Skip to main content
Advertisement
Browse Subject Areas
?

Click through the PLOS taxonomy to find articles in your field.

For more information about PLOS Subject Areas, click here.

  • Loading metrics

Colonization dynamic and distribution of the endophytic fungus Microdochium bolleyi in plants measured by qPCR

  • Pavel Matušinsky ,

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Resources, Software, Supervision, Validation, Visualization, Writing – original draft, Writing – review & editing

    pavel.matusinsky@upol.cz

    Affiliations Department of Botany, Faculty of Science, Palacký University in Olomouc, Olomouc, Czech Republic, Department of Plant Pathology, Agrotest Fyto, Ltd, Kroměříž, Czech Republic

  • Vendula Florová,

    Roles Formal analysis

    Affiliation Department of Plant Pathology, Agrotest Fyto, Ltd, Kroměříž, Czech Republic

  • Božena Sedláková,

    Roles Validation

    Affiliation Department of Botany, Faculty of Science, Palacký University in Olomouc, Olomouc, Czech Republic

  • Patrik Mlčoch,

    Roles Formal analysis, Methodology, Software, Validation, Visualization

    Affiliation Department of Botany, Faculty of Science, Palacký University in Olomouc, Olomouc, Czech Republic

  • Dominik Bleša

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Software, Visualization, Writing – original draft, Writing – review & editing

    Affiliations Department of Plant Pathology, Agrotest Fyto, Ltd, Kroměříž, Czech Republic, Department of Experimental Biology, Faculty of Science, Masaryk University, Brno, Czech Republic

Abstract

Microdochium bolleyi is a fungal endophyte of cereals and grasses proposed as an ideal model organism for studying plant-endophyte interactions. A qPCR-based diagnostic assay was developed to detect M. bolleyi in wheat and Brachypodium distachyon tissues using the species-specific primers MbqITS derived from the ITS of the ribosomal gene. Specificity was tested against 20 fungal organisms associated with barley and wheat. Colonization dynamics, endophyte distribution in the plant, and potential of the seed transmission were analyzed in the wheat and model plant B. distachyon. The colonization of plants by endophyte starts from the germinating seed, where the seed coats are first strongly colonized, then the endophyte spreads to the adjacent parts, crown, roots near the crown, and basal parts of the stem. While in the lower distal parts of roots, the concentration of M. bolleyi DNA did not change significantly in successive samplings (30, 60, 90, 120, and 150 days after inoculation), there was a significant increase over time in the roots 1 cm under crown, crowns and stem bases. The endophyte reaches the higher parts of the base (2–4 cm above the crown) 90 days after sowing in wheat and 150 days in B. distachyon. The endophyte does not reach both host species’ leaves, peduncles, and ears. Regarding the potential for seed transmission, endophyte was not detected in harvested grains of plants with heavily colonized roots. Plants grown from seeds derived from parental plants heavily colonized by endophyte did not exhibit any presence of the endophyte, so transmission by seeds was not confirmed. The course of colonization dynamics and distribution in the plant was similar for both hosts tested, with two differences: the base of the wheat stem was colonized earlier, but B. distachyon was occupied more intensively and abundantly than wheat. Thus, the designed species-specific primers could detect and quantify the endophyte in planta.

Introduction

Endophyte communities are found in almost all plants [1, 2]. These communities usually consist mainly of bacteria or fungi [3] that reside inside plant organs and do not damage the host tissue [37]. Some fungal endophytes can spend part of their life cycle as latent pathogens in plant tissues [810]. Fungal endophytes are vital in enhancing plant resilience against various stresses, such as abiotic factors like drought and biotic factors like pathogens and pests [1113].

The presence of endophytes in plant tissues is typically determined using microscopic and media-cultivation methods [1416]. However, DNA-based methods have become increasingly implemented [17]. PCR has been employed to detect endophytes in various grass tissues, including Epichloë, Neotyphodium, and Acremonium [1821]. Microdochium bolleyi and Microdochium phragmitis were detected by nested PCR [22]. In addition to PCR, alternative methods such as PCR-DGEE (polymerase chain reaction and denaturing gradient gel electrophoresis) [23], Droplet Digital PCR Technology [24], microarray [25], sequencing [26], and immunological tests [27] can be employed to detect endophytes and other fungi residing in plant tissues. Furthermore, modern "omics" techniques such as comparative genomics, metagenomics, and metatranscriptomics have the potential to detect endophytes and unravel the relationships between plants and endophytes [28].

Microdochium bolleyi (Mb) is an endophytic fungus that primarily colonizes roots of cereals and grasses, including wheat [29], barley [30, 31], Brachypodium distachyon (Bd) [32], and others [3335]. Mb has been observed to suppress various plant pathogens affecting cereals [29, 31, 3639]. The formation of typical and under light microscopy visible chlamydospore clusters within colonized plant cells makes this endophytic fungus an excellent model organism to study plant endophyte interactions [32].

The first objective of this study was to develop species-specific primers for qPCR detection and quantification of endophyte M. bolleyi in plant tissues and thorough testing of their specificity and capability of detecting the endophyte in plants. The second aim of the study was to assess the colonization dynamic over time, the location of the endophyte in plants, and the seed transmission potential of the endophytic fungus M. bolleyi using wheat and the model plant species B. distachyon.

Materials and methods

Biological materials

All plants were cultivated in a cooled greenhouse (20/18°C, day/night) within pots 8 cm in diameter filled with a 50:50 planting substrate–sand (vol:vol) mixture, the substrate FLORCOM SV (BB Com, Letohrad, Czech Republic). For the experiment were used not treated seeds of the wheat cultivar Bohemia and of B. distachyon line Bd21 obtained from the Joint Genome Institute (https://jgi.doe.gov). Ten seeds were placed into each pot. After germination, the number of plants in every pot was reduced to six.

Six Mb isolates were used (UPOC-FUN-253–258) from the Collection of Phytopathogenic Microorganisms UPOC (Czech Republic). Inoculum of Mb isolates was obtained by cultivation on double sterilized millet seed [32], which was previously tested for the absence of Mb by qPCR. The inoculation of the plants was also carried out according to the method of Matušinsky et al. (2022) by adding inoculum to the seeds during the sowing [32]. Only sterile millet without endophyte was added to the control treatments.

Isolates of other fungi used for testing primers were obtained from four different collections of microorganisms (Table 1).

thumbnail
Table 1. Fungal species used in the analysis with the MbqITS primers designed in this study.

A positive reaction result is indicated by a Cq lower than 30. The Cq levels in the table are the average of three replicates.

https://doi.org/10.1371/journal.pone.0297633.t001

Sampling and assessment

To evaluate endophyte colonization by qPCR, plants were collected 30, 60, 90, 120, and 150 days after sowing/inoculation (dai), surface sterilized for 3 min in 1% NaOCl, and thoroughly rinsed in sterilized distilled water to remove all superficial hyphae and then dried. Plants were cut into individual parts: the lower part of the roots (distal part of the roots), roots 1 cm (part of the roots 1 cm below the crown), crown, base 1 cm (part of the stem base 1 cm above the crown), base 2–4 cm (part of the stem base 2–4 cm above the crown), peduncles, and ears (S1 Fig). To evaluate endophyte colonization by light microscopy, plants were collected in the same terms as above (30, 60, 90, 120, and 150 dai), and then fixed in 70% ethanol. Plants fixed in 70% ethanol were cleared in 2.5% KOH for three days, acidified in 1% HCl and stored in lactoglycerol [14]. Colonization of plant tissues by the endophytic fungus was assessed by microscopic examination (200× magnifications). Results were evaluated as positive when the visible presence of Mb chlamydospores was observed (S2 Fig). Estimation of root colonization level was made by an adapted method according to Trouvelot et al. (1986) [40].

Seeds of massively colonized plants were harvested and subsequently tested for Mb DNA by qPCR and sown to analyze the potential for endophyte transfer by seed. Plants grown from these seeds were tested for Mb chlamydospores by light microscopy and for Mb DNA content by qPCR.

DNA isolation, primers design, and qPCR

Fungal mycelia (approximately 50–100 mg of biomass) of tested fungi (Table 1) were harvested from Petri dishes, ground to a fine powder in a cooled mortar using liquid nitrogen, then homogenized. Total genomic DNA was extracted using the DNeasy Plant Mini Kit (Qiagen, Germany). Plant material (approximately 50 mg of dried biomass per sample) was also ground to powder and extracted as described above. DNA concentration was measured using Qubit fluorometric quantification (ThermoFisher Scientific, Waltham, MA, USA), and DNA was diluted to a concentration of 5 ng μL−1.

Based upon sequences ITS (KP859018), LSU (large subunit of the ribosomal gene; KP858954), and RPB2 gene (RNA 40 polymerase II second-largest subunit (KP859127)) [41] selected from the GenBank database, five primer pairs (MbqITSF/R, MbqLSU1F/R, MbqLSU2F/R, MbqPOL1F/R, and MbqPOL2F/R) were designed using Primer3Plus software [42]. The samples were analyzed using the CFX96TM Real-Time PCR Detection System (Bio-Rad, Hercules, CA, USA). The quantitative PCR mix consisted of 1× SYBR Green (Top-Bio, Vestec, Czech Republic), 0.2 μM forward and reverse primers (the best final primers MbqITSF/R were selected from a preliminary screen of the designed primers; (S1 Table)), 10 ng DNA (2 μL), and water to final volume 15 μL. The reference gene for wheat was TaPAL [43], and for Bd, it was BdFIM [44]. The control samples consisted of DNA of wheat or Bd plants colonized by the known level of root colonization by Mb obtained from the previous experiment; it will be further described in this article as a standard sample (sample of roots 1 cm under crown collected 90 days after sowing; wheat 20.6% and Bd with 23.4% of colonization measured by light microscopy method according to Trouvelot et al. (1986) [40]). The primers’ specificity and presence of primer dimers were verified by melting analysis. For the resulting MbqITSF/R primers, the reaction efficiency was analyzed by qPCR with DNA of Mb (isolate UPOC-FUN-253) and dilution series (1.0, 0.1, 0.01, 0.001, and 0.0001 ng).

Test of primers specificity and diagnostic potential in plant tissues

The primers were tested on the six Mb isolates as described above. The designed primer pairs were tested for their specificity towards the DNA of fungi associated with diseases of wheat and other cereals. Fungal cultures were obtained from four microorganism collections (Table 1) and cultured on PDA. Furthermore, the potential for detection was tested in wheat, and Bd inoculated with the Mb compared with non-inoculated plants (inoculation and sampling as described above). All reactions were repeated three times.

Statistical analysis

The data were analyzed using the 2–ΔΔCq method with CFX Maestro software (Bio-Rad), and three biological and three technical replicates were run. The reaction efficiency was calculated using CFX Maestro software (Bio-Rad) from the slope value of the standard curve according to the formula E = (-1 + (10–1/slope) × 100) [45]. Differences in the relative quantity of the average values of inoculated samples were evaluated and compared to the standard sample (described above) by Tukey’s pairwise comparison test (P < 0.05; CFX Maestro software, Bio-Rad).

Results

Demonstrated root colonization by Mb of wheat and Bd

Using microscopic and molecular methods, the presence of Mb was detected in both wheat and Bd samples. Using light microscopy, chlamydospores were observed in the seed coats, roots, crowns, and stem bases of wheat and Bd (S2, S3 Tables and Fig 1). No presence of chlamydospores was observed in plants without inoculation.

thumbnail
Fig 1.

Chlamydospores of Microdochium bolleyi in wheat leaf sheaths (A, B) and comparison of seed coats of non-inoculated (C) and inoculated (D) Brachypodium distachyon. Scale bar = 100 μm.

https://doi.org/10.1371/journal.pone.0297633.g001

A set of MbqITSF/R best met the specificity requirements of the five designed primer pairs. The primers showed positive results in all six tested isolates of Mb (S4 Table). In the specificity test, the best-selected primers amplified DNA only from Mb, not any other fungal pathogens screened (Table 1). The primer set was therefore used to examine Mb-inoculated and non-inoculated plant material (wheat and Bd). Inoculated plants with chlamydospores present in plant tissues of both wheat and Bd showed positive results after qPCR with primers MbqITSF/R and, on the contrary, non-inoculated plants without chlamydospores in plant tissues showed negative results with Cq values higher than 30. The analysis confirmed that the primers were specific and robust in detecting and quantifying Mb within plant tissues. No false positive results were seen after melting analysis. Other tested primer pairs, MbqLSU1F/R, MbqLSU2F/R, MbqPOL1F/R, and MbqPOL2F/R, failed in the specificity test and were excluded from other parts of this study. The efficiency of the reaction was determined using a dilution series of DNA from the Mb isolate (UPOC-FUN-253). Based on the standard curve obtained, the efficiency (E) of the reaction was found to be 98.5% (S5 Table). The method showed a reliable positive response at 0.001 ng of Mb DNA (S5 Table), and the melting curve gives a single peak, indicating no artifacts.

Dynamics of Mb colonization in wheat and Bd over time

The seed coats, i.e., the already dead tissue, are massively colonized (30 days after inoculation) (Fig 1C and 1D). Only then are the living parts of the plant inhabited, such as the crown and roots, especially the part of the roots near the crown (30 dai). Over time (90 dai), the endophyte spreads to the distal parts of the roots and in wheat to the higher parts of the stem, 2–4 cm, but in Bd, only up to 1 cm and not higher. The DNA of the endophyte increases over time in the aerial parts, in wheat earlier than in Bd (Fig 2). In the underground parts, especially in the lower distal part of the roots, however, the level of endophyte DNA is growing very slowly during sampling periods (30–150 dai) with no significant changes (Fig 2). In the wheat above-ground parts, the presence of chlamydospores is confirmed microscopically in the outer sheath of the oldest leaf sheaths enclosing the stem (60–90 dai). Gradually, chlamydospores can be found in the sheaths of the second-oldest and third-oldest leaves (120–150 dai). Usually, the sheaths of leaves already in the senescence stage are colonized in this way, so a tendency for Mb to inhabit sheaths already in the senescence stage can be observed. On microscopy, it was found that most of the chlamydospores in the aerial part of the plants are in the cells adjacent to the vascular bundles (Fig 1A and 1B). Still, the fungus does not penetrate the vascular bundle itself, and the tissues are free of visual symptoms, i.e., without browning or other signs of damage. At the last sampling date of 150 days after sowing, the plants were already completely dry and dead, and the increase in Mb DNA in crown and roots 1cm above crown in wheat and stem base 2–4 cm above crow in Bd is, therefore, more likely due to the saprotrophic effect of the fungus. No Mb DNA was detected in leaf blades, peduncles, ears, and harvested seeds, nor was the presence of chlamydospores seen in these parts of the plants (S2 and S3 Tables).

thumbnail
Fig 2.

Dynamics of Microdochium bolleyi colonization in (A) wheat and (B) Brachypodium distachyon over time analyzed by qPCR. Statistically significant differences are indicated by asterixis (post hoc Tukey’s test, P < 0.05).

https://doi.org/10.1371/journal.pone.0297633.g002

Distribution of Mb in wheat and Bd plants

The endophyte distribution in the plant varies over time, as indicated in the previous section of the results. Endophyte colonizes the roots, crowns, and stem bases of plants. The parts of the root adjacent to the crown are more intensively colonized than the distal parts of the roots, which are less colonized (S2 and S3 Tables). Crowns are inhabited initially, and the intensity is higher in the last phase of the plant’s life. The first tissues that are colonized after sowing are the seed coats (Fig 1C and 1D). Surprisingly, the plant’s above-ground parts, especially the lowest parts of the base, are inhabited significantly, and the intensity increases over time. The higher parts of the base, 2–4 cm, are also colonized in wheat. In contrast, in Bd, colonization in these higher regions of the base occurs only at the last sampling date, 150 days after sowing/inoculation. Neither leaf blades nor peduncles and ears are colonized even at the late stages of plant development (S2 and S3 Tables).

Transmission by seed was not confirmed

Seeds from ten plants heavily colonized by the endophyte Mb were harvested and sown. These seeds and the plants grown from these seeds were screened for the presence of Mb by both light microscopy and qPCR. Neither the seeds nor any parts of the plants sown from these seeds contained endophyte Mb (S6 Table). Thus, it can be construed that Mb is not transmitted vertically, i.e., by seed.

Discussion

The fungal endophyte of grasses and cereals Mb is a suitable model organism for studying endophyte interactions with plants. Mb can colonize several cereal species, including wheat [29], barley [30], and the model organism Bd [32]. Molecular methods for detecting Mb have previously been described by nested PCR [22] and standard PCR [32]. In the current study, newly developed primers for qPCR were designed and tested to quantify the amount of Mb DNA in host tissues. The colonization dynamics over time, the distribution of endophytes in different parts of the plant, and the potential of seed transfer, i.e., vertical transfer, were tested in wheat and Bd. Although microscopic methods have been used to determine the presence of endophytes, in many cases, they are combined with molecular or other methods [31, 46, 47]. However, light microscopy could be time-consuming and inaccurate compared to molecular techniques. Identifying individual organisms using microscopy requires considerable experience, as species can only be distinguished based on morphological characters in this way. In addition, individual roots are usually not evenly colonized along their entire length, so only a particular part of the roots needs to be selected for microscopy. If the roots are too thick, focusing the slide in all aspects is challenging. Microscopic detection should be used as a companion tool with a more reliable and accurate molecular method but not as a stand-alone method for diagnosing endophytes.

Mb is a partially saprotrophic organism; in some cases, slightly pathogenic effects on wheat have been observed [30, 48]. The presence of Mb has been previously described in roots [30], crowns [49], and stems of cereals [50]. We used newly developed primers to screen the dynamics of Mb colonization of plants. First, massive colonization of the seed coats occurs, and only then are colonized roots near the crown and parts of the stem base within 1 cm of the crown. Over time Mb spreads to the distal parts of the roots and, in wheat, to the higher parts of the stem. In the lower parts of the roots, the level of endophyte DNA is more or less the same at all sampling periods without significant changes. Mb content in roots part 1 cm bellow crown increases in the last sampling term, e.g., 150 dai in wheat and 120 and 150 dai in Bd. In above-ground parts, the presence of chlamydospores is confirmed microscopically in the outer sheath of the oldest leaf enveloping the stem. No Mb DNA was detected in leaf blades, peduncles, ears, and seeds, nor was the presence of chlamydospores in these plant parts. Seeds from plants heavily colonized by Mb endophyte were harvested and sown, and plants from these seeds were examined for the presence of Mb by both microscopic and qPCR methods. No Mb endophyte was detected in any parts of the plants sown from these seeds, and therefore, it can be concluded that Mb is not transmitted vertically, i.e., by seeds.

Our study is not the first to address the quantification of endophytes in plant tissues using qPCR. For example, Chow et al. (2018) [51] used qPCR to investigate the effect of endophytes (Diaporthe phaseolorum, Trichoderma asperellum, and Penicillium citrinum) on the pathogen Ganoderma boninense that infects oil palm (Elaeis guineensis). The analysis showed that when plants were inoculated simultaneously with endophytes, and the pathogen G. boninense, colonization by the pathogen was suppressed by endophytic fungi, as reflected by the higher DNA content of endophytes compared to that of the pathogen. Latz et al. (2021) [52] investigated interactions between plant genotypes and the biotic and abiotic environment affecting the plant microbiome using the ITS1 metabarcoding method. Root microbial communities were less affected by abiotic factors than the phyllosphere microbiome. The diversity was highest in roots and lower in seeds, while leaves harbored the least diverse microbiome. The complete opposite was found in the work of Tao et al. (2008), who studied the distribution of fungal communities within leaf and root tissues of Bletilla ochracea (Orchidaceae) using random cloning, combined with DGGE (denaturing gradient gel electrophoresis) and phylogenetic analysis [53]. The main objective of their study was to compare the diversity of endophytes among roots and leaves of the same plants and to determine whether there is a consistency of communities within different organs of the same host. Fungal communities within leaf and root tissues differed significantly, and diversity within leaves was higher than within roots [53]. Bayman et al. (1997) used traditional isolation techniques and found that Xylaria spp. (Xylariales) and Rhizoctonia-like taxa (Basidiomycota) made up the majority of endophytes in epiphytic orchids, and fungal communities within leaves and roots were surprisingly similar [54].

These studies highlight the importance of conducting thorough investigations on endophytes and developing precise and reliable methods for detecting these organisms within the tissues of host plants. The use of molecular techniques in the determination of endophytes has wide applications. Cook et al. (2009) [55] used qPCR to determine the amount of an endophyte involved in producing a toxic alkaloid (swainsonine) in plants of the genus Oxytropis and Astragalus. The endophyte Undifilum oxytropis, involved in synthesizing the toxic alkaloid swainsonine, was quantified in plants of Oxytropis sericea and Astragalus mollissimus species using quantitative PCR. The amount of endophytes varies from one plant part to another and, in some cases, does not correspond to the concentration of swainsonine in the respective parts. The above works deal with microbial diversity inhabiting different parts of plants. Our work focused on one endophyte species and used qPCR to investigate the colonization dynamics and its distribution in the plant. The current study confirmed the endophyte presence in the roots and stem bases of wheat and Bd. Mb is found in roots, generally typical of the fourth class of endophytes to which this endophyte belongs [4]. Developing these newly designed primers has significantly broadened our understanding of the dynamics and distribution of endophytes within plants. Moreover, their application extends beyond our study, as they hold great potential for confirming successful colonization in commercial inoculations and facilitating further research in this field.

Fungal endophytes could be transmitted from plant to plant horizontally by soil or vertically by seed. A previous study by Sharon et al. (2023) [2] measured the correlation between fungal endophyte communities in seeds and stems of three cereal species under natural conditions and a controlled environment. Although most dominant fungal endophytic taxa could reach the seed internally, external seed infection (horizontal transfer) was found to be a primary source of specific taxa, including the most abundant species, Alternaria infectoria. Vertical transmission is not probably so frequent and is restricted to particular fungal taxa such as Epichloë [56]. Nevertheless, several studies indicate relatively more widespread vertical transfer than previously assumed [57, 58]. Mb belongs to the category of type 4 endophytes, and like other species in this group, it is recognized for its ability to be horizontally transmitted [4]. This has also been confirmed in our study. However, dead plants and debris remain on the ground and can be colonized by Mb. Murray and Gadd (1981) found that colonized seed remnants and roots remain in the ground after the crop is harvested and would provide a source of inoculum for the infection of barley seedlings in the following spring [30]. Regarding possible biological control, a potential approach could involve inoculation and colonization of seeds by Mb before sowing. By colonizing the seed coats, these cortical seed parts could serve as a source of inoculum and provide a favorable environment for seedlings’ inoculation. However, this hypothetical inoculation method has not yet been developed, and it would be helpful to test it in practice.

There are several reasons why the endophytic fungus Mb and plant Bd create an optimal model for studying plant-fungal interactions. The current study confirmed that the correspondence of Mb and Bd is excellent, and the fungus grows readily in Bd tissues. The inoculation method is simple, and the structures of Mb are readily visible in host tissues using light microscopy. In addition, we have developed and optimized species-specific primers for standard PCR [32] and here also for quantification of Mb by qPCR. The plant colonization dynamics by Mb are thoroughly studied in the current study. Previous studies have shown the beneficial effect of Mb on improving host plant immunity. Optimized methods and available primers from the last research [59, 60] allow the analysis of gene expression of specific marker genes in Bd. The advantage of Mb compared to strictly biotrophic symbiotic organisms is the possibility of cultivation on axenic media, and its massive sporulation ability makes its use suitable for large-scale production. Our results show that Bd and endophytic fungus Mb can be a model system for the plant-fungi interaction surveys. In general, a more profound knowledge of colonization dynamics and endophyte transmission could be used to improve agricultural management regarding plant growth promotion and biocontrol.

Conclusion

Mb is a suitable model for studying the interactions of endophytic fungi and plants. Primers for qPCR have been developed for the diagnosis and relative quantification of endophyte Mb in plant tissues and tested in wheat and Bd. The method is suitable for confirming the presence of endophytes in plants, studying colonization dynamics and distribution in plants, and other applications.

Supporting information

S2 Fig. Brachypodium distachyon root with chlamydospores of Microdochium bolleyi.

Scale bar = 100 μm.

https://doi.org/10.1371/journal.pone.0297633.s002

(TIF)

S1 Table. Primer pairs used in the study.

Names, sequences of forward and reverse primers, publication sources of primer pairs, and gene functions are listed.

https://doi.org/10.1371/journal.pone.0297633.s003

(DOCX)

S2 Table. Colonization of wheat tissues by light microscopy and qPCR.

https://doi.org/10.1371/journal.pone.0297633.s004

(DOCX)

S3 Table. Colonization of Brachypodium distachyon tissues by light microscopy and qPCR.

https://doi.org/10.1371/journal.pone.0297633.s005

(DOCX)

S4 Table. Resulting Cq values after qPCR with five primer pairs.

https://doi.org/10.1371/journal.pone.0297633.s006

(DOCX)

S5 Table. The efficiency (E) of the reaction was determined using a dilution series of DNA from the Microdochium bolleyi isolate (UPOC-FUN-253) with primers MbqITS.

https://doi.org/10.1371/journal.pone.0297633.s007

(DOCX)

Acknowledgments

We thank the Collection of Phytopathogenic Microorganisms UPOC for providing fungal isolates.

References

  1. 1. Chen J, Sharifi R, Khan MSS, Islam F, Bhat JA et al. Wheat Microbiome: Structure, Dynamics, and Role in Improving Performance Under Stress Environments. Front Microbiol. 2022; 12:821546. pmid:35095825.
  2. 2. Sharon O, Sun X, Ezrati S, Kagan-Trushina N, Sharon A. Transmission Mode and Assembly of Seed Fungal Endophyte Communities in Wheat and Wheat Wild Relatives. Phytobiomes J. 2023; 7:113–124. https://doi.org/10.1094/PBIOMES-11-22-0084-R.
  3. 3. Bacon CW, White J, editors. Microbial Endophytes. 1st ed. CRC Press; 2000. Available from: https://doi.org/10.1201/9781482277302. vol. 2000.
  4. 4. Rodriguez RJ, White JF Jr, Arnold AE, Redman RS. Fungal endophytes: diversity and functional roles. New Phytol. 2009; 182:314–330. pmid:19236579
  5. 5. Petrini O. Fungal Endophytes of Tree Leaves. in Microbial Ecology of Leaves (eds. Andrews JH, Hirano SS). Springer New York; 1991. p. 179–197.
  6. 6. Delaye L, García-Guzmán G, Heil M. Endophytes versus biotrophic and necrotrophic pathogens—are fungal lifestyles evolutionarily stable traits? Fungal Diversity. 2013; 60:125–135. https://doi.org/10.1007/s13225-013-0240-y.
  7. 7. Sánchez Márquez S, Bills GF, Domínguez Acuña L, Zabalgogeazcoa I. Endophytic mycobiota of leaves and roots of the grass Holcus lanatus. Fungal Diversity. 2010; 41:115–123. https://doi.org/10.1007/s13225-009-0015-7.
  8. 8. Photita W, Lumyong S, Lumyong P, McKenzie EHC, Hyde KD. Are some endophytes of Musa acuminata latent pathogens? Fungal Divers. 2004; 16: 131–140. http://hdl.handle.net/10722/223070.
  9. 9. Stergiopoulos I, Gordon TR. Cryptic fungal infections: the hidden agenda of plant pathogens. Front Plant Sci. 2014; 5:506. pmid:25309571.
  10. 10. Brader G, Compant S, Vescio K, Mitter B, Trognitz F, Ma L-J, et al. Ecology and Genomic Insights into Plant-Pathogenic and Plant-Nonpathogenic Endophytes. Annu. Rev. Phytopathol. 2017; 55:61–83. pmid:28489497.
  11. 11. Hardoim PR, van Overbeek LS, Berg G, Pirttilä AM, Compant S, Campisano A, et al. The hidden world within plants: ecological and evolutionary considerations for defining functioning of microbial endophytes. Microbiol Mol Biol Rev. 2015; 79:293–320. pmid:26136581.
  12. 12. Fadiji AE, Babalola OO. Elucidating Mechanisms of Endophytes Used in Plant Protection and Other Bioactivities With Multifunctional Prospects. Front Bioeng Biotechnol. 2020; 8:532550. pmid:32500068
  13. 13. Lata R, Chowdhury S, Gond SK, White JF Jr. Induction of abiotic stress tolerance in plants by endophytic microbes. Lett Appl Microbiol. 2018; 66:268–276. pmid:29359344.
  14. 14. Bleša D, Matušinský P, Sedmíková R, Baláž M. The potential of Rhizoctonia-like fungi for the biological protection of cereals against fungal pathogens. Plants. 2021; 10:349. https://doi.org/10.3390/plants10020349.
  15. 15. Ghimire SR, Charlton ND, Bell JD, Krishnamurthy YL, Craven KD. Biodiversity of fungal endophyte communities inhabiting switchgrass (Panicum virgatum L.) growing in the native tallgrass prairie of northern Oklahoma. Fungal Diversity. 2011; 47:19–27. https://doi.org/10.1007/s13225-010-0085-6.
  16. 16. Xia Y, Sahib MR, Amna A, Opiyo SO, Zhao Z, Gao YG. Culturable endophytic fungal communities associated with plants in organic and conventional farming systems and their effects on plant growth. Sci Rep. 2019; 9:1669. pmid:30737459
  17. 17. McKinnon AC. Plant tissue preparation for the detection of an endophytic fungus in planta. Methods Mol Biol. 2016; 1477:167–173. pmid:27565499
  18. 18. Groppe K, Boller T. PCR assay based on a microsatellite-containing locus for detection and quantification of Epichloë endophytes in grass tissue. Appl Environ Microbiol. 1997; 63:1543–50. pmid:9097449.
  19. 19. Doss RP, Clement SL, Kuy S-R, Welty RE. A PCR-Based Technique for Detection of Neotyphodium Endophytes in Diverse Accessions of Tall Fescue. Plant Dis. 1998; 82:738–740. pmid:30856941.
  20. 20. Zhou Y, Bradshaw RE, Johnson RD, Hume DE, Simpson WR, Schmid J. Detection and quantification of three distinct Neotyphodium lolii endophytes in Lolium perenne by real time PCR of secondary metabolite genes. Fungal Biol. 2014; 118:316–324. https://doi.org/10.1016/j.funbio.2014.01.003.
  21. 21. Kelemu S, Dongyi H, Guixiu H, Takayama Y. Detecting and differentiating Acremonium implicatum: developing a PCR-based method for an endophytic fungus associated with the genus Brachiaria. Mol. Plant Pathol. 2003; 4:115–118. https://doi.org/10.1046/j.1364-3703.2003.00157.x.
  22. 22. Ernst M, Neubert K, Mendgen KW, Wirsel SGR. Niche differentiation of two sympatric species of Microdochium colonizing the roots of common reed. BMC Microbiol. 2011; 11:242. https://doi.org/10.1186/1471-2180-11-242.
  23. 23. Ma WK, Siciliano SD, Germida JJ. A PCR-DGGE method for detecting arbuscular mycorrhizal fungi in cultivated soils. Soil Biol Biochem. 2005; 37:1589–1597. https://doi.org/10.1016/j.soilbio.2005.01.020.
  24. 24. del Pilar Martínez-Diz M, Andrés-Sodupe M, Berbegal M, Bujanda R, Díaz-Losada E, Gramaje D. Droplet Digital PCR Technology for Detection of Ilyonectria liriodendri from Grapevine Environmental Samples. Plant Dis. 2020; 104:1144–1150. https://doi.org/10.1094/PDIS-03-19-0529-RE.
  25. 25. Dinkins RD, Barnes A, Waters W. Microarray analysis of endophyte-infected and endophyte-free tall fescue. J Plant Physiol. 2010; 167:1197–1203. pmid:20430473
  26. 26. Asaf S, Khan AL, Khan MA, Al-Harrasi A, Lee IJ. Complete genome sequencing and analysis of endophytic Sphingomonas sp. LK11 and its potential in plant growth. 3 Biotech. 2018; 8:389. https://doi.org/10.1007/s13205-018-1403-z.
  27. 27. Koh S, Vicari M, Ball JP, Rakocevic T, Zaheer S, Hik DS, et al. Rapid Detection of Fungal Endophytes in Grasses for Large-Scale Studies. Functional Ecology. 2006; 20:736–742. https://doi.org/3806624.
  28. 28. Kaul S, Sharma T, Dhar MK. “Omics” Tools for Better Understanding the Plant–Endophyte Interactions. Frontiers in Plant Science. 2016; 7. pmid:27446181
  29. 29. Reinecke P, Fokkema NJ. An evaluation of methods of screening fungi from the Haulm base of cereals for antagonism to Pseudocercosporella herpotrichoides in wheat. Trans Br Mycol Soc. 1981; 77:343–350. https://doi.org/10.1016/s0007-1536(81)80036-8.
  30. 30. Murray DIL, Gadd GM. Preliminary studies on Microdochium bolleyi with special reference to colonization of barley. Trans Br Mycol Soc. 1981; 76:397–403. https://doi.org/10.1016/S0007-1536(81)80065-4.
  31. 31. Fatemi L. Shadmani S, Jamali A. Biocontrol activity of endophytic fungus of barley, Microdochium bolleyi, against Gaeumannomyces graminis var. tritici. Mycologia Iranica. 2018; 5:7–14. https://doi.org/10.22043/mi.2019.118205.
  32. 32. Matušinsky P, Sedláková B, Bleša D. Compatible interaction of Brachypodium distachyon and endophytic fungus Microdochium bolleyi. PLoS ONE. 2022; 17:e0265357. https://doi.org/10.1371/journal.pone.0265357.
  33. 33. Hong SK, Kim WG, Choi HW, Lee SY. Identification of Microdochium bolleyi Associated with Basal Rot of Creeping Bent Grass in Korea. Mycobiology. 2008; 36:77–80. pmid:23990737.
  34. 34. Hodges CF, Campbell DA. Infection of adventitious roots of Agrostis palustris by Idriella bolleyi. J. Phytopathol. 1996; 144:265–271. https://doi.org/10.1111/j.1439-0434.1996.tb01527.x.
  35. 35. Damm U, Brune A, Mendgen K. In vivo observation of conidial germination at the oxic–anoxic interface and infection of submerged reed roots by Microdochium bolleyi. FEMS Microbiol Ecol. 2003; 45:293–299. https://doi.org/10.1016/S0168-6496(03)00161-2.
  36. 36. Kirk JJ, Deacon JW. Control of the take–all fungus by Microdochium bolleyi, and interactions involving M. bolleyi, Phialophora graminicola and Periconia macrospinosa on cereal roots. Plant Soil. 1987; 98:231–237. https://doi.org/10.1007/BF02374826.
  37. 37. Sieber T, Riesen T, Müller E, Fried P. Endophytic fungi in four winter wheat cultivars (Triticum aestivum L.) differing in resistance against Stagonospora nodorum (Berk.) Cast. & Germ. = Septoria nodorum (Berk.) Berk. J. Phytopathol. 1988; 122:289–306. https://doi.org/10.1111/j.1439-0434.1988.tb01021.x.
  38. 38. Duczek LJ. Biological control of common root rot in barley by Idriella bolleyi. Can. J. Plant Pathol. 1997; 19:402–405. https://doi.org/10.1080/07060669709501067.
  39. 39. Comby M, Gacoin M, Robineau M, Rabenoelina F, Ptas S, Dupont J, et al. Screening of wheat endophytes as biological control agents against Fusarium head blight using two different in vitro tests. Microbiological Research. 2017; 202: 11–20. pmid:28647118.
  40. 40. Trouvelot A, Kough JL, Gianinazzi-Pearson V. Estimation of VA mycorrhizal infection levels. Research for method having a functional significance. In: Physiological and Genetical Aspects of Mycorrhizae (Gianinazzi-Pearson V and Gianinazzi S, Eds.). INRA Press, Paris. 1986; 217–221. ISBN 2-85340-774-8.
  41. 41. Hernández-Restrepo M, Groenewald JZ, Crous PW. Taxonomic and phylogenetic re-evaluation of Microdochium, Monographella, and Idriella. Persoonia. 2016; 36:57–82. https://doi.org/10.3767/003158516X688676.
  42. 42. Untergasser A, Nijveen H, Rao X, Bisseling T, Geurts R, Leunissen JA. Primer3Plus, an enhanced web interface to Primer3. Nucleic Acids Res. 2007; 35:71–74. pmid:17485472
  43. 43. Walsh K, Korimbocus J, Boonham N, Jennings P, Hims M. Using Real-time PCR to Discriminate and Quantify the Closely Related Wheat Pathogens Oculimacula yallundae and Oculimacula acuformis. J Phytopathol. 2005; 153:715–721. https://doi.org/10.1111/j.1439-0434.2005.01045.x.
  44. 44. Zhu H, Wen F, Li P, Liu X, Cao J, Jiang M, et al. Validation of a Reference Gene (BdFIM) for Quantifying Transgene Copy Numbers in Brachypodium distachyon by Real-Time PCR. Appl Biochem Biotechnol. 2014; 172:3163–3175. pmid:24497043.
  45. 45. Higuchi R, Fockler C, Dollinger G, Watson R. Kinetic PCR Analysis: Real-time Monitoring of DNA Amplification Reactions. Nat Biotechnol. 1993; 11:1026–1030. pmid:7764001
  46. 46. He C, Wang W, Hou J. Characterization of Dark Septate Endophytic Fungi and Improve the Performance of Liquorice Under Organic Residue Treatment. Front Microbiol. 2019; 10:1364. pmid:31275282
  47. 47. Tellenbach C, Grünig CR, Sieber TN. Suitability of quantitative real-time PCR to estimate the biomass of fungal root endophytes. Appl Environ Microbiol. 2010; 76:5764–5772. pmid:20601500.
  48. 48. Sprague R. Gloeosporium decay in Gramineae. Phytopathology. 1948; 38:131–136.
  49. 49. Hoes JA. Dynamics of the microflora of subterranean parts of winter wheat in the dryland area of Washington. Phytopathology. 1962; 52:736.
  50. 50. Reinecke P. Microdochium bolleyi at the stem base of cereals. Journal of Plant Diseases and Protection. 1978; 85:679–685.
  51. 51. Chow YY, Rahman S, Ting ASY. Interaction dynamics between endophytic biocontrol agents and pathogen in the host plant studied via quantitative real-time polymerase chain reaction (qPCR) approach. Biol Control. 2018; 125:44–49. https://doi.org/10.1016/j.biocontrol.2018.06.010.
  52. 52. Latz MAC, Kerrn MH, Sørensen H, Collinge DB, Jensen B, Brown JKM, et al. Succession of the fungal endophytic microbiome of wheat is dependent on tissue-specific interactions between host genotype and environment. Sci Total Environ. 2021; 759:143804. pmid:33340856
  53. 53. Tao G, Liu ZY, Hyde KD, Lui XZ, Yu ZN. Whole rDNA analysis reveals novel and endophytic fungi in Bletilla ochracea (Orchidaceae). Fungal Diversity. 2008; 33:101–122. http://cmuir.cmu.ac.th/jspui/handle/6653943832/60035.
  54. 54. Bayman P, Lebron LL, Tremblay RL, Lodge DJ. Variation in Endophytic Fungi from Roots and Leaves of Lepanthes (Orchidaceae). The New Phytologist. 1997; 135:143–149. https://doi.org/10.1046/j.1469-8137.1997.00618.x.
  55. 55. Cook D, Gardner DR, Ralphs MH, Pfister JA. Swainsoninine Concentrations and Endophyte Amounts of Undifilum oxytropis in Different Plant Parts of Oxytropis sericea. J Chem Ecol. 2009; 35:1272–1278. pmid:19904570.
  56. 56. Liu J, Nagabhyru P, Schardl CL. Epichloë festucae endophytic growth in florets, seeds, and seedlings of perennial ryegrass (Lolium perenne). Mycologia. 2017; 109:691–700. pmid:29293414.
  57. 57. Abdelfattah A, Tack AJM, Lobato C, Wassermann B, Berg G. From seed to seed: the role of microbial inheritance in the assembly of the plant microbiome. Trends Microbiol. 2023; 31:346–355. pmid:36481186.
  58. 58. Hodgson S, de Cates C, Hodgson J, Morley NJ, Sutton BC, Gange AC. Vertical transmission of fungal endophytes is widespread in forbs. Ecol Evol. 2014; 4:1199–1208. pmid:24834319
  59. 59. Kouzai Y, Kimura M, Yamanaka Y, Watanabe M, Matsui H, Yamamoto M, et al. Expression profiling of marker genes responsive to the defence-associated phytohormones salicylic acid, jasmonic acid and ethylene in Brachypodium distachyon. BMC Plant Biol. 2016; 16:1–11. https://doi.org/10.1186/s12870-016-0749-9.
  60. 60. Sandoya GV, Buanafina de Oliveira MM. Differential responses of Brachypodium distachyon genotypes to insect and fungal pathogens. Physiol Mol Plant Pathol. 2014; 85:53–64. https://doi.org/10.1016/j.pmpp.2014.01.001.