, or determinants of changes in the communities of various groups of wheat-associated microorganisms was possible thanks to further research by Karlsson et al. [106,116], Hertz et al. [118], Granzow et al. [112], Gdanetz and Trail [77], Yashiro et al. [119], and Knorr et al. [121]. However, to our knowledge, the first studies of the wheat endosphere mycobiome using high-throughput techniques were published in 2016 by Ofek-Lalzar et al. [39]. The next ones were the work of Vujanovic et al. [108] and Latz et al. [79]. Despite these efforts, knowledge concerning the wheat endosphere mycobiome is still insufficient. Moreover, the synchronization of data obtained over the years and their co-interpretation are problematic, mainly due to the lack of consistent research standards on the wheat microbiome that would allow for a comparison of data from different laboratories, or the revision and integration of data generated from previous methods. With the purpose of mycobiomic research in mind, standardization should include developing the experimental design, adjusting the methodology and strategy for data analysis, interpretation, and integration. When determining the scale, frequency, and time span of sampling, the multidimensional plasticity of the mycobiome should also be taken into account so that both core and transient endophytic fungi in wheat can be identified. Because wheat is a crop, research into its microbiome usually aims to apply knowledge not only regarding the structure but also the function of fungi associated with its tissues. Therefore, comprehensive studies are recommended, including both high-throughput analyses and those based on classical methods enabling the isolation and direct characterization of endophytes. Recently, Kavamura et al. [122] suggested a multi-omic approach for the effective use of the wheat microbiome in efforts to increase the sustainable production of this grain. They proposed metagenomics as describing the structure and diversity of the microbiome, metatrancriptomics for the evaluation of active microorganisms or their genes, and culturomics and phenomics to isolate microorganism of interest and to detect their functional and metabolic activity. However, to prove the functional ability of the selected isolates, Kavamura et al. [122] advised the use of single-cell genomics to target genes of interest. To verify the effect of isolated microorganisms on wheat, they suggest that metaproteomic or metabolomic analyses should be performed on the plant level. Such a comprehensive approach could be adapted to the analysis of the structure and functionality of only the mycobiome of the wheat endosphere, taking into account its multidimensional plasticity. Here, however, a serious obstacle to obtaining complete knowledge is the inability of some endophytic fungi to live outside plant tissues. Therefore, a major challenge for the future is to develop a methodology to capture and maintain such microorganisms outside the plant system.
In order to obtain a complete understanding of the wheat mycobiom, all the factors that affect its formation and functioning should be taken into account. Kavamura et al. [122] declared that four types of factors determine the microbiome of wheat: antropogenic, edaphic, environmental, and host. When studying the fungi associated with the wheat endosphere, both these factors and the multidirectional interactions of plant–microorganism–microorganism, in line with the latest concept of meta-organisms or the so-called holobiont theory, are worth considering [171]. Currently, endophytic fungi, due to their “type of interaction” with plants or the remaining dead organic matter, are referred to as pathogens, symbiotes, saprophytes, or those whose function has not yet been understood. Until sufficient knowledge regarding the real interactions of these fungi with plants in the endosphere, including the molecular basis of these interactions and their effects on the plant, is presented, this classification should be abandoned. Moreover, the result of the interaction is determined by the entire holobiont of the plant and species, which, apart from this holobiont, exhibits pathogenic features and may perform completely different functions in its endosphere.
Ultimately, research on mycobiomes aims to improve the functioning of the wheat holobiont, stimulating plant germination and growth, providing nutrients, increasing resistance to biotic and abiotic stress factors, and increasing productivity, i.e., yielding. Precision farming, which aims to use a new generation of targeted inoculants based on microorganisms or their metabolites, is a new perspective. Inoculants based on one microorganism or consortia of different microorganisms, closely matched to the host organism or the growing conditions, are considered. The development of such products requires a large amount of research; recognition of their effects on the plant and environment; stability in the target environment; and in the case of microbial consortia, no antagonistic behavior. Such an approach is necessary for the microbiome to be successfully and fully implemented in agriculture and precision farming.
Another perspective based on symbiotic communication between endophytic fungi and wheat is provided by RNAi technologies. This technology uses an internal RNA interference mechanism (RNAi) that occurs in nearly all eukaryotes in which target mRNAs are degraded or functionally suppressed [172]. Two dsRNA delivery strategies are suggested to protect wheat from pathogenic fungal diseases: the transgene-based host-induced gene silencing (HIGS) strategy, which is based on the expression of hairpin RNA or small RNA directed to silence genes in pathogens and pests in a plant, and spray-induced gene silencing (SIGS), which use RNAi-based products such as dsRNA or sRNA that are derived from microorganisms and, accordingly, capable of controlling pests and pathogens present in the phylosphere [170,172]. The latter strategy is considered environmentally friendly and was the first applied by Koch et al. [173], who used foliar application of dsRNA targeting the cytochrome P450 (CYP3) gene in F. graminearum and observed reductions in pathogen growth in directly sprayed leaves as well as in distal untreated leaves of barley plants. Unfortunately, in wheat, research on the use of the SIGS strategy is still in its infancy, especially in terms of interactions with symbiotic endophytic fungi.

8. Conclusions

Due to the use of wheat as food for humans and livestock, its importance for global food production and global food security and the risks caused by adverse environmental conditions, changing climate, emerging pathogens, and diseases have been extensively discussed. Modern agriculture and the cultivation of wheat therefore face many challenges in order to avoid these threats. One of the ways to reduce the effects of abiotic stresses and to reduce the occurrence of pathogens and related diseases is to use symbiotic endophytic fungi inhabiting the internal tissues of wheat. Numerous studies have been carried out on the use of these microorganisms in biological control or as plant growth biostimulants, although, in the case of wheat, the scope of this work is still insufficient and has not resulted in beneficial solutions in disease management and integrated plant protection. These and the abovementioned studies could contribute to the provision of new tools that can be used for modern agriculture and the management of large-scale wheat cultivation in the world.

Author Contributions

Conceptualization, L.B. and S.S.; writing—original draft preparation, S.S. (Section 1, Section 4 and Section 5); K.M. (Section 2, Section 3 and Section 6); L.B. (Section 7 and Section 8); writing—review and editing, L.B.; visualization, S.S. and K.M.; supervision, L.B.; funding acquisition, L.B. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the National Science Center in Poland grant No. 2017/27/B/NZ9/01591.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable. No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Grigoriev, I. Fungal genomics for energy and environment. In Genomics of Soil- and Plant-Associated Fungi: Soil Biology; Horwitz, B.A., Mukherjee, P.K., Mukherjee, M., Kubicek, C.P., Eds.; Springer: Berlin/Heidelberg, Germany, 2013; Volume 36, pp. 11–27. [Google Scholar]
  2. Petrini, O. Fungal endophytes of tree leaves. In Microbial Ecology of Leaves; Andrews, J.H., Hirano, S.S., Eds.; Brock/Springer Series in Contemporary Bioscience; Springer: New York, NY, USA, 1991; pp. 179–197. [Google Scholar]
  3. Hyde, K.D.; Soytong, K. The fungal endophyte dilemma. Fungal Divers. 2008, 33, 163–173. [Google Scholar]
  4. Porras-Alfaro, A.; Bayman, P. Hidden fungi, emergent properties: Endophytes and microbiomes. Annu. Rev. Phytopathol. 2011, 49, 291–315. [Google Scholar] [CrossRef] [Green Version]
  5. Schardl, C.L.; Leuchtmann, A.; Spiering, M.J. Symbioses of grasses with seedborne fungal endophytes. Annu. Rev. Plant Biol. 2004, 55, 315–340. [Google Scholar] [CrossRef] [PubMed]
  6. Parniske, M. Arbuscular mycorrhiza: The mother of plant root endosymbioses. Nat. Rev. Microbiol. 2008, 6, 763–775. [Google Scholar] [CrossRef]
  7. Druzhinina, I.S.; Seidl-Seiboth, V.; Herrera-Estrella, A.; Horwitz, B.A.; Kenerley, C.M.; Monte, E.; Mukherjee, P.K.; Zeilinger, S.; Grigoriev, I.V.; Kubicek, C.P. Trichoderma: The genomics of opportunistic success. Nat. Rev. Microbiol. 2011, 9, 749–759. [Google Scholar] [CrossRef] [PubMed]
  8. Hardoim, P.R.; van Overbeek, L.S.; Berg, G.; Pirttiläd, A.M.; Compante, S.; Campisanof, A.; Döringg, M.; Sessitsche, A. The hidden world within plants: Ecological and evolutionary considerations for defining functioning of microbial endophytes. Microbiol. Mol. Biol. Rev. 2015, 79, 293–320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Saikkonen, K.; Wali, P.R.; Helander, M. Genetic compatibility determines endophyte-grass combinations. PLoS ONE 2010, 5, e11395. [Google Scholar] [CrossRef] [Green Version]
  10. Kale, S.D.; Tyler, B.M. Entry of oomycete and fungal effectors into plant and animal host cells. Cell. Microbiol. 2011, 13, 1839–1848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Rodriguez, R.J.; White, J.F., Jr.; Arnold, A.E.; Redman, R.S. Fungal endophytes: Diversity and functional roles. N. Phytol. 2009, 182, 314–330. [Google Scholar] [CrossRef]
  12. Bacon, C.W.; White, J. (Eds.) Physiological adaptations in the evolution of endophytism in the Clavicipitaceae. In Microbial Endophytes; Taylor & Francis: New York, NY, USA, 2000; pp. 237–263. [Google Scholar]
  13. Saikkonen, K.I.D.; Gyllenberg, M. The persistence of vertically transmitted fungi in grass metapopulations. Proc. Biol. Sci. 2002, 269, 1397–1403. [Google Scholar] [CrossRef] [Green Version]
  14. Tadych, M.; Bergen, M.; Dugan, F.M.; White, J.F. Evaluation of the potential role of water in spread of conidia of the Neotyphodium endophyte of Poa ampla. Mycol. Res. 2007, 111, 466–472. [Google Scholar] [CrossRef]
  15. Gao, K.; Mendgen, K. Seed-transmitted beneficial endophytic Stagonospora sp can penetrate the walls of the root epidermis, but does not proliferate in the cortex, of Phragmites australis. Can. J. Bot. 2006, 84, 981–988. [Google Scholar] [CrossRef] [Green Version]
  16. Redman, R.S.; Sheehan, K.B.; Stout, R.G.; Rodriguez, R.J.; Henson, J.M. Thermotolerance generated by plant/fungal symbiosis. Science 2002, 298, 1581. [Google Scholar] [CrossRef] [PubMed]
  17. Higgins, K.L.; Arnold, A.E.; Miadlikowska, J.; Sarvate, S.D.; Lutzoni, F. Phylogenetic relationships, host affinity, and geographic structure of boreal and arctic endophytes from three major plant lineages. Mol. Phylogenet. Evol. 2007, 42, 543–555. [Google Scholar] [CrossRef] [PubMed]
  18. Davis, E.C.; Shaw, A.J. Biogeographic and phylogenetic patterns in diversity of liverwort associated endophytes. Am. J. Bot. 2008, 95, 914–924. [Google Scholar] [CrossRef]
  19. Kumar, D.S.S.; Hyde, K.D. Biodiversity and tissue-recurrence of endophytic fungi in Tripterygium wilfordii. Fungal Divers. 2004, 17, 69–90. [Google Scholar]
  20. O’Dell, T.E.; Massicotte, H.B.; Trappe, J.M. Root colonization of Lupinus latifolius Agardh. and Pinus contorta Dougl. by Phialocephala fortinii Wang & Wilcox. N. Phytol. 1993, 124, 93–100. [Google Scholar]
  21. Espinosa-Garcia, F.J.; Langenheim, J.H. The endophytic fungal community in leaves of a coastal redwood population diversity and spatial patterns. N. Phytol. 1990, 116, 89–97. [Google Scholar] [CrossRef]
  22. Li, W.C.; Zhou, J.; Guo, S.Y.; Guo, L.D. Endophytic fungi associated with lichens in Baihua mountain of Beijing, China. Fungal Divers. 2007, 25, 69–80. [Google Scholar]
  23. Leuchtmann, A. Systematics, distribution, and host specificity of grass endophytes. Nat. Toxins 1993, 1, 150–162. [Google Scholar] [CrossRef] [PubMed]
  24. Kai, W.; Zhiwei, Z. Occurrence of arbuscular mycorrhizas and dark septate endophytes in hydrophytes from lakes and streams in southwest China. Int. Rev. Hydrobiol. 2006, 91, 29–37. [Google Scholar] [CrossRef]
  25. Arnold, A.E.; Maynard, Z.; Gilbert, G.S.; Coley, P.D.; Kursar, T.A. Are tropical fungal endophytes hyperdiverse? Ecol. Lett. 2000, 3, 267–274. [Google Scholar] [CrossRef]
  26. Li, H.Y.; Shen, M.; Zhou, Z.P.; Li, T.; Wei, Y.L.; Lin, L.B. Diversity and cold adaptation of endophytic fungi from five dominant plant species collected from the Baima Snow Mountain, Southwest China. Fungal Divers. 2012, 54, 79–86. [Google Scholar] [CrossRef]
  27. Innerebner, G.; Knief, C.; Vorholt, J.A. Protection of Arabidopsis thaliana against leaf-pathogenic Pseudomonas syringae by Sphingomonas strains in a controlled model system. Appl. Environ. Microbiol. 2011, 77, 3202–3210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Ritpitakphong, U.; Falquet, L.; Vimoltust, A.; Berger, A.; Métraux, J.P.; L’Haridon, F. The microbiome of the leaf surface of Arabidopsis protects against a fungal pathogen. N. Phytol. 2016, 210, 1033–1043. [Google Scholar] [CrossRef] [Green Version]
  29. Herre, E.A.; Mejía, L.C.; Kyllo, D.A.; Rojas, E.; Maynard, Z.; Butler, A.; Van Bael, S.A. Ecological implications of anti-pathogen effects of tropical fungal endophytes and mycorrhizae. Ecology 2007, 88, 550–558. [Google Scholar] [CrossRef] [Green Version]
  30. Gao, F.K.; Dai, C.C.; Liu, X.Z. Mechanisms of fungal endophytes in plant protection against pathogens. Afr. J. Microbiol. Res. 2010, 4, 1346–1351. [Google Scholar]
  31. Hubbard, M.; Germida, J.; Vujanovic, V. Fungal endophytes improve wheat seed germination under heat and drought stress. Botany 2012, 90, 137–149. [Google Scholar] [CrossRef]
  32. Hubbard, M.; Germida, J.J.; Vujanovic, V. Fungal endophytes enhance wheat heat and drought tolerance in terms of grain yield and second-generation seed viability. J. Appl. Microbiol. 2014, 116, 109–122. [Google Scholar] [CrossRef]
  33. Waqas, M.; Khan, A.L.; Kamran, M.; Hamayun, M.; Kang, S.M.; Kim, Y.H.; Lee, I.J. Endophytic fungi produce gibberellins and indoleacetic acid and promotes host-plant growth during stress. Molecules 2012, 17, 10754–10773. [Google Scholar] [CrossRef] [PubMed]
  34. © FAO. Crops and Livestock Products. 2021. Available online: http://www.fao.org/faostat/en/#data/QCL/visualize (accessed on 6 October 2021).
  35. Mäkinen, H.; Kaseva, J.; Trnka, M.; Balek, J.; Kersebaum, K.C.; Nendel, C.; Gobin, A.; Olesen, J.E.; Bindi, M.; Ferrise, R.; et al. Sensitivity of European wheat to extreme weather. Field Crop. Res. 2018, 222, 209–217. [Google Scholar] [CrossRef]
  36. Araus, J.L.; Slafer, G.A.; Royo, C.; Serret, M.D. Breeding for yield potential and stress adaptation in cereals. Crit. Rev. Plant Sci. 2008, 27, 377–412. [Google Scholar] [CrossRef]
  37. Sieber, T.N. Endophytic fungi in forest trees: Are they mutualists? Fungal Biol. Rev. 2007, 21, 75–89. [Google Scholar] [CrossRef]
  38. Comby, M.; Lacoste, S.; Baillieul, F.; Profizi CDupont, J. Spatial and temporal variation of cultivable communities of co-occurring endophytes and pathogens in wheat. Front. Microbiol. 2016, 7, 403. [Google Scholar] [CrossRef] [Green Version]
  39. Ofek-Lalzar, M.; Gur, Y.; Ben-Moshe, S.; Sharon, O.; Kosman, E.; Mochli, E.; Sharon, A. Diversity of fungal endophytes in recent and ancient wheat ancestors Triticum dicoccoides and Aegilops sharonensis. FEMS Microbiol. Ecol. 2016, 92, 152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Ripa, F.A.; Cao, W.D.; Tong, S.; Sun, J.G. Assessment of plant growth promoting and abiotic stress tolerance properties of wheat endophytic fungi. Biomed. Res. Int. 2019, 2019, 6105865. [Google Scholar] [CrossRef]
  41. Bouzouina, M.; Kouadria, R.; Lotmani, B. Fungal endophytes alleviate salt stress in wheat in terms of growth, ion homeostasis and osmoregulation. J. Appl. Microbiol. 2021, 130, 913–925. [Google Scholar] [CrossRef] [PubMed]
  42. Gao, X.-X.; Zhou, H.; Xu, D.-Y.; Yu, C.-H.; Chen, Y.-Q.; Qu, L.-H. High diversity of endophytic fungi from the pharmaceutical plant, Heterosmilax japonica Kunth revealed by cultivation-independent approach. FEMS Microbiol. Lett. 2005, 249, 255–266. [Google Scholar] [CrossRef] [Green Version]
  43. Arnold, A.E.; Henk, D.A.; Eells, R.L.; Lutzoni, F.; Vilgalys, R. Diversity and phylogenetic affinities of foliar fungal endophytes in loblolly pine inferred by culturing and environmental PCR. Mycologia 2007, 99, 185–206. [Google Scholar] [CrossRef]
  44. Sun, X.; Kosman, E.; Sharon, O.; Ezrati, S.; Sharon, A. Significant host- and environment-dependent differentiation among highly sporadic fungal endophyte communities in cereal crops-related wild grasses. Environ. Microbiol. 2020, 20, 3357–3374. [Google Scholar] [CrossRef]
  45. Gamboa, M.A.; Laureano, S.; Bayman, P. Measuring diversity of endophytic fungi in leaf fragments: Does size matter? Mycopathologia 2003, 156, 41–45. [Google Scholar] [CrossRef]
  46. Larran, S.; Perello, A.; Simon, M.R.; Moreno, V. Isolation and analysis of endophytic microorganisms in wheat (Triticum aestivum L.) leaves. World J. Microbiol. Biotech. 2002, 18, 683–686. [Google Scholar] [CrossRef]
  47. Larran, S.; Perello, A.; Simon, M.R.; Moreno, V. The endophytic fungi from wheat (Triticum aestivum L.). World J. Microbiol. Biotechnol. 2007, 23, 565–572. [Google Scholar] [CrossRef]
  48. Cłapa, T.; Mikołajczak, K.; Błaszczyk, L.; Narożna, D. Development of high-resolution melting PCR (HRM-PCR) assay to identify native fungal species associated with the wheat endosphere. J. Appl. Genet. 2020, 61, 629–635. [Google Scholar] [CrossRef]
  49. Salamon, S.; Mikołajczak, K.; Błaszczyk, L.; Ratajczak, K.; Sulewska, H. Changes in root-associated fungal communities in Triticum aestivum ssp. spelta L. and Triticum aestivum ssp. vulgare L. under drought stress and in various soil processing. PLoS ONE 2020, 15, e0240037. [Google Scholar] [CrossRef] [PubMed]
  50. Rojas, E.C.; Jensen, B.; Jørgensen, H.J.; Latz, M.A.; Esteban, P.; Ding, Y.; Collinge, D.B. Selection of fungal endophytes with biocontrol potential against Fusarium head blight in wheat. Biol. Control. 2020, 144, 104222. [Google Scholar] [CrossRef]
  51. Sun, X.; Guo, L.-D. Endophytic fungal diversity: Review of traditional and molecular techniques. Mycology 2012, 3, 65–76. [Google Scholar]
  52. Torres, M.; White, J.F.; Tadych, M.; Bills, G. Isolation and Identification of Fungal Endophytes. Prospects and Applications for Plant-Associated Microbes. A laboratory Manual, Part B: Fungi; Paimo, BBI (Biobien Innovations): Turku, Finland, 2011; pp. 1–28. [Google Scholar]
  53. Comby, M.; Gacoin, M.; Robineau, M.; Rabenoelina, F.; Ptas, S.; Dupont, J.; Profizi, C.; Baillieul, F. Screening of wheat endophytes as biological control agents against Fusarium head blight using two different in vitro tests. Microbiol. Res. 2017, 202, 11–20. [Google Scholar] [CrossRef] [PubMed]
  54. Trivedi, P.; Leach, J.E.; Tringe, S.G.; Sa, T.; Singh, B.K. Plant–microbiome interactions: From community assembly to plant health. Nat. Rev. Microbiol. 2020, 18, 607–662. [Google Scholar] [CrossRef]
  55. Geiser, D.M. Agriculture, and medicine. In Advances in Fungal Biotechnology for Industry; Springer: Berlin/Heidelberg, Germany, 2004; pp. 3–14. [Google Scholar]
  56. Lutzoni, F.; Kauff, F.; Cox, C.J.; McLaughlin, D.; Celio, G.; Dentinger, B.; Padamsee, M.; Hibbett, 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] [PubMed]
  57. Hughes, K.W.; Petersen, R.H.; Lodge, D.J.; Bergemann, S.E.; Baumgartner, K.; Tulloss, R.E.; Lickey, E.; Cifuentes, J. Evolutionary consequences of putative intra-and interspecific hybridization in agaric fungi. Mycologia 2013, 105, 1577–1594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Brun, S.; Silar, P. Evolutionary Biology—Concepts, Molecular and Morphological Evolution; Pontarotti, P., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; pp. 317–328. [Google Scholar]
  59. 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]
  60. Letourneau, A.; Seena, S.; Marvanová, L.; Bärlocher, F. Potential use of barcoding to identify aquatic hyphomycetes. Fungal Divers. 2010, 40, 51–64. [Google Scholar] [CrossRef]
  61. White, T.J.; Bruns, T.; Lee, S.; Tailor, 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]
  62. Ma, Z.; Michailides, T.J. A PCR-based technique for identification of Fusicoccum sp. from pistachio and various other hosts in California. Plant Dis. 2002, 86, 515–520. [Google Scholar] [CrossRef]
  63. Kuzdraliński, A.; Kot, A.; Szczerba, H.; Nowak, M.; Muszyńska, M. A review of Conventional PCR assays for the detection of selected phytopathogens of wheat. J. Mil. Microbiol. Biotechnol. 2017, 27, 175–189. [Google Scholar] [CrossRef]
  64. Sharma, R.; Polkade, A.V.; Shouche, Y.S. “Species Concept” in microbial taxonomy and systematics. Curr. Sci. 2015, 108, 1804–1814. [Google Scholar]
  65. Vu, D.; Groenewald, M.; de Vries, M.; Gehrmann, T.; Stielow, B.; Eberhardt, U.; Al-Hatmi, A.; Groenewald, J.Z.; Cardinali, G.; Houbraken, J.; et al. Large-scale generation and analysis of filamentous fungal DNA barcodes boosts coverage for kingdom fungi and reveals thresholds for fungal species and higher taxon delimitation. Stud. Mycol. 2019, 92, 135–154. [Google Scholar] [CrossRef]
  66. Nilsson, R.H.; Hyde, K.D.; Pawłowska, J.; Ryberg, M.; Tedersoo, L.; Aas, A.B.; Alias, S.A.; Alves, A.; Anderson, C.L.; Antonelli, A.; et al. Improving ITS sequence data for identification of plant pathogenic fungi. Fungal Divers. 2014, 67, 11–19. [Google Scholar] [CrossRef]
  67. Sun, X.; Guo, L.D.; Hyde, K.D. Community composition of endophytic fungi in Acer truncatum and their role in decomposition. Fungal Divers. 2011, 47, 85–95. [Google Scholar] [CrossRef]
  68. Glass, N.L.; Donaldson, G.C. Development of primer sets designed for use with the PCR to amplify conserved genes from filamentous ascomycetes. Appl. Environ. Microbiol. 1995, 61, 1323–1330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Carbone, I.; Kohn, L.M. A method for designing primer sets for speciation studies in filamentous ascomycetes. Mycologia 1999, 91, 553–556. [Google Scholar] [CrossRef]
  70. Hirt, R.P.; Logsdon, J.M., Jr.; Healy, B.; Dorey, M.W.; Doolittle, W.F.; Embley, T.M. Microsporidia are related to Fungi: Evidence from the largest subunit of RNA polymerase II and other proteins. Proc. Natl. Acad. Sci. USA 1999, 19, 580–585. [Google Scholar] [CrossRef] [Green Version]
  71. Hong, S.B.; Go, S.J.; Shin, H.D.; Frisvad, J.C.; Samson, R.A. Polyhphasic taxonomy of Aspergillus fumigatus and related species. Mycologia 2005, 97, 1316–1329. [Google Scholar] [CrossRef] [PubMed]
  72. Maciá-Vicente, J.G.; Jansson, H.B.; Abdullah, S.K.; Descals, E.; Salinas, J.; Lopez-Llorca, L.V. Fungal root endophytes from natural vegetation in Mediterranean environments with special reference to Fusarium spp. FEMS Microbiol. Ecol. 2008, 64, 90–105. [Google Scholar] [CrossRef] [Green Version]
  73. Brun, S.; Madrid, H.; Gerrits Van Den Ende, B.; Andersen, B.; Marinach-Patrice, C.; Mazier, D.; De Hoog, G.S. Multilocus phylogeny and MALDI-TOF analysis of the plant pathogenic species Alternaria dauci and relatives. Fungal Biol. 2013, 117, 32–40. [Google Scholar] [CrossRef] [PubMed]
  74. Sharma, R.; Kulkarni, G.; Sonawane, M.S.; Shouche, Y.S. A new endophytic species of Chaetomium from Jatropha podagrica. Mycotaxon 2013, 124, 117–126. [Google Scholar] [CrossRef]
  75. Slippers, B.; Boissin, E.; Phillips, A.J.L.; Groenewald, J.Z.; Wingfield, M.J.; Postma, A.; Burgess, T.; Crous, P.W. Phylogenetic lineages in the Botryosphaeriales: A systematic an evolutionary framework. Stud. Mycol. 2013, 76, 31–49. [Google Scholar] [CrossRef] [Green Version]
  76. U’Ren, J.M.; Dalling, J.W.; Gallery, R.E.; Maddison, D.R.; Davis, E.C.; Gibson, C.M.; Arnold, A.E. Diversity and evolutionary origins of fungi associated with seeds of a neotropical pioneer tree: A case study for analysing fungal environmental samples. Mycol. Res. 2009, 113, 432–449. [Google Scholar] [CrossRef]
  77. Gdanetz, K.; Trail, F. Wheat microbiome under four management strategies, and potential for endophytes in disease protection. Phytobiomes 2017, 1, 158–168. [Google Scholar] [CrossRef] [Green Version]
  78. Llorens, E.; Sharon, O.; Camañes, G.; García-Agustín, P.; Sharon, A. Endophytes from wild cereals protect wheat plants from drought by alteration of physiological responses of the plants to water stress. Environ. Microbiol. 2019, 21, 3299–3312. [Google Scholar] [CrossRef]
  79. Latz, M.A.; Jensen, B.; Collinge, D.B.; Jørgensen, H.J.L. Identification of two endophytic fungi that control Septoria tritici blotch in the field, using a structured screening approach. Biol. Control. 2020, 141, 104128. [Google Scholar] [CrossRef]
  80. Dissanayake, A.J.; Purahong, W.; Wubet, T.; Hyde, K.D.; Zhang, W.; Xu, H.; Zhang, G.; Fu, C.; Liu, M.; Xing, Q.; et al. Direct comparison of culture-dependent and culture-independent molecular approaches reveal the diversity of fungal endophytic communities in stems of grapevine (Vitis vinifera). Fungal Divers. 2018, 90, 85–107. [Google Scholar] [CrossRef]
  81. Kraková, L.; Šoltys, K.; Otlewska, A.; Pietrzak, K.; Purkrtová, S.; Savická, D.; Puškárová, A.; Bučková, M.; Szemes, T.; Budiš, J.; et al. Comparison of methods for identification of microbial communities in book collections: Culture-dependent (sequencing and MALDI-TOF MS) and culture-independent (Illumina MiSeq). Int. Biodeterior. Biodegrad. 2018, 131, 51–59. [Google Scholar] [CrossRef]
  82. Mendoza, L.M.; Neef, A.; Vignolo, G.; Belloch, C. Yeast diversity during the fermentation of Andean chicha: A comparison of high-throughput sequencing and culturedependent approaches. Food Microbiol. 2017, 67, 1–10. [Google Scholar] [CrossRef]
  83. Malcolm, G.M.; Kuldau, G.A.; Gugino, B.K.; Jimenez-Gasco, M.D. Hidden host plant associations of soil borne fungal pathogens: Anecological perspective. Phytopathology 2013, 103, 538–544. [Google Scholar] [CrossRef] [Green Version]
  84. Zabalgogeazcoa, I. Fungal endophytes and their interaction with plant pathogens: A review. Span. J. Agric. 2008, 6, 138–146. [Google Scholar] [CrossRef] [Green Version]
  85. Dingle, J.; Mcgee, P.A. Some endophytic fungi reduce the density of pustules of Puccinia recondita f. sp. tritici in wheat. Mycol. Res. 2003, 107, 310–316. [Google Scholar] [CrossRef] [PubMed]
  86. Liu, C.H.; Zou, W.X.; Lu, H.; Tan, R.X. Antifungal activity of Artemisia annua endophyte cultures against phytopathogenic fungi. J. Biotechnol. 2001, 88, 277–282. [Google Scholar] [CrossRef]
  87. Rivera-Varas, V.V.; Freeman, T.A.; Gudmestad, N.C.; Secor, G.A. Mycoparasitism of Helminthosporium solani by Acremonium strictum. Phytopathology 2007, 97, 1331–1337. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Schulz, B.; Boyle, C. The endophytic continuum. Mycol. Res. 2005, 109, 661–686. [Google Scholar] [CrossRef] [Green Version]
  89. Card, S.D.; Faville, M.J.; Simpson, W.R.; Johnson, R.D.; Voisey, C.R.; de Bonth, A.C.M.; Hume, D.E. Mutualistic fungal endophytes in the Triticeae—Survey and description. FEMS Microbiol. Ecol. 2014, 88, 94–106. [Google Scholar] [CrossRef] [Green Version]
  90. Xu, K.G.; Jiang, Y.M.; Li, Y.K.; Xu, Q.Q.; Niu, J.S.; Zhu, X.X.; Li, Q.Y. Identification and pathogenicity of fungal pathogens causing black point in wheat on the north China plain. Indian J. Microbiol. 2018, 58, 159–164. [Google Scholar] [CrossRef]
  91. Zhan, G.; Tian, Y.; Wang, F.; Chen, X.; Guo, J.; Jiao, M.; Huang, L.; Kang, Z. A novel fungal hyperparasite of Puccinia striiformis f. sp. tritici, the causal agent of wheat stripe rust. PLoS ONE 2014, 9, e111484. [Google Scholar] [CrossRef]
  92. Vujanovic, V.; Mavragani, D.; Hamel, C. Fungal communities associated with durum wheat production system: A characterization by growth stage, plant organ and preceding crop. Crop Prot. 2012, 37, 26–34. [Google Scholar] [CrossRef]
  93. Taheri, A.E.; Hamel, C.; Gan, Y. Cropping practices impact fungal endophytes and pathogens in durum wheat roots. Appl. Soil Ecol. 2016, 100, 104–111. [Google Scholar] [CrossRef]
  94. Al-Sadi, A.M. Bipolaris sorokiniana-induced black point, common root rot, and spot blotch diseases of wheat: A review. Front. Cell. Infect. Microbiol. 2021, 11, 118. [Google Scholar] [CrossRef] [PubMed]
  95. Yue, H.-M.; Wang, M.; Gong, W.-F.; Zhang, L.-O. The screening and identification of the biological control fungi Chaetomium spp. against wheat common root rot. FEMS Microbiol. Lett. 2018, 365, fny242. [Google Scholar] [CrossRef]
  96. Istifadah, N.; McGee, P.A. Endophytic Chaetomium globosum reduces development of tan spot in wheat caused by Pyrenophora tritici-repentis. Australas. Plant Pathol. 2006, 35, 411–418. [Google Scholar] [CrossRef]
  97. Roberti, R.; Veronesi, A.R.; Cesari, A.; Cascone, A.; Di Berardina, I.; Bertini, L.; Caruso, C. Induction of PR proteins and resistance by the biocontrol agent Clonostachys rosea in wheat plants infected with Fusarium culmorum. Plant Sci. 2008, 175, 339–347. [Google Scholar] [CrossRef]
  98. Banerjee, A.; Mittra, B. Morphological modification in wheat seedlings infected by Fusarium oxysporum. Eur. J. Plant Pathol. 2018, 152, 521–524. [Google Scholar] [CrossRef]
  99. Alkan, M.; Özer, G.; İmren, M.; Özdemir, F.; Morgounov, A.; Dababat, A.A. First Report of Fusarium culmorum and Microdochium bolleyi Causing Root Rot on Triticale in Kazakhstan. Plant Dis. 2021, 1. [Google Scholar] [CrossRef]
  100. Littlefield, L.J. Biology of the Plant Rust: An Introduction; Iowa State University Press Ames: Iowa City, IA, USA, 1981; p. 103. [Google Scholar]
  101. Bartosiak, S.F.; Arseniuk, E.; Szechyńska-Hebda, M.; Bartosiak, E. Monitoring of natural occurrence and severity of leaf and glume blotch diseases of winter wheat and winter triticale incited by necrotrophic fungi Parastagonospora spp. and Zymoseptoria tritici. Agronomy 2021, 11, 967. [Google Scholar] [CrossRef]
  102. Abdel-Kareem, M.M.; Zohri, A.N.A.; Nasr, S.A.E.E. Novel marine yeast strains as plant growth-promoting agents improve defense in wheat (Triticum aestivum) against Fusarium oxysporum. J. Plant Dis. Prot. 2021, 128, 973–988. [Google Scholar] [CrossRef]
  103. Ponomarenko, A.; Goodwin, S.B.; Kema, G.H.J. Septoria tritici blotch (STB) of wheat. Plant Health Instr. 2011. [Google Scholar] [CrossRef]
  104. Błaszczyk, L.; Basińska-Barczak, A.; Ćwiek-Kupczyńska, H.; Gromadzka, K.; Popiel, D.; Stępień, Ł. Suppressive effect of Trichoderma spp. on toxigenic Fusarium species. Pol. J. Microbiol. 2017, 66, 85–100. [Google Scholar] [CrossRef] [Green Version]
  105. Rascovan, N.; Carbonetto, B.; Perrig, D.; Diaz, M.; Canciani, W.; Abalo, M.; Alloati, J.; Gonzalez-Ant, G.; Vazquez, M.P. Integrated analysis of root microbiomes of soybean and wheat from agricultural fields. Sci. Rep. 2016, 6, 1–12. [Google Scholar] [CrossRef]
  106. Karlsson, I.; Friberg, H.; Kolseth, A.K.; Steinberg, C.; Persson, P. Organic farming increases richness of fungal taxa in the wheat phyllosphere. Mol. Ecol. 2017, 26, 3424–3436. [Google Scholar] [CrossRef]
  107. Sapkota, R.; Jørgensen, L.N.; Nicolaisen, M. Spatiotemporal variation and networks in the mycobiome of the wheat canopy. Front. Plant Sci. 2017, 8, 1357. [Google Scholar] [CrossRef] [Green Version]
  108. Vujanovic, V.; Kim, S.H.; Lahlali, R.; Karunakaran, C. Spectroscopy and SEM imaging reveal endosymbiont-dependent components changes in germinating kernel through direct and indirect coleorhiza-fungus interactions under stress. Sci. Rep. 2019, 9, 1665. [Google Scholar] [CrossRef]
  109. Mahoney, A.K.; Yin, C.; Hulbert, S.H. Community structure, species variation, and potential functions of rhizosphere-associated bacteria of different winter wheat (Triticum aestivum) cultivars. Front. Plant Sci. 2017, 8, 132. [Google Scholar] [CrossRef] [Green Version]
  110. Ofek, M.; Voronov-Goldman, M.; Hadar, Y.; Minz, D. Host signature effect on plant root-associated microbiomes revealed through analyses of resident vs. active communities. Environ. Microbiol. 2013, 16, 2157–2167. [Google Scholar] [CrossRef]
  111. Yin, C.; Mueth, N.; Hulbert, S.; Schlatter, D.; Paulitz, T.C.; Schroeder, K.; Prescott, A.; Dhingra, A. Bacterial communities on wheat grown under long-term conventional tillage and no-tiol in the Pacific Northwest of the United States. Phytobiomes 2017, 1, 83–90. [Google Scholar] [CrossRef]
  112. Granzow, S.; Kaiser, K.; Wemheuer, B.; Pfeiffer, B.; Daniel, R.; Vidal, S.; Wemheuer, F. The effects of cropping regimes on fungal and bacterial communities of wheat and faba bean in a greenhouse pot experiment differ between plant species and compartment. Front. Microbiol. 2017, 8, 902. [Google Scholar] [CrossRef] [Green Version]
  113. Huang, Y.; Kuang, Z.; Wang, W.; Cao, L. Exploring potential bacterial and fungal biocontrol agents transmitted from seeds to sprouts of wheat. Biol. Control. 2016, 98, 27–33. [Google Scholar] [CrossRef]
  114. Kothe, E.; Turnau, K. Mycorrhizosphere communication: Mycorrhizal fungi and endophytic fungus-plant interactions. Front. Microbiol. 2018, 9, 3015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Nicolaisen, M.; Justesen, A.F.; Knorr, K.; Wang, J.; Pinnschmidt, H.O. Fungal communities in wheat grain show significant co-existence patterns among species. Fungal Ecol. 2014, 11, 145–153. [Google Scholar] [CrossRef]
  116. Karlsson, I.; Friberg, H.; Steinberg, C.; Persson, P. Fungicide effects on fungal community composition in the wheat phyllosphere. PLoS ONE 2014, 9, e0111786. [Google Scholar] [CrossRef] [Green Version]
  117. Sapkota, R.; Knorr, K.; Jørgensen, L.N.; O’Hanlon, K.A.; Nicolaisen, M. Host genotype is an important determinant of the cereal phyllosphere mycobiome. N. Phytol. 2015, 207, 1134–1144. [Google Scholar] [CrossRef]
  118. Hertz, M.; Jensen, I.R.; Jensen, L.Ø.; Thomsen, S.N.; Winde, J.; Dueholm, M.S.; Sørensen, L.H.; Wollenberg, R.D.; Sørensen, H.O.; Sondergaard, T.E.; et al. The fungal community changes over time in developing wheat heads. Int. J. Food Microbiol. 2016, 222, 30–39. [Google Scholar] [CrossRef] [PubMed]
  119. Yashiro, E.; Savova-Bianchi, D.; Niculita-Hirzel, H. Major differences in the diversity of mycobiomes associated with wheat processing and domestic environments: Significant findings from high-throughput sequencing of fungal barcode ITS1. Int. J. Environ. Res. Public Health 2019, 16, 2335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Schiro, G.; Colangeli, P.; Müller, M.E.H. A metabarcoding analysis of the mycobiome of wheat ears across a topographically heterogeneous field. Front Microbiol. 2019, 10, 2095. [Google Scholar] [CrossRef] [PubMed]
  121. Knorr, K.; Jørgensen, L.N.; Nicolaisen, M. Fungicides have complex effects on the wheat phyllosphere mycobiome. PLoS ONE 2019, 14, e0213176. [Google Scholar] [CrossRef] [Green Version]
  122. Kavamura, V.N.; Mendes, R.; Bargaz, A.; Mauchline, T.H. Defining the wheat microbiome: Towards microbiome-facilitated crop production. Comput. Struct. Biotechnol. 2021, 19, 1200. [Google Scholar] [CrossRef] [PubMed]
  123. Simonin, M.; Dasilva, C.; Terzi, V.; Ngonkeu, E.L.M.; Diouf, D.; Kane, A.; Béna, G.; Moulin, L. Influence of plant genotype and soil on the wheat rhizosphere microbiome: Evidences for a core microbiome across eight African and European soils. FEMS Microbiol. Ecol. 2020, 96, fiaa067. [Google Scholar] [CrossRef]
  124. Schlatter, D.C.; Yin, C.; Hulbert, S.; Paulitz, T.C. Core rhizosphere microbiomes of dry land wheat are influenced by location and land use history. Appl. Environ. Microbiol. 2019, 86, e02135-19. [Google Scholar]
  125. Rossmann, M.; Pérez-Jaramillo, J.E.; Kavamura, V.N.; Chiaramonte, J.B.; Dumack, K.; Fiore-Donno, A.M.; Mendes, L.W.; Ferreira, M.M.C.; Bonkowski, M.; Raaijmakers, J.M.; et al. Multitrophic interactions in the rhizosphere microbiome of wheat: From bacteria and fungi to protists. FEMS Microbiol. Ecol. 2020, 96, fiaa032. [Google Scholar] [CrossRef]
  126. Toju, H.; Peay, K.G.; Yamamichi, M.; Narisawa, K.; Hiruma, K.; Naito, K.; Fukuda, S.; Ushio, M.; Nakaoka, S.; Onoda, Y.; et al. Core microbiomes for sustainable agroecosystems. Nat. Plants 2018, 4, 247. [Google Scholar] [CrossRef]
  127. Fiorilli, V.; Belmondo, S.; Khouja, H.R.; Abbà, S.; Faccio, A.; Daghino, S.; Lanfranco, L. RiPEIP1, a gene from the arbuscular mycorrhizal fungus Rhizophagus irregularis, is preferentially expressed in planta and may be involved in root colonization. Mycorrhiza 2016, 26, 609–621. [Google Scholar] [CrossRef]
  128. Hartmann, M.; Frey, B.; Mayer, J.; Mäder, P.; Widmer, F. Distinct soil microbial diversity under longterm organic and conventional farming. ISME J. 2014, 9, 1177–1194. [Google Scholar] [CrossRef] [Green Version]
  129. Martienssen, R.A.; Colot, V. DNA methylation and epigenetic inheritance in plants and filamentous fungi. Science 2001, 293, 1070–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Feng, S.; Jacobsen, S.E.; Reik, W. Epigenetic reprogramming in plant and animal development. Science 2010, 330, 622–627. [Google Scholar] [CrossRef] [Green Version]
  131. Asgari, S. Epigenetic modifications underlying symbiont–host interactions. Adv. Genet. 2014, 86, 253–276. [Google Scholar] [PubMed]
  132. Varga, S.; Soulsbury, C.D. Arbuscular mycorrhizal fungi change host plant DNA methylation systemically. Plant Biol. 2019, 21, 278–283. [Google Scholar] [CrossRef]
  133. Ichida, H.; Matsuyama, T.; Abe, T.; Koba, T. DNA adenine methylation changes dramatically during establishment of symbiosis. FEBS J. 2007, 274, 951–962. [Google Scholar] [CrossRef] [Green Version]
  134. Kinoshita, T.; Motoaki, S. Epigenetic memory for stress response and adaptation in plants. Plant Cell Physiol. 2014, 55, 1859–1863. [Google Scholar] [CrossRef]
  135. Thiebaut, F.; Hemerly, A.S.; Ferreira, P.C.G. A role for epigenetic regulation in the adaptation and stress responses of non-model plants. Front. Plant Sci. 2019, 10, 246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Appels, R.; Eversole, K.; Stein, N.; Feuillet, C.; Keller, B.; Rogers, J.; Khurana, J.P. Shifting the limits in wheat research and breeding using a fully annotated reference genome. Science 2018, 361, eaar7191. [Google Scholar]
  137. Ramírez-González, R.H.; Borrill, P.; Lang, D.; Harrington, S.A.; Brinton, J.; Venturini, L. International Wheat Genome Sequencing Consortium. The transcriptional landscape of polyploid wheat. Science 2018, 361, eaar7191. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Saripalli, G.; Sharma, C.; Gautam, T.; Singh, K.; Jain, N.; Prasad, P.; Roy, J.K.; Sharma, J.B.; Sharma, P.K.; Prabhu, K.V.; et al. Complex relationship between DNA methylation and gene expression due to Lr28 in wheat-leaf rust pathosystem. Mol. Biol. Rep. 2020, 47, 1339–1360. [Google Scholar] [CrossRef]
  139. Geng, S.; Kong, X.; Song, G.; Jia, M.; Guan, J.; Wang, F.; Qin, Z.; Wu, L.; Lan, X.; Li, A.; et al. DNA methylation dynamics during the interaction of wheat progenitor Aegilops tauschii with the obligate biotrophic fungus Blumeria graminis f. sp. tritici. N. Phytol. 2019, 221, 1023–1035. [Google Scholar] [CrossRef] [Green Version]
  140. Sun, X.; Lin, L.; Sui, N. Regulation mechanism of microRNA in plant response to abiotic stress and breeding. Mol. Biol. Rep. 2019, 46, 1447–1457. [Google Scholar] [CrossRef] [PubMed]
  141. Xin, M.; Wang, Y.; Yao, Y.; Xie, C.; Peng, H.; Ni, Z.; Sun, Q. Diverse set of microRNAs are responsive to powdery mildew infection and heat stress in wheat (Triticum aestivum L.). BMC Plant Biol. 2010, 10, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Kamthan, A.; Chaudhuri, A.; Kamthan, M.; Datta, A. Small RNAs in plants: Recent development and application for crop improvement. Front Plant Sci. 2015, 6, 208. [Google Scholar] [CrossRef] [Green Version]
  143. Gupta, O.P.; Permar, V.; Koundal, V.; Singh, U.D.; Praveen, S. MicroRNA regulated defense responses in Triticum aestivum L. during Puccinia graminis f. sp. tritici infection. Mol. Biol. Rep. 2012, 39, 817–824. [Google Scholar] [CrossRef]
  144. Kumar, D.; Singh, D.; Kanodia, P.; Prabhu, K.V.; Kumar, M.; Mukhopadhyay, K. Discovery of novel leaf rust responsive microRNAs in wheat and prediction of their target genes. J. Nucleic Acids. 2014, ID 570176. [Google Scholar] [CrossRef] [Green Version]
  145. Cai, Q.; He, B.; Jin, H. A safe ride in extracellular vesicles—Small RNA trafficking between plant hosts and pathogens. Curr. Opin. Plant Biol. 2019, 52, 140–148. [Google Scholar] [CrossRef]
  146. Lauressergues, D.; Delaux, P.M.; Formey, D.; Lelandais-Brière, C.; Fort, S.; Cottaz, S.; Bécard, G.; Niebel, A.; Roux, C.; Combier, J.P. The microRNA miR171h modulates arbuscular mycorrhizal colonization of Medicago truncatula by targeting NSP2. Plant J. 2012, 72, 512–522. [Google Scholar] [CrossRef] [PubMed]
  147. Bazin, J.; Khan, G.A.; Combier, J.P.; Bustos-Sanmamed, P.; Debernardi, J.M.; Rodriguez, R.; Sorin, C.; Palatnik, J.; Hartmann, C.; Crespi, M.; et al. miR396 affects mycorrhization and root meristem activity in the legume Medicago truncatula. Plant J. 2013, 74, 920–934. [Google Scholar] [CrossRef]
  148. Couzigou, J.M.; Lauressergues, D.; André, O.; Gutjahr, C.; Guillotin, B.; Bécard, G.; Combier, J.P. Positive gene regulation by a natural protective miRNA enables arbuscular mycorrhizal symbiosis. Cell Host Microbe 2017, 21, 106–112. [Google Scholar] [CrossRef] [Green Version]
  149. Etemadi, M.; Gutjahr, C.; Couzigou, J.-M.; Zouine, M.; Lauressergues, D.; Timmers, A.; Audran, C.; Bouzayen, M.; Bécard, G.; Combier, J.-P. Auxin perception is required for arbuscule development in arbuscular mycorrhizal symbiosis. Plant Physiol. 2014, 166, 281–292. [Google Scholar] [CrossRef] [Green Version]
  150. Jiao, J.; Peng, D. Wheat microRNA1023 suppresses invasion of Fusarium graminearum via targeting and silencing FGSG_03101. J. Plant Interact. 2018, 13, 514–521. [Google Scholar] [CrossRef] [Green Version]
  151. Zhang, H.; Chen, X.; Wang, C.; Xu, Z.; Wang, Y.; Liu, X.; Kang, Z.; Ji, W. Long non-coding genes implicated in response to stripe rust pathogen stress in wheat (Triticum aestivum L.). Mol. Biol. Rep. 2013, 40, 6245–6253. [Google Scholar] [CrossRef]
  152. Zhang, H.; Hu, W.; Hao, J.; Lv, S.; Wang, C.; Tong, W.; Wang, Y.; Wang, Y.; Liu, X.; Ji, W. Genome-wide identification and functional prediction of novel and fungi-responsive lincRNAs in Triticum aestivum. BMC Genom. 2016, 17, 238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Jain, N.; Sinha, N.; Krishna, H.; Singh, P.K.; Gautam, T.; Prasad, P.; Gupta, P.K. A study of miRNAs and lncRNAs during Lr28-mediated resistance against leaf rust in wheat (Triticum aestivum L.). Physiol. Mol. Plant Pathol. 2020, 112, 101552. [Google Scholar] [CrossRef]
  154. Han, G.; Cheng, C.; Zheng, Y.; Wang, X.; Xu, Y.; Wang, W.; Zhu, S.; Cheng, B. Identification of long non-coding RNAs and the regulatory network responsive to arbuscular mycorrhizal fungi colonization in maize roots. Int. J. Mol. Sci. 2019, 20, 4491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. World Population Prospects 2019, Volume I: Comprehensive Tables; United Nations Department of Economic and Social Affairs/Population Division, United Nations New York: New York, NY, USA, 2019; ISBN 978-92-1-148327-7. eISBN: 978-92-1-004642-8.
  156. Larran, S.; Simon, M.R.; Moreno, M.V.; Siurana, M.S.; Perello, A. Endophytes from wheat as biocontrol agents against tan spot disease. Biol. Control. 2016, 92, 17–23. [Google Scholar] [CrossRef]
  157. Fernández-Pastor, I.; González-Menéndez, V.; Annang, F.; Toro, C.; Mackenzie, T.A.; Bosch-Navarrete, C.; Reyes, F. Pipecolisporin, a novel cyclic peptide with antimalarial and antitrypanosome activities from a wheat endophytic nigrospora oryzae. Pharmaceuticals 2021, 14, 268. [Google Scholar] [CrossRef]
  158. Reddy, G.V.P.; Tangtrakulwanich, K.; Wu, S.; Miller, J.H.; Ophus, V.L.; Prewett, J.; Jaronski, S.T. Evaluation of the effectiveness of entomopathogens for the management of wireworms (Coleoptera. Elateridae) on spring wheat. J. Invertebr. Pathol. 2014, 120, 43–49. [Google Scholar] [CrossRef] [Green Version]
  159. Keyser, C.A.; Jensen, B.; Meyling, N.V. Dual effects of Metarhizium spp. and Clonostachys rosea against an insect and a seed-borne pathogen in wheat. Pest Manag. Sci. 2015, 72, 517–526. [Google Scholar] [CrossRef]
  160. Herreraa, S.D.; Grossia, C.; Zawoznika, M.; Groppaa, M.D. Wheat seeds harbour bacterial endophytes with potential as plant growth promoters and biocontrol agents of Fusarium graminearum. Microbiol. Res. 2016, 186–187, 37–43. [Google Scholar] [CrossRef]
  161. Colla, G.; Rouphael, Y.; Bonini, P.; Cardarelli, M. Coating seeds with endophytic fungi enhance growth, nutrient uptake, yield, and grain quality of winter wheat. Int. J. Plant Prot. 2015, 9, 171–189. [Google Scholar]
  162. Serfling, A.; Wirsel, S.G.R.; Lind, V.; Deising, H.B. Performance of the BioControl fungus piriformospora indica on wheat under greenhouse and field conditions. Phytopathology 2007, 97, 523–531. [Google Scholar] [CrossRef] [Green Version]
  163. Malik, A.; Butt, T.A.; Naqvi, S.T.A.; Yousaf, S.; Qureshi, M.K.; Zafar, M.I.; Farooq, G.; Nawaz, I.; Iqbal, M. Lead tolerant endophyte Trametes Hirsuta improved the growth and lead accumulation in the vegetative parts of Triticum aestivum L. Heliyon 2020, 6, e04188. [Google Scholar] [CrossRef] [PubMed]
  164. Ikram, M.; Ali, N.; Jan, G.; Jan, F.G.; Rahman, I.U.; Iqbal, A.; Hamayun, M. IAA producing fungal endophyte Penicillium roqueforti Thom., enhances stress tolerance and nutrients uptake in wheat plants grown on heavy metal contaminated soils. PLoS ONE 2018, 13, e0208150. [Google Scholar] [CrossRef] [PubMed]
  165. De Silva, N.I.; Brooks, S.; Lumyong, S.; Hyde, K.D. Use of endophytes as biocontrol agents. Fungal Biol. Rev. 2019, 33, 133–148. [Google Scholar] [CrossRef]
  166. Junaid, J.M.; Dar, N.A.; Bhat, T.A.; Bhat, A.H.; Bhat, M.A. Commercial biocontrol agents and their mechanism of action in the management of plant pathogens. Int. J. Mod. Plant Anim. Sci. 2013, 1, 39–57. [Google Scholar]
  167. Kiss, L. A review of fungal antagonists of powdery mildews and their potential as biocontrol agents. Pestic. Manag. Sci. 2003, 59, 475–483. [Google Scholar] [CrossRef]
  168. Butt, T.; Jackson, C.; Magan, N. Fungi as Biocontrol Agents: Progress, Problems and Potential; CABI: Wallingford, UK, 2001; p. 390. [Google Scholar]
  169. Degenkolb, T.; Fog-Nielsen, K.; Dieckmann, R.; Branco-Rocha, F.; Chaverri, P.; Samuels, G.J.; Thrane, U.; von Dohren, H.; Vilcinskas, A.; Bruckner, H. Peptaibol, secondary metabolite, and hydrophobin pattern of commercial biocontrol agents formulated with species of the Trichoderma harzianum complex. Chem. Biodivers. 2015, 12, 662–684. [Google Scholar] [CrossRef]
  170. Wang, M.; Jin, H. Spray-induced gene silencing: A powerful innovative strategy for crop protection. Trends Microbiol. 2017, 25, 4–6. [Google Scholar] [CrossRef] [Green Version]
  171. Simon, J.C.; Marchesi, J.R.; Mougel, C.; Selosse, M.A. Host-microbiota interactions: From holobiont theory to analysis. Microbiome 2019, 7, 5. [Google Scholar] [CrossRef] [PubMed]
  172. Liu, S.; Geng, S.; Li, A.; Mao, Y.; Mao, L. RNAi technology for plant protection and its application in wheat. aBIOTECH 2021. [Google Scholar] [CrossRef]
  173. Koch, A.; Biedenkopf, D.; Furch, A.; Weber, L.; Rossbach, O.; Abdellatef, E.; Linicus, L.; Johannsmeier, J.; Jelonek, L.; Goesmann, A.; et al. An RNAi-based control of Fusarium graminearum infections through spraying of long dsRNAs involves a plant passage and is controlled by the fungal silencing machinery. PLoS Pathog. 2016, 12, e1005901. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Network of identified endophytic fungal species in the organs of common wheat.
Figure 1. Network of identified endophytic fungal species in the organs of common wheat.
Pathogens 10 01288 g001
Figure 2. Network of identified endophytic fungal species in the organs of Triticum durum.
Figure 2. Network of identified endophytic fungal species in the organs of Triticum durum.
Pathogens 10 01288 g002
Table 1. Characteristic of fungal endophytes identified in wheat (Triticaceae).
Table 1. Characteristic of fungal endophytes identified in wheat (Triticaceae).
Species 1Tissue TypeRole 2 LocalizationWheatReferences
Alternaria alternataroots, stems, leavessaprophyte/
pathogen
South AfricaTriticum aestivum[46,47,90,91]
leavesArgentina
leaves, stems, glumes, grains
Alternaria infectoriarachis, leaves, glumes, anthers, stems, grainspathogenFrance[38]
leaves, glumes, grainsArgentina[53]
Alternaria triticimaculansleaves, glumes,
stems, grains
pathogenFrance[38]
Arthopyrenia salicisrootsunrecognizedPoland[49]
Athelia bombacinaglumesunrecognizedFrance[38]
Aureobasidium proteaeaerial organsunrecognized[92]
rachis, anthers, stems, grains[38]
Bipolaris cynodontisleavesunrecognizedArgentina[47]
Bipolaris sorokinianaroots, stems, leaves, crownspathogenCanadaTriticum durum[47,53,93,94]
leavesArgentinaTriticum aestivum
stems, grains
Biscogniauxia nummularialeavesunrecognizedFrance[38]
Botrytis cinerealeaves, glumes, antherspathogen[38]
Candida albicansgrainsunrecognizedArgentina[53]
Chaetomium globosumleavesunrecognized
mycoparasites
France[38,47,53,92,95,96]
leavesArgentina
aerial organsFrance
leaves, grainsArgentina
Cladorrhinum australerootsunrecognizedPoland[49]
Cladosporium alliigrains, rachis, roots, leaves, anthersunrecognizedFrance[38]
Cladosporium cladosporoidesroots, stems, leaves, awns, crownsunrecognized
mycoparasite
CanadaTriticum durum[93]
Cladosporium halotoleransgrainsunrecognizedFranceTriticum aestivum[38,92]
grains
Cladosporium herbarumleavessaprophyte/
pathogen
Argentina[47]
leaves, stems glumes, grains[53]
Cladosporium iridisglumesunrecognizedFrance[38]
Cladosporium minouraeroots, stems, leaves, awns, crownsunrecognizedCanadaTriticum durum[93]
Clonostachys rosearootsunrecognized
mycoparasite
FranceTriticum aestivum[38,92,97]
Cochliobolus sativus
(Bipolaris sorokiniana)
pathogenCanadaTriticum durum[93]
Cochliobolus spicifier
(Curvularia spicifera)
grainspathogenArgentinaTriticum aestivum[53,90]
Coriolopsis gallicaglumes, stemsunrecognizedFrance[38]
Curvularia lunataleaves, glumespathogenArgentina[53]
Cytospora chrysospermastemsunrecognizedFrance[38]
Diaporthe eres
(Phomopsis velata)
grainsunrecognized
Dichotomomyces cejpii
(Aspergillus cejpii)
rootsunrecognized
Didymella exitialis
(Neoascochyta exitialis)
leaves, glumes, antherspathogen
Doratomyces microsporus
(Cephalotrichum microsporum)
rootsunrecognized
Drechslera poae
(Pyrenophora poae)
grainspathogen
Epicoccum nigrumroots, stems, leavessaprophyte/
pathogen
South Africa[38,46,47,53]
leaves, anthers, grainspathogenFrance
leavessaprophyte/
pathogen
Argentina
Argentina
leaves, stems, glumes, grains
Eutypa maurarachisunrecognizedFrance[38]
Funalia trogii
(Trametes trogii)
stemsunrecognized
Fusarium tricinctumroots, stems,
awns, crowns
pathogenCanadaTriticum durum[93]
Fusarium acuminatumrootspathogen
Fusarium avenaceumroots, stems, leaves, awns, crownspathogen[46,49,93]
roots, stems, leavespathogenSouth AfricaTriticum aestivum
rootspathogenPolandTriticum aestivum
Fusarium culmorumpathogenCanadaTriticum durum[94]
Fusarium graminearumstemspathogenFranceTriticum aestivum[38,49,53]
leaves, stemsArgentina
rootsPolandTriticum aestivum spp. spelta
Fusarium oxysporumleaves, stemspathogenArgentinaTriticum aestivum[49,53,98]
rootsPoland
Fusarium redolenspathogenFrance[38,49]
Poland
Fusarium reticulatumroots, stems, leaves, awns, crownspathogenCanadaTriticum durum[93]
Fusarium torulosumpathogen
Fusarium tricinctumleavespathogenFranceTriticum aestivum[38,49]
rootsPoland
Gaeumannomyces graminispathogenCanadaTriticum durum[93]
FranceTriticum aestivum[38]
Ganoderma carnosumunrecognized
Geomyces pannorum
(Pseudogymnoascus pannorum)
roots, stems, leaves, awns, crownsunrecognizedCanadaTriticum durum[93]
Gnomoniopsis idaeicolaglumesunrecognizedFranceTriticum aestivum[38]
Hyphodermella rosaeroots, stemsunrecognized
Ilyonectria macrodidyma
(Dactylonectria macrodidyma)
rootsunrecognized
Magnaporthiopsis panicorumrootsunrecognizedPolandTriticum aestivum spp. spelta[49]
Microdochium bolleyiroots,
stems, leaves
unrecognized
pathogen
South AfricaTriticum aestivum[38,46,49,92,99]
rootsFrance
Poland
Microdochium nivaleroots, leaves, glumes, stems, anthers, grainspathogen
mycoparasite
France[38,100]
Mortierella hyalinaawnsunrecognizedCanadaTriticum durum[93]
Mortierella alpinarootsunrecognizedFranceTriticum aestivum[38]
Mycosphaerella graminicola
(Zymoseptoria tritici)
leavespathogen
Nigrospora sphaerica
(Nigrospora oryzae)
unrecognizedSouth Africa[46]
Oxyporus latemarginatusstemsFrance[38,101]
Parastagonospora avenaeleavesunrecognized/
pathogen
Penicillium aurantiogriseumroots, stems, leaves, awns, crownsCanadaTriticum durum[93]
Penicillium crustosum
(Penicillium solitum)
rootsunrecognizedPolandTriticum aestivum spp. vulgare[49]
Penicillium griseofulvumroots stems, leaves, awns, crownsCanadaTriticum durum[93]
Peniophora cinerealeaves, stemsFranceTriticum aestivum[38]
Periconia macrospinosarootsCanadaTriticum durum[38,49,93]
roots, leavesFranceTriticum aestivum
rootsPoland
Phlebia subserialisleavesFrance[38]
Phoma caloplacae
(Diederichomyces caloplacae)
stems
Phoma glomerata
(Didymella glomerata)
roots, stems, leavessaprophyte/
pathogen
South Africa[46,92]
aerial organsFrance
Pleospora herbarum
(Stemphylium vesicarium)
roots, stems, leavessaprophyte/
pathogen
South Africa[46,47,53]
leavesArgentina
Podospora fimbriata
(Schizothecium fimbriatum)
rootsunrecognizedFrance[38]
Podospora glutinans
(Schizothecium glutinans)
roots, leaves
Polyporus lepideusglumes, stems
Pyrenophora tritici-repentisleaves, awnspathogenCanadaTriticum durum[93]
leavesFranceTriticum aestivum[38]
Rhizoctonia solanistems, roots, glumes
rootsPoland[49]
Rhodosporidium kratochvilovae
(Rhodotorula kratochvilovae)
grainsunrecognizedFrance[38]
Rhodotorula rubra
(Rhodotorula mucilaginosa)
leavesunrecognized/
mycoparasites
Argentina[47,53,102]
leaves, stems, glumes
Sarocladium kiliensestemsunrecognizedFrance[38,92]
aerial organs
Sclerotinia sclerotiorumroots, stemspathogen[38]
Septoria tritici
(Zymoseptoria tritici)
leavesunrecognized/
pathogen
Argentina[53,103]
Setophoma terrestrisrootsunrecognizedPoland[49]
Stagonospora nodorum
(Parastagonospora nodorum)
stems, leavespathogenSouth Africa[46]
Stemphylium botryosumleaves, grainsunrecognizedArgentina[53]
Stereum hirsutumroots, glumesunrecognizedFrance[38]
Talaromyces flavusrootsunrecognized
Thielavia hyalocarpa
(Cladorrhinum hyalocarpum)
awnsunrecognizedCanadaTriticum durum[93]
Trametes gibbosarootsunrecognizedFranceTriticum aestivum[38]
Trametes hirsutagrainsunrecognized
Trametes versicolorglumesunrecognized
Trichoderma harzianumrootsmycoparasitesCanadaTriticum durum[93,104]
Trichoderma hamatumleaves, stems, glumesunrecognized/
mycoparasite
ArgentinaTriticum aestivum[53,104]
Trichoderma longibrachiatumrootsunrecognized/
mycoparasite
Poland[49,104]
Truncatella angustataroots, stems, leavesunrecognizedSouth Africa[46]
Xylaria longipesleavesunrecognizedFrance[38]
Waitea circinatarootsunrecognizedPoland[49]
Zopfiella piliferarootsunrecognized
1—name of the species indicated in the references and valid (in brackets) according to Index Fungorum; 2—the role of the species indicated only in relation to wheat or its pathogens (for antagonistic fungi, mycoparasites).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Błaszczyk, L.; Salamon, S.; Mikołajczak, K. Fungi Inhabiting the Wheat Endosphere. Pathogens 2021, 10, 1288. https://doi.org/10.3390/pathogens10101288

AMA Style

Błaszczyk L, Salamon S, Mikołajczak K. Fungi Inhabiting the Wheat Endosphere. Pathogens. 2021; 10(10):1288. https://doi.org/10.3390/pathogens10101288

Chicago/Turabian Style

Błaszczyk, Lidia, Sylwia Salamon, and Katarzyna Mikołajczak. 2021. "Fungi Inhabiting the Wheat Endosphere" Pathogens 10, no. 10: 1288. https://doi.org/10.3390/pathogens10101288

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