Publication Cover
Mycology
An International Journal on Fungal Biology
Volume 14, 2023 - Issue 4
2,117
Views
0
CrossRef citations to date
0
Altmetric
Review Article

Biology, taxonomy, genetics, and management of Zymoseptoria tritici: the causal agent of wheat leaf blotch

Pages 292-315 | Received 20 Apr 2023, Accepted 22 Jul 2023, Published online: 08 Aug 2023

ABSTRACT

Septoria tritici blotch or Septoria leaf blotch has been used for long time, but leaf blotch is a correct disease name. Moreover, Lb resistant gene is the correct name, but, not Stb gene. It has sexual and asexual parts on the mycelia, known as heterothallic fungi. Its pathogenic diversity ranged from 40% to 93% and has produced a wide variety of AvrLb6 haplotypes. M. graminicola has a plasmogamy and karyogamy sexual process. The pathogen can use macropycnidiospores, micropycnidiospores, and pycnidia vegetative growths for infection and overwintering. Synthetic M3, Kavkaz-K4500, Synthetic 6×, and TE9111 wheat genotypes have horizontal resistance. Avirulence (Avr) genes in Z. tritici and their matching wheat (R) genes indicate gene for gene mechanisms of resistance. Twenty-two R genes (vertical resistance) have been identified. In both horizontal and vertical resistance, different Lb genes have been broken down due to new Z.tritici virulent gene and currently Lb19 resistant gene is being recommended. Mixing of resistant and susceptible cultivars is also the most effective management strategy. Moreover, different cultural practices and biological control have been proposed. Lastly, different fungicides are also available. However, in developing countries cultivar mixture, isolates diversity, biological control, and epidemic studies have been greatly missed.

1. Introduction

Wheat is a highly productive crop and produced all over the world, but the production of this crop is hindered by different biotic and abiotic factors. Leaf blotch disease is one of the most important fungal diseases. Quaedvlieg et al. (Citation2004) identified Septoria tritici as the causal agent of leaf blotch disease, but, later advocated Zymoseptoria tritici (teleomorph: Mycosphaerella graminicola) (Quaedvlieg et al. Citation2011). Z. tritici is an apoplastic pathogen species with a hemibiotrophic life cycle (Ponomarenko et al. Citation2011; Fones and Gurr Citation2015; Savary et al. Citation2019).

Small chlorotic patches on the foliage are the first signs of leaf blotch disease. The lesions turn light tan as they grow and produce fruiting bodies that are deeper in colour. Especially on seedlings or leaves that were young when infected, lesions on mature leaves are frequently long, narrow, and defined by leaf veins, but they can also be shaped irregularly or be oval (Ponomarenko et al. Citation2011).

Leaf blotch disease is currently a significant and ongoing threat to wheat growers worldwide (Zhan et al. Citation2003; Ponomarenko et al. Citation2011). In fields with wheat cultivars susceptible to leaf blotch disease during severe epidemics, yield losses were 30% to 54% (Ponomarenko et al. Citation2011), 50.1% (Berraies et al. Citation2014). Furthermore, important wheat production areas in Ethiopia had a 25% to 82% loss in wheat output as a result of the frequency and intensity of leaf blotch disease infestations (Bekele Citation1985). However, another report showed a 41% yield loss in grain production (Takele et al. Citation2015).

Leaf blotch disease is widely distributed around the world, including in the largest wheat-producing countries of Argentina, Ethiopia, the United States, the Netherlands, Russia, New Zealand, Iran, Tunisia, Morocco, and Australia (Ponomarenko et al. Citation2011). The sexual reproduction of the pathogen regulates both pathogen populations and the quantity of primary inoculum (wind-dispersed ascospores) (Suffert et al. Citation2011). Hosseinnezhad et al. (Citation2014) discovered that Iranian Z. tritici isolates showed significantly different virulence patterns, with virulence spectra ranging from 40% to 90% of the tested genotypes. Eight Iranian Z. tritici isolates were tested on 45 durum wheat landraces and the isolates’ pathogenicity spectrum ranged from 42% to 82% (Ghaneie et al. Citation2012). In Ethiopia, Mekonnen et al. (Citation2020) found Z. tritici to have broad-spectrum virulence, successfully infecting 71% to 93% of the differential lines. In another country, Burcu et al. (Citation2022) identified 28 pathotypes from 60 isolates in four provinces of the central Anatolia region of Turkey.

Regarding its significant importance and wide distribution where the wheat is cultivated, various management approaches have been investigated to control this disease. The most effective techniques for reducing Z. tritici infections are to grow mixed cultivars. The traditional method of leaf blotch disease protection has been the extensive application of fungicides supported by the use of a few resistant genes. However, the reproduction cycle of Z. tritici may contribute to the high genetic diversity of the pathogen, which could hasten the fungicides rapid loss of efficacy. There are currently resistant strains for each significant fungicide group used to combat them (Stammler and Semar Citation2011; Hillocks Citation2012; Van den Berg et al. Citation2013).

Wheat resistance at adult growth stage has been promoted for management of leaf blotch disease. Studies on the gene Lb16q revealed that it encodes a cysteine-rich receptor-like kinase that confers resistance of wheat at this growth stage (Zhong et al. Citation2017; Kema et al. Citation2018; Saintenac et al. Citation2018). The resistance of wheat crop to leaf blotch disease has been investigated using quantitative genetic resistance. According to genome-wide association study (GWAS), one or more chromosomes (1A, 1B, 1D, 2B, 3B, 4A, 5A, 6A, and 6B) may be associated with dult plant resistance (Muqaddasi et al. Citation2019; Alemu et al. Citation2021). Therefore, EBW174, Lorikeet, Synthetic M3, Kavkaz-K4500, Synthetic 6×, TE9111, Selam, Mangudo, EBW174, Borstvete, Ankar, Hereford, and Aring resistant materials have been suggested (Mekonnen et al. Citation2021, Alemu et al. Citation2021; Tidd et al. Citation2023). The resistance of durum and bread wheat crop in Ethiopia has been examined; therefore, Ababa et al. (Citation2022) suggested Danda’a, HONQOLO, Digalu, Dashen, EJERSA, Alemtena, Mosobo, Hitosa, Robe, and Lelisso wheat cultivars showed resistant reaction to more than two isolates and considrered as horizontal resistance.

Moreover, for specific interactions between wheat cultivars and Z. tritici isolates, 22 Lb genes have been observed (Brading et al. Citation2002; Goodwin Citation2007; Ghaffary Citation2011; Orton et al. Citation2011; Brown et al. Citation2015). The gene Lb6 encodes wall-associated kinases that confer gene–gene mechanisms of resistance (Saintenac et al. Citation2021). Consequently, scholars reported gene for gene or specific resistance found in KK variety, Synthetic 6×, SO852, Arina, TE 9111, Veranopolis, Olaf, and Shafir wheat genotypes (Eyal et al. Citation1985; Arraiano et al. Citation2001; Adhikari et al. Citation2004; Chartrain et al. Citation2004). However, the broad-spectrum resistance (Lb16q) gene transferred to wheat cultivars has been broken down (Dalvand et al. Citation2018; Kildea et al. Citation2020; Orellana-Torrejon et al. Citation2022). Moreover, the race-specific resistance (Lb4) gene is overcome (Jackson et al. Citation2000).

Although it is a debatable idea whether cultural practices can control leaf blotch disease, its severity was decreased as the ploughing frequency increased (Bailey et al. Citation2001; Gilbert and Woods Citation2001; Bankina et al. Citation2014; Fernandez et al. Citation2016; Ababa et al. Citation2021). Crop rotation is essential to prevent the sowing of wheat in paddocks with high levels of stubble-borne inoculum (Eyal and Levy Citation1987; Ponomarenko et al. Citation2011).

Despite the lack of information on the use of microbes as leaf blotch disease control agents, some of them have been published, such as Paecilomyces lilacinus, Nigrospora sphaerica, Cryptococcus sp., Bacillus sp., and Bacillus (Perello et al. Citation2002; Kildea et al. Citation2008). As a last option, in the absence of high-resistance wheat cultivars and significant disease pressure, fungicides are the primary means of disease control (Fraaije et al. Citation2012). A variety of management techniques, including crop rotation, resistant cultivars, biocontrol, the use of fungicides, and appropriate fertilisation can be utilised to reduce disease in conservation tillage systems (Bockus and Shroyer Citation1998; Krupinsky et al. Citation2002, Citation2004, Citation2007).

Different management strategies have been developed for this disease; nevertheless, the research on the most efficient time to apply fungicides has been mostly overlooked in developing countries. As a result, farmers have applied fungicides without being aware of the best time. Cultivar mixes are also one of the most important control method, but this is also underutilised in various developing countries. Different scholars have used “Septoria tritici blotch” as a disease name and Z. tritici (former S. tritici) as a causal agent (Makhdoomi et al. Citation2015; Teferi and Gebreslassie Citation2015; McDonald and Mundt Citation2016). However, the correct name for this disease is leaf blotch disease, and its causal agent is Z. tritici (Lillemo et al. Citation2011; Salgado and Paul Citation2016; Scala et al. Citation2020). Stb name was given for resistant gene in wheat based on Septoria tritici blotch disease. Moreover, the disease and R gene names and causal agent are not clearly used in different reports. Several Lb genes have been frequently broken down by new virulent isolates, but the status of the new virulent isolates is not available in different countries. Therefore, this review was conducted to state the correct disease name and causal agent, to demonstrate new virulent isolates, and to assess the biology of leaf blotch disease and control techniques.

2. Wheat blotch (leaf and glume blotch)

2.1. Causal agent of wheat blotch (leaf and glume blotch)

Most plant diseases are named after their symptoms (grey mould, downy mildew, Powdery mildew, ring rot, white silk, sclerotinia, etc.), characteristics (soft rot, wilt, damping off, scab, leaf blight, scab, bud blight, black rot, etc.), or pathogens (Pythium, Anthrax virus disease, etc.). Wheat blotch has been named in different ways for a long time. The disease was termed as “Septoria leaf blotch” (Makhdoomi et al. Citation2015; Teferi and Gebreslassie Citation2015; McDonald and Mundt Citation2016, Citation2016), “Septoria tritici blotch” (Figueroa et al. Citation2018). In these cases, “Septoria tritici blotch” or “Septoria leaf blotch” has been used as a disease name. However, Septoria represents a genus of plant pathogenic fungi and is extremely large in number (Quaedvlieg et al. Citation2004). Moreover, fungal species belonging to Septoria are among the most widespread leaf-spotting fungi worldwide. Therefore, according to these scientists, Septoria tritici is the causal agent of leaf blotch disease (Quaedvlieg et al. Citation2004). Later, Septoria was reclassified by the genus Zymoseptoria (Quaedvlieg et al. Citation2011). Leaf blotch disease has been used as a disease name (Lillemo et al. Citation2011; Salgado and Paul Citation2016; Scala et al. Citation2020). These indicate that different scholars have used leaf blotch name in different ways.

Accordingly, Stb name was given for the resistant gene in wheat based on the Septoria tritici blotch disease and has been used for a long time (Chartrain et al. Citation2004; Arraiano et al. Citation2009; Stephens et al. Citation2021). However, since leaf blotch disease is the correct disease name, Lb is also the correct name for resistant gene in wheat. In agreement with this review, scholars have used Lr gene and Sr gene for leaf rust and stem rust, respectively (Moore et al. Citation2015; Olivera et al. Citation2018).

Similarly, Septoria nodorum blotch disease has been used wrongly (Downie et al. Citation2021), but some of the scholars used glume blotch disease (Melville and Jemmett Citation1971). Septoria nodorum is the causal agent of glume blotch disease (Lillemo et al. Citation2011; Salgado and Paul Citation2016; Scala et al. Citation2020). For few years, genus Stagnospora was also used. S. nodorum was replaced by Parastagnospora nodorum and Parastagnospora avenae f. sp. triticea (Nedyalkova et al. Citation2019). As a result, Nedyalkova et al. (Citation2019) suggested that P. nodorum and P. avenae f. sp. triticea are responsible for the glume blotch disease. S. nodorum was differentiated from S. tritici based on the length-to-width ratio of their conidia; the first two species (S. nodorum and S. avenae f. sp. triticea) are different from S. tritici. According to Cunfer and Ueng (Citation1999), conidia of the genus Septoria may be 10 times longer than broad compared to those in the genus Stagonospora. Nedyalkova et al. (Citation2019) used internal transcribed spacer (ITS) and suggested the 100% similarity of the P. nodorum and P. avenae f. sp. triticea species.

In another study, the species of Z. halophila, Z. pseudotritici, Z. tritici, Z. ardabiliae, and Z. brevis have been reported in Zymoseptoria genus (Stukenbrock et al. Citation2012). Of them, Z. ardabiliae and Z. pseudotritici have been isolated from Elymus repens, Dactylis glomerata, and Lolium perenne in Iran (Stukenbrock et al. Citation2012). As a heterothallic ascomycete, the teleomorph of Z. tritici is Mycosphaerella graminicola (Hunter et al. Citation1999; Linde et al. Citation2002; Zhan et al. Citation2003; Orton et al. Citation2011).

Under this review, my target is Z. tritici species. Mycosphaerella graminicola (Fuckel) (anamorph: S. tritici Roberge in Desmaz.) is the sexual form and is a species of filamentous fungus that belongs to kingdom Mycota, Phylum of Ascomycota, Class Dothideomycetes, in the family of Dothideaceae, Genus Mycosphaerella, Species graminicola (Orton et al. Citation2011).

2.2. Genetics and biology of the pathogen

2.2.1. Sexual reproduction of the M. graminicola

This pathogen known as heterothallic fungi due to both sexual and asexual parts are present on the mycelia. During the sexual reproduction of Mycosphaerella graminicola, the spermatia and trichogynes were produced by spermatogonia and ascogonia, respectively, and a gametes formation occurs regularly in nature during each growing season (Crous Citation1998; Hunter et al. Citation1999). Genetic diversity of M. graminicola populations has been reported all over the world (Linde et al. Citation2002; Zhan et al. Citation2003).

Plasmogamy (combination of cytoplasm): when the cytoplasm of two-parent mycelia unites, a spermatium fertilises a trichogyne. Spermatogonia gametes are produced from the male spermatia and female trichogynes (Crous Citation1998). The first mitosis started to divide the nuclei of spermatogonia gametes. After the first mitosis results, the septa are produced; then, the septa resulted in a dikaryotic cell having two haploid nuclei (a dikaryote) ().

Figure 1. Vegetative growth forms of Zymoseptoria tritici. (a) Mycelium contains pair of nuclei. (b) the arrow indicates lateral budding of macropycnidiospores to form a single cell.

Figure 1. Vegetative growth forms of Zymoseptoria tritici. (a) Mycelium contains pair of nuclei. (b) the arrow indicates lateral budding of macropycnidiospores to form a single cell.

Karyogamy: a combination of nuclei and zygote (nuclei fuse into a zygote); the second mitosis started to divide the nuclei of dikaryotic cell. The mitosis process resulted in ascogenous hyphae containing nuclei from each parent. From ascogenous hyphae, the ascus mother cells are produced. Then, ascus mother cells fuse into a zygote. Ascus mother cells are genetically distinct nuclei. The zygote nucleus undergoes meiosis, resulting in four haploid nuclei. Each of these nuclei will then undergo mitosis to yield four twin pairs of daughter cells (ascospores), which form within asci within fructifications called pseudothecia (Alexopoulos et al. Citation1996) ().

Figure 2. Process of sexual reproduction in Mycosphaerella graminicola.

Figure 2. Process of sexual reproduction in Mycosphaerella graminicola.

Chromosome re-assortment and crossing over can both lead to recombination during meiosis. Re-assortment of homologous chromosomes results in the redistribution of entire chromosomes from the parents in the progeny. When non-sister chromatids of homologous chromosomes cross over, genetic material is transferred, resulting in the production of new chromosomes that include genetic material from both parents (Alexopoulos et al. Citation1996).

Ascocarp is the genetic name for several asci formed in the pathogen fruiting body, and it determines the pathogen sexual state. This ascocarp comes in a variety of shapes and sizes, and it can be found in various parts of the host’s tissue. One of the ascocarp forms of this disease is pseudothecia (ball shape). It is made inside lesions below the host epidermis and embedded in the tissue’s protective layer, the stroma. Pseudothecia has a dark brown globose appearance, and it is structured to produce sexual fruit. Perithecia are other form of M. graminicola sexual states that are flask-shaped with a pore at the tip and sub-epidermal to the plant leaf.

After the macropycnidiospore or micropycnidiospore asexual state of M. graminicola infected the leaves, ascospores arise (Eriksen et al. Citation2003). This means that M. graminicola sexual reproduction begins after the asexual spore germinates. Each ascus produces eight ascospores at the time of ascospore development, and hyaline (clear) comprises two cells of irregular extent (Ponomarenko et al. Citation2011). The sexual mating system of M. graminicola requires the union of two compatible partners with different mating types to create sexual spores (Sanderson et al. Citation1985).

The mating system is heterothallic, which means that sexual spores must be produced by two compatible partners. In reaction to the mating pheromone released by unlike mating types, two fungal strains of opposite mating types sense each other’s existence when they come together. The homothallic ones can reproduce sexually (Zhan et al. Citation2002). For sexual reproduction, the two compatible mating types must be there in the same geographical area at a time. Ascospore is formed when two compatible haploid nuclei of different mating types come together. When strains from several isolates that have different mating types get together, ascospores are produced (Kema and Vansilfhout Citation1997); hence, their occurrence is influenced by the severity of epidemics. According to Ponomarenko et al. (Citation2011), the fungus has a bipolar, heterothallic mating system. To cause sexual reproduction, individuals of the mat1–1 and mat1–2 must come together. Sexual reproduction in M. graminicola, requires a physical encounter between two compatible strains (Mat1–1 and Mat1–2) (Kema et al. Citation1996, Citation2000; Zhan et al. Citation2002), and takes place on senescent tissues, with pseudothecia appearing 46 to 76 days after infection in field conditions (Eriksen et al. Citation2003; Suffert et al. Citation2011).

2.2.2. Asexual reproduction of the Z. tritici

Z. tritici name was given for asexual part of mycelia of the causal agent of leaf blotch disease. It is available in three different vegetative growth forms (Sanderson et al. Citation1985). The first vegetative growth form of this pathogen is macropycnidiospore with three to five septa being the most common cell form. The “Yeast-like” stage of the macropycnidiospore form has been frequently mentioned (Mehrabi et al. Citation2006). However, yeasts are single-celled organisms, whereas macropycnidiospores are multicellular () (Steinberg Citation2015). Individual cells in this multicellular structure range in size from 1.5 μm to 3.5 μm wide to 40 μm to 100 μm (Sanderson et al. Citation1985). Macropycnidiospores germinate near the apex of the plant (end). The extension of tip growth to generate thin hyphae, which are made up of highly elongated cells, is the germination process.

Micropycnidiospores are the second vegetative growth stage (Eyal et al. Citation1987). It has unicellular structures without septa (1 μm wide, 5 μm to 10 μm long). As a result, it fulfils the criteria of a “Yeast” growth type. Lateral budding from hyphae or macropycnidiospores produces these cells () (Steinberg Citation2015). Another study also showed that the yeast-like growth of this pathogen. One mutant of Z. tritici exhibiting blastosporulation (yeast-like growth) grew fatter than the other mutant on the nutrient-rich PDA medium (Francisco et al. Citation2020). The fungus can be changed morphologically through switching between hyphal growth and yeast-like development as a result of conducive environment (Mehrabi et al. Citation2006; Motteram et al. Citation2011; Francisco et al. Citation2019).

Pycnidia are the other morphological form and asexual fruiting parts of Z. tritici. They are used for dissemination by rain splashes but not used for infection. They range in size from 60 μm to 200 μm, depending on the fungal strain, infection density, and wheat cultivar stomata size variations (Sanderson et al. Citation1985; Kema et al. Citation1996). The pycnidia are implanted in mesophyll and epidermal tissue on both sides of the leaf, with an aperture (ostiole) on top. The pathogen asexual condition is represented by these three types of growth.

2.3. Life cycle of the pathogen

After coming into touch with the host leaf, both the sexual ascospores and asexual pycnidiospores germinate and penetrate plant tissues under suitable environmental conditions (Palmer and Skinner Citation2002). A major source of infection for leaf blotch disease is ascospores (Cook et al. Citation1999). After landing on the leaf, the spores nearly completely pass through the stomata (Cohen and Eyal Citation1993; Duncan and Howard Citation2000). Moreover, the fungus initially infects the host through hyphal development through stomatal openings (Kema et al. Citation1996; Shetty et al. Citation2003). Then, it smoothly expands into the mesophyll cells with hyphae spreading out in the intercellular spaces in between the mesophyll cells. In other words, after penetrating the leaf, the fungus grows intercellularly to colonise the mesophyll cells, but it does not produce any feeding structures like haustoria (Palmer and Skinner Citation2002). After 3 to 11 days of infection, hyphae start to fill the substomatal space and pre-pycnidia appear in these openings (Duncan and Howard Citation2000; Shetty et al. Citation2003). The infection continues to be asymptomatic during this phase of leaf colonisation. Dead plant cells are infrequently observed, and the leaves appear healthy (Hilu and Bever Citation1957). This stage is called the latent phase also known as the biotrophic phase. The biotrophic phase atypically long might last between 6 and 36 days, depending on the wheat genotype-fungal isolate combination (Ponomarenko et al. Citation2011).

The fungus quickly transit to necrotrophic development associated with disease lesions on the leaf surface 10 days after inoculation (dai) (Shetty et al. Citation2003). In other words, the crop cells collapse as hyphae grow intercellularly and transition from a biotrophic to a necrotrophic state, leaving behind yellow lesions or blotches. The host cell walls have not been known to be broken, because any specialised penetrations or feeding structures have not been present for Z. tritici (Kema et al. Citation1996; Shetty et al. Citation2003). Throughout the whole infection cycle, it stays in the apoplastic area to collect nutrients (Rohel et al. Citation2001). When cells break, necrotic areas of lesions release conidia into the gelatinous hygroscopic cirrhi, and pycnidia appear at necrotic locations (Eyal et al. Citation1987; Ponomarenko et al. Citation2011). When the procedure is complete, the pycnidia mature and return to the inoculums (Ponomarenko et al. Citation2011) ().

2.4. Ecology of the pathogen

On stubble, Z. tritici sustains from one season to the next. Windborne spores (ascospores) are discharged from fruiting bodies (perithecia) stayed in the stubble of previously infected plants in late fall and early winter after rain or heavy dew (Ponomarenko et al. Citation2011). These spores have a long range of dispersal (Chen et al. Citation1994). It can exist in seeds, straw, stubble, volunteers and debris, exogenous debris, adult plants, and grass species.

Once more, wheat seedlings that have grown on other wheat residues help the fungi survive. They can be discovered as pycnidia on wheat residues from one crop to the next or as mycelium in straw plants. Up to 3 years, the fungi can survive in the wheat stubble on the soil’s surface. Debris from severely diseased leaves and stems remain in fields after harvest to serve as inoculum for the following growing season (Eyal et al. Citation1987; Ponomarenko et al. Citation2011). The fungus typically overwinters in defective volunteer wheat seed, infected wheat straw from past crops, and other susceptible grasses from year to year. The fungi grow in contaminated straw and can survive for up to a year in seed form.

2.5. Epidemiology of the pathogen

Leaf blotch disease of wheat epidemics is linked to favourable weather (regular rainfall and moderate temperatures) (Campbell and Madden Citation1990), specific cultural techniques (Lovell et al. Citation2004), inoculum availability, and the presence of susceptible wheat cultivars.

2.5.1. Conducive environment

The latent period can last from 14 to 21 days at an optimal temperature (15°C to 20°C) to 40 days at 5°C, depending on the cultivar and environmental conditions including temperature and leaf moisture (Eyal et al. Citation1987; Shaw Citation1990). Z. tritici infection is temperature-dependent and necessitates wet, chilly surroundings. If the right circumstances are present, an infection can happen at any stage of a plant’s development.

Because of the temperature and moisture needs, Z. tritici infections are most likely in the early to mid-season, when temperatures are at their coolest. Infections normally begin at the base of the plant and progress upwards until high temperatures become a limiting factor. Temperatures of 15°C to 20°C are favourable for Z. tritici activity. Wet and windy weather favours epidemics of leaf blotch disease due to its specific temperature and moisture needs, but dry weather generally slows or stops growth (Lucas Citation1998).

From the lower leaves of the wheat, rain-splattered pycnidiospores move vertically up to the higher leaves. Leaf blotch disease progression from lower infected leaves to the upper plant section is frequently disrupted by long rain intervals with high temperatures (Eyal et al. Citation1987). When humidity is high and there is free water on the leaf, secondary spread from conidia/ascospore is effective for epidemics. The quantity of ascospores emitted is influenced by the length and intensity of rain, temperature, and wind (Lovell et al. Citation2004).

2.5.2. Susceptible materials of wheat

The usual vertical progression of leaf blotch disease from lower to upper leaves is impacted by the “ladder effect.” The first three to four developing leaves on both short and tall cultivars are equally spaced; however, on tall varieties, the distance between each leaf increases towards the flag leaf (Shaw Citation1990; Shaw and Royle Citation1993; Lovell et al. Citation2004).

The proximity of the upper leaves to the lower leaves in dwarf cultivars (70 cm to 90 cm) enables contact between freshly developing leaves and splashing pycnidiospores. The studies on leaf blotch disease indicated that dwarf cultivars with higher plant parts are more susceptible to the disease than taller wheat because they are closer to inoculum sources (Eyal Citation1971). Thus, it is made easier for the disease to spread from infected lower leaves. Pycnidia consequently appear earlier on the higher plant portions of dwarf cultivars than they do on the leaves of taller cultivars. As a result, both resistance and morphology-related genetic variables have an impact on leaf blotch disease spread and severity (Lovell et al. Citation2004).

As indicated in several crop disease studies, crop architecture is being used to increase disease resistance. Numerous studies have discovered that some architectural features enable pathogens to escape by preventing the pathogen from making direct contact with the host, creating an unfavourable environment for the disease, or inhibiting the pathogen’s ability to start an infection (Ando et al. Citation2007). Leaf positioning should be considered when new wheat cultivars are issued in wheat-growing locations where leaf blotch disease is a possible threat (Shaw and Royle Citation1993; Pietravalle et al. Citation2003; Lovell et al. Citation2004). According to the same theory, disease severity is significantly influenced by plant stance (Saindon et al. Citation1995), growing conditions (Ando and Grumet Citation2006), and plant height (Hilton et al. Citation1999). Numerous factors have been shown to affect disease severity (Ando et al. Citation2007), including the spatial distribution of leaf area in the canopy (Schwartz et al. Citation1978), the number of leaves (Jurke and Fernando Citation2008), and leaf shape (Gan et al. Citation2007).

With a slower stem extension, the leaf blotch disease has more time to spread from older to younger leaves (Shaw and Royle Citation1993; Lovell et al. Citation2004). Furthermore, in diseased plants, too long a maturation time results in another Z. tritici cycle of proliferation (Shaw and Royle Citation1993). Leaf blotch disease resistance was also found to be inversely correlated with plant height using quantitative trait locus (QTL) analysis. As a result, the plant height-controlling Rht-D1 gene has been found to have a considerable effect on leaf blotch disease resistance (Miedaner et al. Citation2012). Negative correlations of leaf blotch disease resistance with plant height and heading date were found in groups where Rht-D1 was not segregating (Risser et al. Citation2011; Miedaner et al. Citation2012; Andersson et al. Citation2022). However, it was discovered that tall and late wheat genotypes are less vulnerable to Z. tritici infection than short and early wheat genotypes. The co-localisation of QTL for leaf blotch disease resistance and dwarfing gene loci (Baltazar et al. Citation1990; Risser et al. Citation2011), in addition to additional plant height QTL (Eriksen et al. Citation2003), supported this morphological resistance at the gene level.

At the seedling growth stage, many bread and durum wheat-susceptible materials have been reported (Medini and Hamza Citation2008; Makhdoomi et al. Citation2015; Mekonnen et al. Citation2020; Ababa et al. Citation2022; Burcu et al. Citation2022). Shafir has Lb6 resistant gene, and Estanzuela Federal has Lb7 resistant gene; however, they did not show resistance reaction to any of the tested isolates (Makhdoomi et al. Citation2015). Mekonnen et al. (Citation2020) also stated that commercial cultivars like Laketch, Et-13A2, Gassay, Dembal, Tay, Ogolcho, Kakaba, King Bird, and Millinum were susceptible to leaf blotch disease, in addition to the differential lines Salamouni, Veranopolis, Israel-493, Tadinia, Shafir, Estanzuela Federal, Kavkaz-K4500, and Km7. Additionally, K6295-4A, LEMU, Yerer, Laketch, and ET-A132 are only a few of the commercial wheat cultivars that were identified as being susceptible (Ababa et al. Citation2022). Both durum and bread wheat were examined at the adult growth stage for susceptibility. According to Kidane et al. (Citation2017), Tossa line was more susceptible both at heading and maturity on the severity of leaf blotch disease. The most vulnerable genotypes at heading, mid-maturity, and maturity phases were EBW104 (39.8%), EBW076 (48.5%), and Paven-76 (65.3%) (Mekonnen et al. Citation2020).

2.5.3. Virulent isolates or genetic variability of Z. tritici

The sexual reproduction of Z. tritici regulates both the variety of regional populations and the quantity of primary inoculum (wind-dispersed ascospores) available to trigger subsequent epidemics (Suffert et al. Citation2011). Different virulent Z. tritici isolates from multiple countries have been reported (Grieger et al. Citation2005). The Canadian (Can1, Can2, and Can3) isolates were virulent on the same differential lines, except on the differential line has Lb6 gene (Medini and Hamza Citation2008). In another country, RM155 isolate was virulent on 14 out of 18 differential lines (Makhdoomi et al. Citation2015). K-5 isolate from the Konya province displayed virulent reaction on 10 of the 12 wheat differential set (Burcu et al. Citation2022). Mekonnen et al. (Citation2020) also reported that the I3 isolate in Ethiopia was virulent on all differential lines, including Salamouni, Veranopolis, Israel-493, Tadinia, Shafir, Estanzuela Federal, Kavkaz-K4500, and Km7. Later, EtA-19 which was virulent on the Salamouni, Veranopolis, Israel-493, Tadinia, Estanzuela Federal, Kavkaz-K4500, and Km7 differential lines was reported (unpublished data).

The diversity of Z. tritici isolates has been reported so far. Hosseinnezhad et al. (Citation2014) discovered that Iranian Z. tritici isolates showed significantly different virulence patterns, with virulence spectra ranging from 40% to 90% of the tested genotypes. Eight Iranian Z. tritici isolates were tested on 45 durum wheat landraces, and the isolate pathogenicity spectrum ranged from 42% to 82% (Ghaneie et al. Citation2012). Ghaneie et al. (Citation2012) found that eight Iranian Z. tritici isolates tested on 45 durum wheat landraces had a wide virulence spectrum. The isolates were virulent from 42% to 82% of the wheat genotypes evaluated. In another country, Burcu et al. (Citation2022) identified 28 pathotypes from 60 isolates in four provinces of the central Anatolia region of Turkey. Burcu et al. (Citation2022) detected a broad spectrum of virulence for the Z. tritici isolates using the differential wheat lines; from these pathotypes, they observed 47% pathogenic diversity of Turkey isolates. Medini and Hamza (Citation2008) identified pathotypes of Z. tritici. Therefore, they suggested 19% pathogenic diversity of Tunisia, Algeria, and Canada isolates from eight pathotypes detected. Mekonnen et al. (Citation2020) found Z. tritici to have a broad-spectrum virulence, successfully infecting 71% to 93% of the differential lines. Currently, in Ethiopia, out of 43 isolates, 25 pathotypes, were identified based on the pycnidia percent parameter; this are implying 58.1% pathogenic diversity of isolates (unpublished data).

Moreover, the diversity of the virulent gene in this pathogen has been reported. The sequencing of AvrLb6 from populations of Z. tritici had been done in two earlier investigations. To examine the diversities of AvrLb6, some of the scholars collected a global population of Z. tritici between 1990 and 2001 (Zhan et al. Citation2005; Brunner and McDonald Citation2018); moreover, other populations collected from France in 2009–2010 (Zhong et al. Citation2017). A wide variety of AvrLb6 haplotypes were discovered, along with proof of positive selection brought on by point mutations and recombination. AvrLb6 was discovered to be positioned in a very dynamic area of Z. tritici chromosome 5, where there is significant transposon activity that contributes to AvrLb6 polymorphism (Sánchez-Vallet et al. Citation2018).

Although the AvrLb6 haplotype distribution in these older collections has been extensively defined, it is uncertain how diverse the AvrLb6 haplotypes are in more recent Z. tritici populations. The changes at two amino acid residues (positions 41 and 43) in the AvrLb6 protein have been suggested as being critical for the pathogenicity of wheat cultivars carrying Lb6, the precise polymorphisms that drive the change from avirulence to virulence phenotype in the AvrLb6 protein have not yet been identified (Kema et al. Citation2018). Stephens et al. (Citation2021) analysed the diversity of the avirulence factor AvrLb6 in the recent global Z. tritici populations. They sequenced the AvrLb6 gene from recent field populations of Z. tritici isolates collected between 2013 and 2017. Therefore, they suggested that from the previous studies, the large shifts in AvrLb6 haplotype prevalence have taken place in multiple global regions over the relatively short period between samplings.

2.6. Wheat leaf blotch disease management practices

2.6.1. Cultivars mixture

Improved epidemics of pests and transferable diseases in wildlife have been linked to decreased biodiversity caused by human activities (King and Lively Citation2012; Ostfeld and Keesing Citation2012; Civitello et al. Citation2015). The deliberate introduction of genetic variation into populations of crop plants has been proposed as a possible counter measure for lowering outbreaks of plant diseases and enhancing crop function (Finckh and Wolfe Citation2006). One strategy for genetically varying agricultural plants is to simultaneously cultivate two or more genotypes of the same crop in the same location. A physical cultivar combination can be created by combining seeds from various cultivars before sowing.

The theory behind cultivar mixtures is that genetic, physiological, structural, and phonological diversity among the components of the mixture specifically, the various cultivars that make up the mixture may promote advantageous interactions not only between genotypes but also between genotypes and environments (Newton et al. Citation2009; Borg et al. Citation2018; Kristoffersen et al. Citation2020). Consequently, cultivar combinations have been shown to increase yield and stability relative to pure stands and increase crop resistance to biotic and abiotic stresses, particularly in low pesticide cropping systems (Smithson and Lenne Citation1996; Borg et al. Citation2018). The greater significance of this idea is that cultivar combinations can stop disease epidemics from spreading when the components of the mixture have various degrees of resistance to the targeted disease (Wolfe Citation1985; Finckh and Wolfe Citation2006; Gigot et al. Citation2013). If the components in the mixture are properly chosen, mixtures can also improve the quality of the final product (Finckh et al. Citation2000; Cowger and Mundt Citation2002).

Therefore, different scholars identified genetic homogeneity as a critical component in the formation of disease outbreaks in crops (Johnson Citation1961; Person Citation1966; Browning and Frey Citation1969). To combat this, multiline combinations have been proposed as a way to deliberately introduce variability into host populations (Jensen Citation1952; Borlaug Citation1959; Browning and Frey Citation1969; Wolfe and Finckh Citation1997).

Combining resistant and susceptible isogenic lines, cultivars, or even species may lessen the severity of epidemics by slowing the spread of secondary diseases via spores released from susceptible plants that have been affected. Combinations are particularly effective in combating wind-dispersed diseases with shallow spore dispersal gradients, according to simulation modelling. However, diseases like barley rhynchosporium, which have steep spore dispersion gradients and less obvious gene-for-gene interactions due to secondary transmission by splash-dispersed spores, have demonstrated that combinations can enable effective control (McDonald et al. Citation1988).

Cultivar mixes are effective against splash-dispersed pathogens such as Z. tritici, R. secalis, and S. nodorum (Jeger et al. Citation1981; McDonald et al. Citation1988; Newton et al. Citation1997). It is believed that their mode of action involves the dilution of vulnerable cultivars, the barrier effect of resistant cultivars, and induced resistance (Chin and Wolfe Citation1984). It is supposed that induced resistance can help control epidemics brought on by biotrophic, airborne diseases with highly specialised host–pathogen interactions (Calonnec et al. Citation1996). R. secalis, on the other hand, exhibits a genetically diverse population structure (McDermott et al. Citation1989). In rusts and powdery mildews, the host-pathogen connection is substantially less specific (Newton and Thomas Citation1993). Furthermore, in commercial cultivars, there are fewer sources of particular resistance to R. secalis. Cultivar combinations are therefore probably less effective against R. secalis than against biotrophic diseases. However, non-specific resistance can be used in cultivar mixtures to reduce infection (Jeger et al. Citation1981; McDonald et al. Citation1988; Mundt et al. Citation1994).

Some three-component mixtures with two susceptible and one resistant lines, as well as some two-component mixes with one resistant and one susceptible line, did not exhibit any greater scald disease than the resistant component of the mixture grown alone (McDonald et al. Citation1988). When compared to pure stands, cultivar combinations reduced leaf blotch disease by 10.6% (Kristoffersen et al. Citation2020). The susceptible pure stands received only 25% of a resistant component, which led to a surprising 48% reduction in disease levels (Ben et al. Citation2020). Different leaf blotch disease resistance scores, similar earliness, and plant height, as indicated by several trials, were taken into consideration while selecting the treatment arrangements. Their research suggests that the testing should be conducted using pure stands, two-way mixtures, and three-way mixtures (). The ratio of resistant cultivars to susceptible combinations was always 25%, 50%, or 75% (Ben et al. Citation2020). The cultivar proportion suggested by different scholars was reviewed (Ababa et al. Citation2023).

Table 1. Cultivar proportions have been employed in pure stands, two-way mixtures, and three-way mixtures planted in three replicates.

2.6.2. Resistant materials of wheat

During leaf blotch disease infection, wheat additionally secretes small proteins into the apoplast to aid in the detection of Z. tritici proteins and/or to elicit defence responses. Nuclear-binding leucine-rich repeat (NB-LRR) proteins externally bind to pathogen recognition receptors (PRRs) on the wheat cell surface to recognise Z. tritici pathogen-associated molecular pattern (PAMP) chitin after infection and start the effector-triggered immunity (ETI). Wheat has evolved a multi-layered immune system to recognise and resist Z. tritici (Jones and Dangl Citation2006). ETI or pathogen-associated effector-triggered immunity is the initial layer of plant immunity. A hypersensitive reaction to effector recognition that results in ETI is sometimes accompanied by salicylic acid (SA) signalling and systemic acquired resistance (SAR) (Kombrink and Schmelzer Citation2001). A growing portion of the wheat called apoplast acts as a barrier between the wheat and Z. tritici. For ETI, it is the area outside the plasma membrane (Block et al. Citation2014; Jashni et al. Citation2015; Wang and Wang Citation2018; Schellenberger et al. Citation2019).

Wheat can activate genes involved in pathogenesis-related (PR), including those involved in the generation of reactive oxygen species (ROS) and the activation of transcription factors. Once more, it secretes a variety of PR proteins into the apoplast that have been shown to hydrolyse glucans, chitin, and polypeptides (Ilyas et al. Citation2015; Ali et al. Citation2018), inhibit pathogen-secreted enzymes (Jashni et al. Citation2015; Wang et al. Citation2017), and phytochemically inhibit pathogen growth (Wirthmueller et al. Citation2013). Small secreted proteins (SSPs) are effectors that Z. tritici can exploit to stop or lessen ETI-induced defence responses (Palma-Guerrero et al. Citation2016). Comparative genomes and transcriptome studies have led to the identification of numerous possible Z. tritici effector genes (Gohari Citation2015; Rudd et al. Citation2015; Palma-Guerrero et al. Citation2016; Kettles et al. Citation2017; Plissonneau et al. Citation2018; Zhou et al. Citation2020).

For all of these defence mechanisms of wheat to leaf blotch disease, several models have been utilised; invasion or spatial invasion model is one of them, the condensed form reaction against apoplastic leaf pathogens (Stotz et al. Citation2014; Cook et al. Citation2015; Kanyuka and Rudd Citation2019). For this invasion model, different Lb genes have been cloned, and the resistance gene was introduced into wheat. Lb16q is the second Lb gene cloned, that is responsible for broad-spectrum resistance, and it encodes a cysteine-rich receptor-like kinase (Saintenac et al. Citation2021). Lb16q and Lb17 are two Lb genes found in Synthetic M3 that are responsible for broad-spectrum resistance. According to earlier studies, Lb17 is effective only at adult plant resistance (Tabib Ghaffary et al. Citation2012), indicating that Lb16q, which is known to confer broad resistance against Z. tritici, is principally responsible for Synthetic M3 resistance. It should be highlighted that the resistance offered by Lb16q in the field is probably less complete. The field effectiveness of Lb16q will likely decline over the following years (as was the case with Lb6 and Lb15 previously) due to the selection of Z. tritici isolates with virulence against the line carrying Lb16q resistance gene (Dalvand et al. Citation2018; Kildea et al. Citation2020). When important resistance genes break, agricultural systems are vulnerable because wheat lacks broad-spectrum leaf blotch disease resistance. Both Lb6 and Lb15 have been extensively utilised in Northern Europe and were initially very effective; however, because of the selection pressures brought up by their extensive usage, Z. tritici has now extensively broken both of them (Chartrain et al. Citation2004; Arraiano et al. Citation2009; Stephens et al. Citation2021).

It has been demonstrated that Kavkaz-K4500, one of the most reliable sources of field resistance utilised in breeding, possesses at least five qualitative resistance genes, including Lb6, Lb7, Lb10, and Lb12 (Chartrain et al. Citation2005). Although many international Z. tritici isolates are virulent on it in laboratory tests (Chartrain et al. Citation2004, Citation2005), this combination of Lb genes appears to be sufficient to make Kavkaz-K4500 resistant to leaf blotch disease under field conditions. This may indicate high genetic diversity differences between UK and international Z. tritici populations or could be related to the different levels of inoculum used in laboratory vs field trials.

Additionally, some lines including TE9111 (containing Lb6, Lb7, and Lb11) and Lorikeet (containing Lb19) were resistant to pycnidia production from every Z. tritici strain (Tidd et al. Citation2023). Breeders should be most interested in these lines. Previous studies indicated that Lb6 and Lb7 genes did not show resistance to UK Z. tritici populations; therefore, Lb5, Lb11, and either Lb10 or Lb12 are responsible (Czembor et al. Citation2011; Makhdoomi et al. Citation2015; Stephens et al. Citation2021). Currently, Lb5 and Lb11 appear to be the optimal resistances to protect the durability of Lb19 in future wide use (Tidd et al. Citation2023).

Black and Gallegly (Citation1957) described field resistance (also known as horizontal resistance). Eyal et al. (Citation1985) proposed that KK variety resistance could be governed by up to seven genes. Other lines include TE 9111, Veranopolis, and Olaf have four, whereas Chaucer and Catbird have two resistant genes (Chartrain et al. Citation2004). KK variety has been used as a source of resistance to leaf blotch disease for many years and is one of the most resistant in the field (Eyal et al. Citation1985; Kema and Vansilfhout Citation1997; Arraiano et al. Citation2001). Again, Chartrain et al. (Citation2004) suggested the most resistant line was Senat, followed by Gene, Milan, Israel 493, and Chaucer.

Germplasm from China (Synthetic 6×), Latin America (SO852), and Europe (Arina, and Shafir) has been identified as potential resistance sources (Arraiano et al. Citation2001, Citation2007; Adhikari et al. Citation2004; Chartrain et al. Citation2004). However, several sources are not suitable for commercial breeding (Arraiano et al. Citation2001, Citation2007; Adhikari et al. Citation2004; Chartrain et al. Citation2004). Makhdoomi et al. (Citation2015) proposed five wheat genotypes for breeding efforts, all of which were resistant to the six isolates tested in Iran. According to (Mekonnen et al. Citation2020), the Hidase cultivar displayed three to five resistance levels in Ethiopia. Danda’a, HONQOLO, Digalu, Dashen, EJERSA, Alemtena, Mosobo, Hitosa, Robe, and Lelisso materials were proposed as broad-spectrum materials (Ababa et al. Citation2022).

The adult plant resistance of wheat crop against leaf blotch disease has been discovered by many investigators. Genome-wide association study (GWAS) has been successfully used to mine multiple putative QTLs/genes related to agronomically important features in a variety of plants, including disease resistance (Bartoli and Roux Citation2017; Kidane et al. Citation2017; Juliana et al. Citation2018; Odilbekov et al. Citation2019; Alemu et al. Citation2021; Mekonnen et al. Citation2021). For leaf blotch disease, several valuable QTLs/genes have been discovered by GWAS and linkage mapping (Schilly et al. Citation2011; Tabib Ghaffary et al. Citation2012; Miedaner et al. Citation2013; Dreisigacker et al. Citation2015; Mirdita et al. Citation2015; Muqaddasi et al. Citation2019; Odilbekov et al. Citation2019).

Again, adult plant resistance was detected at wheat heading, mid-maturity, and maturity growth stages (Dreisigacker et al. Citation2015; Kidane et al. Citation2017; Muqaddasi et al. Citation2019; Alemu et al. Citation2021; Mekonnen et al. Citation2021). They proposed that the disease was most severe at the maturity growth stage, and the response of each genotype was considerably different at each stage. Selam showed lower leaf blotch disease severity at the heading and maturity growth stages. Mangudo expressed lower leaf blotch disease severity at maturity and heading growth stages. Therefore, both cultivars have more resistance than other cultivars (Kidane et al. Citation2017). EBW174 showed resistance at all growth stages having a leaf blotch disease severity of 5.3% (Mekonnen et al. Citation2020). Some genotypes performed better for leaf blotch disease resistance than others. Borstvete was more resistant than Gotland and Ankar. Again, Hereford was more resistant than Sweden and Aring from Denmark (Alemu et al. Citation2021). When compared to the line IAS20 × 5/H567.71, RPB709.71/COC parent displayed greater resistance. In the CIMMYT wheat lines IAS20 × 5/H567.71 and RPB709.71/COC, there are a total of five consistent QTL for leaf blotch disease resistance across all environmental conditions (Dreisigacker et al. Citation2015).

A genome-wide association (GWA) scan utilising SDS data obtained at the heading showed likely QTLs on chromosomes 1D, 2A, 3A, 3D, 5A, 7A, and 7D. Therefore, 33 potential quantitative trait loci (QTLs) were discovered in bread wheat (Mekonnen et al. Citation2021). Furthermore, effective putative QTLs on chromosomes 1B, 3D, and 7B were revealed in the association analysis for SDS at the mid-maturity growth stage. Similarly, at the maturity stage, potential QTLs on chromosomes 1D, 4A, and 6A were discovered. On durum wheat, the GWA scan identified five significant potential QTL for leaf blotch disease resistance (Kidane et al. Citation2017).

After, Flor (Citation1942) reported the gene for gene hypothesis in flux and rust pathosystem, different scholars reported this type of resistance in different crop pathosystems. The specificity in Z. tritici and wheat pathosystems had been debated for more than 20 years. Then, in 1973, the specificity of this pathogen was confirmed (Eyal et al. Citation1973). Of 89 genomic regions, 27 had been detected at the seedling growth stage. Moreover, 22 Lb genes have been observed for specific resistance (Brading et al. Citation2002; Brown et al. Citation2015), including 12 isolate-specific genes and 10 non-isolate-specific genes from wheat (Ghaffary et al. Citation2018; Yang et al. Citation2018).

Lb6 is the one of Lb gene cloned that is responsible for race-specific resistance and it encodes a wall-associated-like receptor kinase (Zhong et al. Citation2017; Kema et al. Citation2018; Saintenac et al. Citation2018). Lb1 is the first; it is available to growers in the cultivars Oasis and Sullivan in 1975. The efficacy of this gene had been used for 3 years (Cowger et al. Citation2000; Adhikari et al. Citation2004; Singh et al. Citation2016). Saw is a cross between Tadorna, Cleo, and Inia 66, then released in 1984 with Lb4 gene (Somasco et al. Citation1996). Lb4 also demonstrated a respectable level of durability, lasting approximately 15 years. The Cougar variety was released in 1992, but it had served for 3 years (Cowger et al. Citation2000), due to Cougar-virulent strains of Z. tritici in the UK (Kildea et al. Citation2021).

In general, Lb4 and Lb6 genes have lost their resistance due to Lb6q avr gene (Brading et al. Citation2002; Adhikari et al. Citation2004; Stephens et al. Citation2021), whereas Lb2/11/WW and Lb18 genes have lost their resistance due to other new virulent isolates (Tabib Ghaffary et al. Citation2011; Liu et al. Citation2013; Dreisigacker et al. Citation2015).

The majority of commercial wheat cultivars are susceptible to the disease despite the discovery of 22 Lb resistant genes. Other scholars discovered 17 significant resistance genes (Lb genes) (Goodwin Citation2007; Ghaffary Citation2011; Orton et al. Citation2011). To date, 60 major leaf blotch disease resistance genes have been mapped and found, ranging from Lb 1 to Lb 22 for race specific resistance (Brading et al. Citation2002; Brown et al. Citation2015). Today commercially available varieties are largely partially resistant to leaf blotch disease (i.e. they are moderately susceptible).

2.6.3. Cultural practices

In the concept of crop rotation, prevention of planting wheat in farmlands with high levels of stubble-borne inoculum is very crucial (Eyal et al. Citation1987; Ponomarenko et al. Citation2011). Two to three years of crop rotation, tilling, and the eradication of volunteer are very important to reduce the leaf blotch disease. At different times, a disease outbreak can be achieved with a one-year rotation. Furthermore, Pedersen (Citation1992) discovered that a 1-year break between wheat crops was effective in reducing leaf blotch disease severity under unfavourable environmental conditions, whereas a 2-year crop rotation between wheat crops was necessary under ideal environmental conditions. However, during extremely dry seasons, the fungus can persist on stubble for up to 18 months (Bankina et al. Citation2014). Krupinsky (Citation1999) recommended crop rotation as a technique to hasten the breakdown of infected crop residue, while non-host crops were being grown. The author suggested that crop rotation will lower the pathogen inoculum level but not eradicate the disease. However, some findings claim that crop rotation does not affect leaf blotch disease. Tan spot severity was significantly higher in repeated wheat sowings; however, crop rotation did not affect leaf blotch disease development (Bankina et al. Citation2014).

As the frequency of ploughing increased, the severity of leaf blotch disease was reduced (Bailey et al. Citation2001; Gilbert and Woods Citation2001; Bankina et al. Citation2014; Fernandez et al. Citation2016; Ababa et al. Citation2021). Numerous studies on the effect of soil tillage on leaf blotch disease have been conducted. Despite the contradicting findings, conventional tillage-ploughed plots had a higher severity of leaf blotch disease than plots with alternative tillage techniques (Gilbert and Woods Citation2001; Krupinsky et al. Citation2007; Bürger et al. Citation2012).

Surprisingly, Huber et al. (Citation1987) in Indiana discovered that the intensity of tan spots on wheat cultivars was reduced as the N rate increased. However, it has been noted that using a lot of N fertiliser can make the severity of leaf and glume blotch infections on winter wheat worse (Broscious et al. Citation1985; Ditsch and Grove Citation1991; Howard et al. Citation1994). Depending on the geography and the local environment, increasing N fertiliser rates either appear to have a positive, negative, or no effect on the severity of the leaf blotch disease (Krupinsky Citation1999). As N fertiliser increased, the plant height was also increased, therefore resulted in the reduction of leaf blotch disease severity. In reverse, as nitrogen increased, the tiller number was also increased and then resulted in the increments of the leaf blotch disease severity (Krupinsky Citation1999). Leaf blotch disease levels can also be reduced by wheat traits such as taller plant height and late heading date or flowering time that contributes to disease escape (Simó et al. Citation2004; Arraiano et al. Citation2009; Brown et al. Citation2015).

Studies on crop diseases indicate that cropping practices, such as nitrogen fertilisation, planting density, and sowing date, have a variety of effects on disease development, including changes to the canopy’s architectural structure. However, these effects are intermittent, and it may not be clear how to interpret them. According to some authors, increasing planting density caused the foliar disease to occur more frequently (Ando and Grumet Citation2006; Gan et al. Citation2007; Jurke and Fernando Citation2008).

The development of a favourable microclimate, such as elevated relative humidity, or modifications in canopy architecture was thought to be the causes of the density effect (Tompkins et al. Citation1993). According to Pielaat et al. (Citation2002), denser canopies increase leaf-to-leaf contact and facilitate the spread of disease through the canopy. The substantial increase in wheat tiller numbers with increased plant density may facilitate the deposition of Z. tritici spores (Broscious et al. Citation1985). However, there are some contradictions in the literature when it comes to the density impact. Pfleeger and Mundt (Citation1998) discovered that plant density has only a little impact on disease development. Less dense plant stands had a sparser canopy as a result of lower plant densities, which increased the chance of rainfall splashing on lower leaves and accelerated the spread of disease (Eyal Citation1981).

According to various scientists, the planting date has a significant impact on disease (Shaner et al. Citation1975; Thomas et al. Citation1989; Shaw and Royle Citation1993; Hailemariam et al. Citation2020), and the crops planted early in the season have a higher risk of infection. After a leaf blotch disease outbreak, postponing of wheat early planting is very important since a lot of ascospores are discharged early in the season. Early-planted crops are more susceptible to infection. This could be early-sown plants develop more leaves, which means more inoculums are present (Shaw and Royle Citation1993). Furthermore, with early-sown crops, the infection has more time to migrate from older to younger leaves due to the slower stem expansion (Shaw and Royle Citation1993; Lovell et al. Citation2004). Because early-sown plants have a lot of leaves, early-planted fields might have more disease activity (Shaw and Royle Citation1993), which leads to a higher occurrence of inocula (Shaner et al. Citation1975; Shaw and Royle Citation1993).

Some plants are only susceptible to a pathogen at a specific growth stage (young leaves, stems, or fruits; blossoming or fruiting; maturity and early senescence); as a result, if the pathogen is absent or inactive at this specific stage, such plants avoid infection, though the latent infection may occur (Agrios Citation2005).

2.6.4. Biological control

Global populations of Z. tritici have developed resistance to the most frequently employed fungicides, including azole and quinone outside inhibitors (QoI) (Fraaije et al. Citation2005). To combat the rise in fungicide resistance, alternative control measures including biological control are becoming increasingly important. Despite these obstacles, only a small portion of the pesticide industry, which is still dominated by synthetic chemicals, uses safe and environmentally acceptable plant protection methods (Lynch et al. Citation2016).

Biological control agents are being developed as an alternate control approach. Treatments for biological control include living microorganisms or abiotic substances that can (i) protect plants by creating antibiotics or other compounds that hinder the growth of pathogens; (ii) compete with pathogens for nutrients and space; (iii) cause plant resistance. Few studies have been reported on the use of Paecilomyces lilacinus, Nigrospora sphaerica, Cryptococcus sp., Bacillus sp. or their metabolites as leaf blotch disease control agents (Perello et al. Citation2002; Kildea et al. Citation2008).

According to Lynch et al. (Citation2016), biological organisms such as lactic bacterial strains have a strong inhibitory action against leaf blotch disease. They also suggested that LAB can stop the leaf blotch disease from growing. Lactobacillus brevis JJ2P, Lactobacillus arizonensis R13, and Lactobacillus reuteri R2 suppressed as seen by massive mycelium on modified MRS (De Man, Rogosa, and Sharpe) agar. Lactic acid bacteria (LAB) were used as natural biocontrol agents in a variety of foods and feeds (Stiles Citation1996; Carr et al. Citation2002; Schnürer and Magnusson Citation2005; Broberg et al. Citation2007). Antifungal activity of LAB has been demonstrated against an extensive range of fungi (Corsetti et al. Citation1998; Stiles et al. Citation2002). Again, various data suggested that Trichoderma spss was found to have the ability to control the growth and severity of leaf blotch disease. Trichoderma harzianum and Gliocladium roseum were used as biological controls in the greenhouse and in vitro. Therefore, the severity of leaf blotch disease was significantly reduced by these microbes (Perelló et al. Citation1997) but, T. harzianum reduced Z. tritici growth more effectively than G. roseum. Both biocontrols were capable of completing and covering the colony expansion of leaf blotch disease. Some studies showed that lipopeptides from B. subtilis controlled Z. tritici. Therefore, as Z. tritici treated by mycosubtilin alone or in mixture with surfactin or with both surfactin and fengycin up to 82% reduction was resulted (Mejri et al. Citation2018). Bacillus subtilis ATCC 10,783, B. cereus ATCC 11,778, B. licheniformis NRRLB-510, B. pumilus ATCC 7061, Brevibacillus laterosporus BLA170, and Paenibacillus polymyxa NA are some of the microbes recommended for Z. tritici management (Alippi et al. Citation2000; Dutilloy et al. Citation2022).

2.6.5. Fungicides

Fungicides are the principal way of disease control during the lack of high-resistance wheat cultivars and high disease pressure. The use of foliar fungicide sprays and seed treatment can both provide chemical control. To manage leaf blotch disease, three main fungicide groups have been suggested: Succinate dehydrogenase inhibitors (SDHI), 14a-demethylase inhibitors (DMIs), and Quinone outside inhibitors (QoI).

Since many years ago, compounds from those three families have been used effectively (Mäe et al. Citation2020). 14a-demethylase inhibitors (DMIs), like epiconazole and prothioconazole, and Sterol 14-demethylation inhibitors (DMI), also known as triazoles, are a type of sterol biosynthesis inhibitor (SBI). The other family is Succinate dehydrogenase inhibitors (SDHI) like Fluxapyroxad, isopyrazam, bixafen, and biscalid (Fraaije et al. Citation2012).

Due to Benzimidazole fungicides (Garnault et al. Citation2019) and quinone outside inhibitors (QoIs) (Fraaije et al. Citation2005) resistance developed, systemic demethylation inhibitors (DMIs; azole fungicides) and the protective multi-site inhibitor chlorothalonil have been recommended. Again, azole fungicide resistance developed in Z. tritici populations, field performance of several products has been negatively impacted (Clark Citation2006). Therefore, further changes in susceptibility to various azoles due to the ongoing evolution of novel CYP51 (sterol 14ademethylase) variations (Brunner et al. Citation2008; Cools and Fraaije Citation2008; Cools et al. Citation2011; Leroux and Walker Citation2011), and guarantee long-term sustainable disease management, new modalities of intervention are urgently required. Moreover, sensitivity in Z. tritici populations has been declining over time (Clark Citation2006). Recently, carboxamide fungicides such as boscalid (2005), isopyrazam (2010), and bixafen (2011) are entered the global cereal market. These fungicides inhibited succinate dehydrogenase (Sdh) of the mitochondrial respiratory chain (Mäe et al. Citation2020).

Various scientists have proposed that the criteria should be considered when utilising fungicides. The tan spot spread quickly soon after flowering. Therefore, early sprayings are ineffective under these circumstances, and these factors are essential in determining the best time to employ fungicides (Bankina and Priekule Citation2011). Other scholars have discovered similar patterns: After growth stage 59, leaf necrotic spots appeared quickly, and a single fungicide handling was effective (Wyczling et al. Citation2010). Wegulo et al. (Citation2009) established that leaf blotch disease (induced by many diseases) developed rapidly just after flowering, with disease severity growing exponentially until milk maturity. At the flowering stage, the severity of the disease had the largest association with yield reductions (Wegulo et al. Citation2009).

The time of fungicide applications will be crucial for achieving effective disease control in high-risk areas. In early-sown susceptible varieties, a fungicide application at development stages 31 to 32 may be necessary to control the disease and protect new leaves. Once the flag leaf has fully developed at GS39, another application of fungicide may be necessary to safeguard the upper canopy. To find the best fungicide treatment options, numerous studies have been undertaken all around the world. Because lower fungicide doses are not allowed, research must focus on establishing the best time to spray and disease damage thresholds. Compared to one application at GS 31 or GS 39, two applications of trifloxystrobin propiconazole (at GS 31 and again at GS 39) reduced disease severity and AUDPC while increasing yield. This was a done deal because the two applications extended the period of disease control compared to the single application (Wegulo et al. Citation2009).

Bankina et al. (Citation2014) found that the fungicide application yielded 9.8% to 13.5% more than control treatment. Wegulo et al. (Citation2012) established that fungicides containing strobilurins considerably increased yield, with a mean yield difference between treated and untreated regions ranging from 12.6% to 29.4%. Different fungicide spraying techniques are utilised, but for rigorous control of winter wheat, two (and occasionally three) applications are often made (Bankina et al. Citation2014).

Because Septoria spp. has evolved resistance to a variety of fungicides, it is recommended to employ SDHI fungicides in combination with a maximum of two sprays per season as a preventive measure. Fungicide-resistant isolates are less common when fungicidal compounds with diverse modes of action, such as azoles and SDHI (Fraaije et al. Citation2012).

There are a few approaches that have been proposed to minimise the selection rate for new mutations. The first method is to use various triazoles since Z. tritici fungus mutations do not affect all triazole fungicides equally. The same triazole fungicide should not be applied if many treatments are necessary for single season. The second choice is to utilise recommended fungicides such as triazoles, flutriafol, propiconazole and cyproconazole or tebuconazole and flutriafol.

Utilising fungicides with different modes of action is the third approach; however, there is a small selection of fungicides with various modes of action. Utilising solutions that combine triazole and strobilurin fungicide may assist to reduce the probability of resistance. Considering their distinct mechanism of action, strobilurins are expected to have a high likelihood of developing resistance. Again, two to four applications of fungicide mixtures during the growing season increase yields by about two tonnes per hectare (Berry et al. Citation2008). The best method for managing leaf blotch disease is an integrated strategy that includes variety selection, cultural methods, crop rotation, and fungicides.

3. Conclusion

The current review suggests that leaf blotch disease is the correct disease name and it is caused by Zymoseptoria tritici. Moreover, leaf blotch (Lb) gene is the correct name of resistant gene than Septoria tritici blotch (Stb) gene. Different cultural, resistant wheat materials and biological methods of leaf blotch disease management have been advocated. Lastly, when we miss using the resistance materials and at high disease pressure, fungicide application at Gs 39 and after flowering is critical because the pathogen can harm the crop at this growth stage.

Pathogenicity levels, population diversity, and pathotype identification are also critical for investigating host resistance, but these are similarly restricted in some parts of the world. In developing countries, studies on epidemiological factors such as soil types, plant architecture, and other cultural practices are seriously lacking. Other work waiting for wheat pathologists includes molecular investigations on pathogen diversity and pathotype identification as well as spatial and temporal distribution study of the pathogen. Wheat durable and race-specific resistances have been weakened globally; as a result, regular breeding is crucial. Furthermore, it should be taken into account in future studies since prompt detection of virulent isolates is not always observed in certain countries. The probability of vertical resistance of wheat materials has been broken down, it may occur due to high genetic variability. Therefore, timely pathotype identification is very important. Different genes in horizontal and vertical resistance have been broken down; therefore, cloning or transferring resistant genes are very important. Gene pyramiding is very important to make the wheat gene diversities. Genomic regions of wheat cultivars need more attention to map the resistant gene areas on the chromosome.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been republished with minor changes. These changes do not impact the academic content of the article.

References

  • Ababa G, Adugna G, Hundie B. 2021. Prevalence, intensity, and morphological variability of wheat blotch (Zymoseptoria tritici) in Oromia, Ethiopia. Int J Phytopathol. 10(3):167–180. doi: 10.33687/phytopath.010.03.3899.
  • Ababa G, Adugna G, Hundie B. 2022. Seedling resistance of wheat cultivars to Zymoseptoria tritici disease in Ethiopia. Indian Phytopathol. 75(4):1–12. doi: 10.1007/s42360-022-00549-x.
  • Ababa G, Kesho A, Tadesse Y, Amare D. 2023. Reviews of taxonomy, epidemiology, and management practices of the barley scald (Rhynchosporium graminicola) disease. Heliyon. 9(3):e14315. doi: 10.1016/j.heliyon.2023.e14315.
  • Adhikari T, Cavaletto J, Dubcovsky J, Gieco J, Schlatter A, Goodwin S. 2004. Molecular mapping of the Stb4 gene for resistance to Septoria tritici blotch in wheat. Phytopathol. 94(11):1198–1206. doi: 10.1094/PHYTO.2004.94.11.1198.
  • Adhikari TB, Wallwork H, Goodwin SB. 2004. Microsatellite markers linked to the Stb2 and Stb3 genes for resistance to Septoria tritici blotch in wheat. Crop Sci. 44(4):1403–1411. doi: 10.2135/cropsci2004.1403.
  • Adhikari TB, Yang X, Cavaletto JR, Hu X, Buechley G, Ohm HW, Shaner G, Goodwin SB. 2004. Molecular mapping of Stb1, a potentially durable gene for resistance to Septoria tritici blotch in wheat. Theor Appl Genet. 109(5):944–953. doi: 10.1007/s00122-004-1709-6.
  • Agrios GN. 2005. Plant pathology. 5th ed. USA: Elsevier Academic Press p. 79–103
  • Alemu A, Brazauskas G, Gaikpa DS, Henriksson T, Islamov B, Jørgensen LN, Koppel M, Koppel R, Liatukas Ž, Svensson JT, et al. 2021. Genome-wide association analysis and genomic prediction for adult-plant resistance to Septoria tritici blotch and powdery mildew in winter wheat. Front Genet. 12:627. doi: 10.3389/fgene.2021.661742.
  • Alexopoulos CJ, Mims CW, Blackwell M. 1996. Introductory Mycology. New York: John Willey and Sons. Inc. p. 868.
  • Ali S, Ganai BA, Kamili AN, Bhat AA, Mir ZA, Bhat JA, Tyagi A, Islam ST, Mushtaq M, Yadav P, et al. 2018. Pathogenesis-related proteins and peptides as promising tools for engineering plants with multiple stress tolerance. Microbiol. 212:29–37. doi:10.1016/j.micres.2018.04.008.
  • Alippi A, Perelló A, Sisterna GN, Cordo C. 2000. Potential of spore-forming bacteria as biocontrol agents of wheat foliar diseases under laboratory and greenhouse conditions. J Plant Dis Prot. 107:155–169.
  • Andersson B, Djurle A, Ørum JE, Jalli M, Ronis A, Ficke A, Jørgensen LN. 2022. Comparison of models for leaf blotch disease management in wheat based on historical yield and weather data in the Nordic-Baltic region. Agron Sustain Dev. 42(3):42. doi: 10.1007/s13593-022-00767-7.
  • Ando K, Grumet R. 2006. Evaluation of altered cucumber plant architecture as a means to reduce Phytophthora capsici disease incidence on cucumber fruit. J Amer Soc Hortic Sci. 131(4):491–498. doi: 10.21273/JASHS.131.4.491.
  • Ando K, Grumet R, Terpstra K, Kelly JD. 2007. Manipulation of plant architecture to enhance crop disease control. CAB Reviews: Perspectives In Agri, Veterinary Sci, Nutri And Nat Res. 2(8):1–8. doi: 10.1079/PAVSNNR20072026.
  • Arraiano LS, Balaam N, Fenwick PM, Chapman C, Feuerhelm D, Howell P, Smith SJ, Widdowson JP, Brown JKM. 2009. Contributions of disease resistance and escape to the control of Septoria tritici blotch of wheat. Plant Pathol. 58(5):910–922. doi: 10.1111/j.1365-3059.2009.02118.x.
  • Arraiano LS, Brading PA, Brown JKM. 2001. A detached seedling leaf technique to study resistance to Mycosphaerella graminicola (anamorph Septoria tritici) in wheat. Plant Pathol. 50(3):339–346. doi: 10.1046/j.1365-3059.2001.00570.x.
  • Arraiano LS, Chartrain L, Bossolini E, Slatter HN, Keller B, Brown JKM. 2007. A gene in European wheat cultivars for resistance to an African isolate of Mycosphaerella graminicola. Plant Pathol. 56(1):73–78. doi: 10.1111/j.1365-3059.2006.01499.x.
  • Bailey KL, Gossen BD, Lafond GP, Watson PR, Derksen DA. 2001. Effect of tillage and crop rotation on root and foliar diseases of wheat and pea in Saskatchewan from 1991 to 1998: Univariate and multivariate analyses. Can J Plant Sci. 81(4):789–803. doi: 10.4141/P00-152.
  • Baltazar BM, Scharen AL, Kronstad WE. 1990. Association between dwarfing genes ‘Rht1’and ‘Rht2’and resistance to Septoria tritici blotch in winter wheat (Triticum aestivum L. em Thell). Theor Appl Genet. 79(3):422–426. doi: 10.1007/BF01186089.
  • Bankina B, Gaile Z, Balodis O, Bimšteine G, Katamadze M, Kreita D, Paura L, Priekule I. 2014. Harmful winter wheat diseases and possibilities for their integrated control in Latvia. Acta Agric Scand B Soil Plant Sci. 64(7):615–622. doi: 10.1080/09064710.2014.949296.
  • Bankina B, Priekule I. 2011. A review of tan spot research in the Baltic countries: occurrence, biology and possibilities of control. Žemdirbystė (Agriculture). 98(1):3–10.
  • Bartoli C, Roux F. 2017. Genome-wide association studies in plant pathosystems: toward an ecological genomics approach. Front Plant Sci. 8:763. doi: 10.3389/fpls.2017.00763.
  • Bekele E. 1985. A review of research on diseases of barley, tef and wheat in Ethiopia. A review of crop protection research in Ethiopia. Addis Ababa: Institute of Agricultural Research (IAR), Ethiopia. p. 79–107.
  • Ben MS, Karisto P, Abdedayem W, Laribi M, Fakhfakh M, Kouki H, Mikaberidze A, Yahyaoui A. 2020. Improved control of Septoria tritici blotch in durum wheat using cultivar mixtures. Plant Pathol. 69(9):1655–1665. doi: 10.1111/ppa.13247.
  • Berraies S, Gharbi MS, Rezgui S, Yahyaoui A. 2014. Estimating grain yield losses caused by septoria leaf blotch on durum wheat. Chil J Agric Res. 74(4):432–437. doi: 10.4067/S0718-58392014000400009.
  • Berry PM, Kindred DR, Paveley ND. 2008. Quantifying the effects of fungicides and disease resistance on greenhouse gas emissions associated with wheat production. Plant Pathol. 57(6):1000–1008. doi: 10.1111/j.1365-3059.2008.01899.x.
  • Black W, Gallegly ME. 1957. Screening of Solanum species for resistance to physiologic races of Phytophthora infestans. Am Potato J. 34:273–281. doi: 10.1007/BF02854886.
  • Block A, Toruño TY, Elowsky CG, Zhang C, Steinbrenner J, Beynon J, Alfano JR. 2014. The Pseudomonas syringae type III effector HopD1 suppresses effector-triggered immunity, localizes to the endoplasmic reticulum, and targets the Arabidopsis transcription factor NTL9. New Phytol. 201(4):1358–1370. doi: 10.1111/nph.12626.
  • Bockus WW, Shroyer JP. 1998. The impact of reduced tillage on soilborne plant pathogens. Annu Rev Phytopathol. 36(1):485–500. doi: 10.1146/annurev.phyto.36.1.485.
  • Borg J, Kiær LP, Lecarpentier C, Goldringer I, Gauffreteau A, Saint-Jean S, Barot S, Enjalbert J. 2018. Unfolding the potential of wheat cultivar mixtures: A meta-analysis perspective and identification of knowledge gaps. Field Crops Res. 221:298–313. doi:10.1016/j.fcr.2017.09.006.
  • Borlaug NE 1959. The use of multilineal or composite varieties to control airborne epidemic diseases of self-pollinated crop plants. In: Jenkins BC, editor. Proceedings of the First International Wheat GeneticsSymposium; Winnipeg, Canada: Universityof Manitoba. p 12–27.
  • Brading PA, Verstappen ECP, Kema GHJ, Brown JKM. 2002. A gene-for-gene relationship between wheat and Mycosphaerella graminicola, the Septoria tritici blotch pathogen. Phytopathol. 92(4):439–445. doi: 10.1094/PHYTO.2002.92.4.439.
  • Broberg A, Jacobsson K, Ström K, Schnürer J. 2007. Metabolite profiles of lactic acid bacteria in grass silage. Appl Environ Microbiol. 73(17):5547–5552. doi: 10.1128/AEM.02939-06.
  • Broscious SC, Frank JA, Frederick JR. 1985. Influence of winter wheat management practices on the severity of powdery mildew and Septoria blotch in Pennsylvania. Phytopathol. 75(5):538–542. doi: 10.1094/Phyto-75-538.
  • Brown JKM, Chartrain L, Lasserre-Zuber P, Saintenac C. 2015. Genetics of resistance to Zymoseptoria tritici and applications to wheat breeding. Fungal Genet Biol. 79:33–41. doi: 10.1016/j.fgb.2015.04.017.
  • Browning JA, Frey KJ. 1969. Multiline cultivars as a means of disease control. Annu Rev Phytopathol. 7(1):355–382. doi: 10.1146/annurev.py.07.090169.002035.
  • Brunner PC, McDonald BA. 2018. Evolutionary analyses of the avirulence effector AvrStb6 in global populations of Zymoseptoria tritici identify candidate amino acids involved in recognition. Mol Plant Pathol. 19:1836–1846. doi:10.1111/mpp.12662.
  • Brunner PC, Stefanato FL, McDonald BA. 2008. Evolution of the CYP51 gene in Mycosphaerella graminicola: evidence for intragenic recombination and selective replacement. Mol Plant Pathol. 9(3):305–316. doi: 10.1111/j.1364-3703.2007.00464.x.
  • Burcu E, Orhan T, Ali B, Yıldırım F, Ölmez F, Akci N, Karakaya A. 2022. Pathotype diversity of Zymoseptoria tritici (Desm. Quaedvlieg & Crous) isolates collected from Central Anatolia, Turkey nine of the isolates were virulent on the resistance. J Phytopathol. 21:1–10. doi: 10.1111/jph.13122.
  • Bürger J, Günther A, de Mol F, Gerowitt B. 2012. Analysing the influence of crop management on pesticide use intensity while controlling for external sources of variability with linear mixed effects models. Agric Syst. 111:13–22. doi: 10.1016/j.agsy.2012.05.002.
  • Calonnec A, Goyeau H, de Vallavieille-Pope C. 1996. Effects of induced resistance on infection efficiency and sporulation of Puccinia striiformis on seedlings in varietal mixtures and on field epidemics in pure stands. null. 102(8):733–741. doi: 10.1007/BF01877147.
  • Campbell CL, Madden LV. 1990. Introduction to plant disease epidemiology. New York: John Wiley & Sons.
  • Carr FJ, Chill D, Maida N. 2002. The lactic acid bacteria: a literature survey. Crit Rev Microbiol. 28(4):281–370. doi: 10.1080/1040-840291046759.
  • Chartrain L, Berry ST, Brown JKM. 2005. Resistance of wheat line kavkaz-K4500 L.6.A.4 to Septoria tritici blotch controlled by isolate-specific resistance genes. Phytopathol. 95(6):664–671. doi: 10.1094/PHYTO-95-0664.
  • Chartrain L, Brading PA, Makepeace JC, Brown JKM. 2004. Sources of resistance to Septoria tritici blotch and implications for wheat breeding. Plant Pathol. 53(4):454–460. doi: 10.1111/j.1365-3059.2004.01052.x.
  • Chartrain L, Brading PA, Widdowson JP, Brown JKM. 2004. Partial resistance to Septoria tritici blotch (Mycosphaerella graminicola) in wheat cultivars arina and riband. Phytopathol. 94(5):497–504. doi: 10.1094/PHYTO.2004.94.5.497.
  • Chen R, Boeger JM, McDonald BA. 1994. Genetic stability in a population of a plant pathogenic fungus over time. Mol Ecol. 3(3):209–218. doi: 10.1111/j.1365-294X.1994.tb00054.x.
  • Chin KM, Wolfe MS. 1984. The spread oi Erysiphe graminis f. sp. hordei in mixtures of barley varieties. Plant Pathol. 33(1):89–100. doi: 10.1111/j.1365-3059.1984.tb00592.x.
  • Civitello DJ, Cohen J, Fatima H, Halstead N, Liriano J, McMahon TA, Ortega CN, Sauer EL, Sehgal T, Young S, et al. 2015. Biodiversity inhibits parasites: broad evidence for the dilution effect. Proc Natl Acad Sci. 112(28):8667–8671. doi:10.1073/pnas.1506279112.
  • Clark WS. 2006. Septoria tritici and azole performance. Aspects Appl Biol. 78:127–132.
  • Cohen L, Eyal Z. 1993. The histology of processes associated with the infection of resistant and susceptible wheat cultivars with Septoria tritici. null. 42(5):737–743. doi: 10.1111/j.1365-3059.1993.tb01560.x.
  • Cook RJ, Hims MJ, Vaughan TB. 1999. Effects of fungicide spray timing on winter wheat disease control. null. 48(1):33–50. doi: 10.1046/j.1365-3059.1999.00319.x.
  • Cook DE, Mesarich CH, Thomma BP. 2015. Understanding plant immunity as a surveillance system to detect invasion. Annu Rev Phytopathol. 53(1):541–563. doi: 10.1146/annurev-phyto-080614-120114.
  • Cools HJ, Fraaije BA. 2008. Are azole fungicides losing ground against Septoria wheat disease? Resistance mechanisms in Mycosphaerella graminicola. Pest Manag Sci. 64(7):681–684. doi: 10.1002/ps.1568.
  • Cools HJ, Mullins JGL, Fraaije BA, Parker JE, Kelly DE, Lucas JA, Kelly SL. 2011. Impact of recently emerged sterol 14α-demethylase (CYP51) variants of Mycosphaerella graminicola on azole fungicide sensitivity. Appl Environmental Microbiol. 77(11):3830–3837. doi: 10.1128/AEM.00027-11.
  • Corsetti A, Gobbetti M, Rossi J, Damiani P. 1998. Antimould activity of sourdough lactic acid bacteria: identification of a mixture of organic acids produced by Lactobacillus sanfrancisco CB1. Appl Microbiol Biotechnol. 50(2):253–256. doi: 10.1007/s002530051285.
  • Cowger C, Hoffer ME, Mundt CC. 2000. Specific adaptation by Mycosphaerella graminicola to a resistant wheat cultivar. null. 49:445–451. doi:10.1046/j.1365-3059.2000.00472.x.
  • Cowger C, Mundt CC. 2002. Effects of wheat cultivar mixtures on epidemic progression of Septoria tritici blotch and pathogenicity of Mycosphaerella graminicola. Phytopathology. 92(6):617–623. doi: 10.1094/PHYTO.2002.92.6.617.
  • Crous PW. 1998. Mycosphaerella spp. and their anamorphs associated with leaf spot diseases of Eucalyptus. American Phytopathological Society. 21:1–170.
  • Cunfer BM, Ueng PP. 1999. Taxonomy and identification of Septoria and Stagonospora species on small-grain cereals. Annu Rev Phytopathol. 37(1):267. doi: 10.1146/annurev.phyto.37.1.267.
  • Czembor PC, Radecka-Janusik M, Mankowski D. 2011. Virulence spectrum of Mycosphaerella graminicola isolates on wheat genotypes carrying known resistance genes to septoria tritici blotch. J Phytopathol. 159:146–154. doi: 10.1111/j.1439-0434.2010.01734
  • Dalvand M, Soleimani PMJ, Zafari D. 2018. Evaluating the efficacy of STB resistance genes to Iranian Zymoseptoria tritici isolates. J Plant Dis Protect. 125(1):27–32. doi: 10.1007/s41348-017-0143-3.
  • Ditsch DC, Grove JH. 1991. Influence of tillage on plant populations, disease incidence, and grain yield of two soft red winter wheat cultivars. J Prod Agric. 4(3):360–365. doi: 10.2134/jpa1991.0360.
  • Downie RC, Lin M, Corsi B, Ficker A, Lillemo M, Oliver RP, Phan HTT, Tan KC, Cockram J. 2021. Septoria nodorum blotch of wheat: disease management and resistance breeding in the face of shifting disease dynamics and a changing environment. Phytopathol. 11(6):906–920. doi: 10.1094/PHYTO-07-20-0280-RVW.
  • Dreisigacker S, Wang X, Martinez CBA, Jing R, Singh PK. 2015. Adult-plant resistance to Septoria tritici blotch in hexaploid spring wheat. Theor Appl Genet. 128(11):2317–2329. doi: 10.1007/s00122-015-2587-9.
  • Duncan KE, Howard RJ. 2000. Cytological analysis of wheat infection by the leaf blotch pathogen Mycosphaerella graminicola. Mycol Res. 104(9):1074–1082. doi: 10.1017/S0953756299002294.
  • Dutilloy E, Oni FE, Esmaeel Q, Clément C, Barka EA. 2022. Plant beneficial bacteria as bioprotectants against wheat and barley diseases. J Fungi. 8(6):632. doi: 10.3390/jof8060632.
  • Eriksen L, Borum F, Jahoor A. 2003. Inheritance and localisation of resistance to Mycosphaerella graminicola causing Septoria tritici blotch and plant height in the wheat (Triticum aestivum L.) genome with DNA markers. Theor Appl Genet. 107(3):515–527. doi: 10.1007/s00122-003-1276-2.
  • Eyal Z. 1971. The kinetics of pycnospore liberation in Septoria tritici. Can J Bot. 49(7):1095–1099. doi: 10.1139/b71-157.
  • Eyal Z. 1981. Integrated control of Septoria diseases of wheat. Plant Dis. 65(9):763–768. doi: 10.1094/PD-65-763.
  • Eyal Z, Amiri Z, Wahl I. 1973. Physiologic specialization of Septoria tritici. Phytopathol. 49(9):1087–1091. doi: 10.1094/Phyto-63-1087.
  • Eyal Z, Levy E. 1987. Variations in pathogenicity patterns of Mycosphaerella graminicola within Triticum spp. in Israel. Euphytica. 36(1):237–250. doi: 10.1007/BF00730670.
  • Eyal Z, Scharen AL, Prescott JM. 1985. Global fingerprinting of Leptosphaeria nodorum (Septoria nodorum) and Mycosphaerella graminicola (Septoria tritici) pathogenicity patterns. ARS-US Department Of Agriculture, Agril Res Serv (USA). 37(12):74–76.
  • Eyal Z, Scharen AL, Prescott JM, Ginkel MV. 1987. The septoria diseases of wheat: concepts and methods of disease management (Mexico, D.F.: CIMMYT).
  • Fernandez MR, Stevenson CF, Hodge K, Dokken-Bouchard F, Pearse PG, Waelchli F, Brown A, Peluola C. 2016. Assessing effects of climatic change, region and agronomic practices on leaf spotting of bread and durum wheat in the western Canadian prairies, from 2001 to 2012. Agron J. 108:1180–1195. doi: 10.2134/agronj2015.0451.
  • Figueroa M, Hammond‐Kosack KE, Solomon PS. 2018. A review of wheat diseases—a field perspective. Mol Plant Pathol. 19(6):1523–1536. doi: 10.1111/mpp.12618.
  • Finckh M, Gacek E, Goyeau H, Lannou C, Merz U, Mundt C, Munk L, Nadziak J, Newton A, De Vallavieille-Pope C, et al. 2000. Cereal variety and species mixtures in practice, with emphasis on disease resistance. Agronomie. 20(7):813–837. doi: 10.1051/agro:2000177.
  • Finckh M, Wolfe M. 2006. Diversification strategies. In: Cooke B, Jones D, and Kate B, editors. The Epidemiology of Plant Diseases. Dordrecht: Springer; pp. 269–307. doi:10.1007/1-4020-4581-6_10.
  • Flor HH. 1942. Inheritance of pathogenicity in Melampsora lini. Phytopathol. 32:653–669.
  • Fones H, Gurr S. 2015. The impact of Septoria tritici blotch disease on wheat: an EU perspective. Fungal Genet Biol. 79:3–7. doi:10.1016/j.fgb.2015.04.004.
  • Fraaije BA, Bayon C, Atkins S, Cools HJ, Lucas JA, Fraaije MW. 2012. Risk assessment studies on succinate dehydrogenase inhibitors, the new weapons in the battle to control Septoria leaf blotch in wheat. Mol Plant Pathol. 13(3):263–275. doi: 10.1111/j.1364-3703.2011.00746.x.
  • Fraaije BA, Cools HJ, Fountaine J, Lovell DJ, Motteram J, West JS, Lucas JA. 2005. Role of ascospores in further spread of QoI-resistant cytochrome b alleles (G143A) in field populations of Mycosphaerella graminicola. Phytopathol. 95(8):933–941. doi: 10.1094/PHYTO-95-0933.
  • Francisco CS, Ma X, Zwyssig MM, McDonald BA, Palma-Guerrero J. 2019. Morphological changes in response to environmental stresses in the fungal plant pathogen Zymoseptoria tritici. Sci Rep. 9(1):9642. doi: 10.1038/s41598-019-45994-3.
  • Francisco CS, Zwyssig MM, Palma-Guerrero J. 2020. The role of vegetative cell fusions in the development and asexual reproduction of the wheat fungal pathogen Zymoseptoria tritici. BMC Biol. 18(1):1–16. doi: 10.1186/s12915-020-00838-9.
  • Gan Y, Gossen BD, Li L, Ford G, Banniza S. 2007. Cultivar type, plant population, and ascochyta blight in chickpea. Agron J. 99(6):1463–1470. doi: 10.2134/agronj2007.0105.
  • Garnault M, Duplaix C, Leroux P, Couleaud G, Carpentier F, David O, Walker A. 2019. Spatiotemporal dynamics of fungicide resistance in the wheat pathogen Zymoseptoria tritici in France. Pest Manag Sci. 75(7):1794–1807. doi: 10.1002/ps.5360.
  • Ghaffary MST 2011. Efficacy and mapping of resistance to Mycosphaerella graminicola in wheat [ Doctoral dissertation]. Wageningen University and Research.
  • Ghaffary SMT, Chawade A, Singh PK 2018. Practical breeding strategies to improve resistance to Septoria tritici blotch of wheat. Euphytica. 214:1–18. doi: 10.1007/s10681-018-2205-4
  • Ghaneie A, Mehrabi R, Safaie N, Abrinbana M, Saidi A, Aghaee M. 2012. Genetic variation for resistance to septoria tritici blotch in Iranian tetraploid wheat landraces. Eur J Plant Pathol. 132(2):191–202. doi: 10.1007/s10658-011-9862-7.
  • Gigot C, Saint‐Jean S, Huber L, Maumené C, Leconte M, Kerhornou B, De Vallavieille‐Pope C. 2013. Protective effects of a wheat cultivar mixture against splash‐dispersed Septoria tritici blotch epidemics. Plant Pathol. 62(5):1011–1019. doi: 10.1111/ppa.12012.
  • Gilbert J, Woods SM. 2001. Leaf spot diseases of spring wheat in southern Manitoba farm fields under conventional and conservation tillage. Can J Plant Sci. 81(3):551–559. doi: 10.4141/P00-088.
  • Gohari AM 2015. Identification and functional characterization of putative (a) virulence factors in the fungal wheat pathogen Zymoseptoria tritici. [ Doctoral dissertation]. Netherlands: Wageningen University and Research.
  • Goodwin SB. 2007. Back to basics and beyond: increasing the level of resistance to Septoria tritici blotch in wheat. Australasian Plant Pathol. 36(6):532–538. doi: 10.1071/AP07068.
  • Grieger A, Lamari L, Brûlé-Babel A. 2005. Physiologic variation in Mycosphaerella graminicola from western Canada. Can J Plant Pathol. 27(1):71–77. doi: 10.1080/07060660509507196.
  • Hailemariam BN, Kidane YG, Ayalew A. 2020. Epidemiological factors of Septoria tritici blotch (Zymoseptoria tritici) in durum wheat (Triticum turgidum) in the highlands of Wollo, Ethiopia. Ecol Process. 9(1):1–11. doi: 10.1186/s13717-020-00258-1.
  • Hillocks RJ. 2012. Farming with fewer pesticides: EU pesticide review and resulting challenges for UK agriculture. Crop Prot. 31:85–93. doi:10.1016/j.cropro.2011.08.008.
  • Hilton H, Jenkinson J, Hollins H, Parry P. 1999. Relationship between cultivar height and severity of Fusarium ear blight in wheat. Plant Pathol. 48(2):202–208. doi: 10.1046/j.1365-3059.1999.00339.x.
  • Hilu HB, Bever WM. 1957. Inoculation, over summering and suscept-pathogen relationships of Septoria tritici on Triticum species. Phytopathol. 72:474–480.
  • Hosseinnezhad A, Khodarahmi M, Rezaee S, Mehrabi R, Roohparvar R. 2014. Effectiveness determination of wheat genotypes and Stb resistance genes against Iranian Mycosphaerella graminicola isolates. Arch Phytopathol Plant Prot. 47(17):2051–2069. doi: 10.1080/03235408.2013.868696.
  • Howard DD, Chambers AY, Logan J. 1994. Nitrogen and fungicide effects on yield components and disease severity in wheat. J Pro Agr. 7(4):448–454. doi: 10.2134/jpa1994.0448.
  • Huber DM, Lee TS, Ross MA, Abney TS 1987. Amelioration of tan spot-infected wheat with nitrogen. Plant Dis. 71(1):49–50.
  • Hunter T, Coker RR, Royle DJ. 1999. The teleomorph stage, Mycosphaerella graminicola, in epidemics of Septoria tritici blotch on winter wheat in the UK. Plant Pathol. 48(1):51–57. doi: 10.1046/j.1365-3059.1999.00310.x.
  • Ilyas M, Hörger AC, Bozkurt TO, van den Burg HA, Kaschani F, Kaiser M, Belhaj K, Smoker M, Joosten MAJ, Kamoun S, et al. 2015. Functional divergence of two secreted immune proteases of tomato. Curr Biol. 25(17):2300–2306. doi: 10.1016/j.cub.2015.07.030.
  • Jackson LF, Dubcovsky J, Gallagher LW. 2000. Regional barley and common and durum wheat performance tests in California. Agron Pro Rep. 272:1–56.
  • Jashni MK, Mehrabi R, Collemare J, Mesarich CH, de Wit PJG. 2015. The battle in the apoplast: further insights into the roles of proteases and their inhibitors in plant–pathogen interactions. Front Plant Sci. 6:584. doi: 10.3389/fpls.2015.00584.
  • Jeger MJ, Griffiths E, Jones DG. 1981. Disease progress of non‐specialised fungal pathogens in intraspecific mixed stands of cereal cultivars. I Models Ann Appl Biol. 98(2):187–198. doi: 10.1111/j.1744-7348.1981.tb00752.x.
  • Jensen NF. 1952. Intra‐varietal diversification in oat breeding 1. Agron J. 44(1):30–34. doi: 10.2134/agronj1952.00021962004400010009x.
  • Johnson T. 1961. Man-guided evolution in plant rusts: Through his modification of the host plants of the cereal rusts, man is also modifying the rusts. Science. 133(3450):357–362. doi: 10.1126/science.133.3450.357.
  • Jones JD, Dangl JL. 2006. The plant immune system. Nature. 444(7117):323–329. doi: 10.1038/nature05286.
  • Juliana P, Singh RP, Singh PK, Poland JA, Bergstrom GC, Huerta-Espino J, Bhavani S, Crossa J, Sorrells ME. 2018. Genome-wide association mapping for resistance to leaf rust, stripe rust and tan spot in wheat reveals potential candidate genes. Theor Appl Genet. 131(7):1405–1422. doi: 10.1007/s00122-018-3086-6.
  • Jurke CJ, Fernando WGD. 2008. Effects of seeding rate and plant density on sclerotinia stem rot incidence in canola. Arch Phytopathol Plant Prot. 41(2):142–155. doi: 10.1080/03235400600679743.
  • Kanyuka K, Rudd JJ. 2019. Cell surface immune receptors: the guardians of the plant’s extracellular spaces. Curr Opin Plant Biol. 50:1–8. doi:10.1016/j.pbi.2019.02.005.
  • Kema GH, Mirzadi GA, Aouini L, Gibriel HA, Ware SB, Van DBF, Manning-Smith R, Alonso-Chavez V, Help J, Ben MS, et al. 2018. Stress and sexual reproduction affect the dynamics of the wheat pathogen effector AvrStb6 and strobilurin resistance. Nat Genet. 50(3):375–380. doi:10.1038/s41588-018-0052-9.
  • Kema GHJ, Vansilfhout C. 1997. Genetic variation for virulence and resistance in the wheat-Mycosphaerella graminicola pathosystem III. Comparative seedling and adult plant experiments. Phytopathol. 87(3):266–272. doi: 10.1094/PHYTO.1997.87.3.266.
  • Kema GHJ, Verstappen ECP, Waalwijk C. 2000. Avirulence in the wheat Septoria tritici leaf blotch fungus Mycosphaerella graminicola is controlled by a single locus. Mol Plant-Microbe Interact. 13(12):1375–1379. doi: 10.1094/MPMI.2000.13.12.1375.
  • Kema GHJ, Yu DZ, Rijkenberg FHJ, Shaw MW, Baayen RP. 1996. Histology of the pathogenesis of Mycosphaerella graminicola in wheat. Phytopathol. 86(7):777–786. doi: 10.1094/Phyto-86-777.
  • Kettles GJ, Bayon C, Canning G, Rudd JJ, Kanyuka K. 2017. Apoplastic recognition of multiple candidate effectors from the wheat pathogen Zymoseptoria tritici in the nonhost plant Nicotiana benthamiana. New Phytol. 213:338–350. doi:10.1111/nph.14215.
  • Kidane YG, Hailemariam BN, Mengistu DK, Fadda C, Pè ME, Dell’Acqua M. 2017. Genome-wide association study of Septoria tritici blotch resistance in Ethiopian durum wheat landraces. Front Plant Sci. 8:1–12. doi:10.3389/fpls.2017.01586.
  • Kildea S, Byrne JJ, Cucak M, Hutton F. 2020. First report of virulence to the Septoria tritici blotch resistance gene Stb16q in the Irish Zymoseptoria tritici population. New Dis Rep. 41(1):13. doi: 10.5197/j.2044-0588.2020.041.013.
  • Kildea S, Ransbotyn V, Khan MR, Fagan B, Leonard G, Mullins E, Doohan FM. 2008. Bacillus megaterium shows potential for the biocontrol of Septoria tritici blotch of wheat. Biol Control. 47(1):37–45. doi: 10.1016/j.biocontrol.2008.07.001.
  • Kildea S, Sheppard L, Cucak M, Hutton F. 2021. Detection of virulence to Septoria tritici blotch resistance conferred by the winter wheat cultivar cougar in the Irish Zymoseptoria tritici population and potential implications for LB control. Plant Pathol. 70(9):2137–2147. doi: 10.1111/ppa.13432.
  • King KC, Lively CM. 2012. Does genetic diversity limit disease spread in natural host populations? Heredity. 109(4):199–203. doi: 10.1038/hdy.2012.33.
  • Kombrink E, Schmelzer E. 2001. The hypersensitive response and its role in local and systemic disease resistance. Eur J Plant Pathol. 107:69–78. doi:10.1023/A:1008736629717.
  • Kristoffersen R, Jørgensen LN, Eriksen LB, Nielsen GC, Kiær LP. 2020. Control of Septoria tritici blotch by winter wheat cultivar mixtures: Meta-analysis of 19 years of cultivar trials. Field Crops Res. 249. doi: 10.1016/j.fcr.2019.107696.
  • Krupinsky JM 1999. Influence of cultural practices on Septoria/Stagonospora diseases. septoria and stagonospora diseases of cereals: A compilation of global research. Proceedings of International Workshop, 5. Mexico: DF Mexico. p. 20–24.
  • Krupinsky JM, Bailey KL, McMullen MP, Gossen BD, Turkington TK. 2002. Managing plant disease risk in diversified cropping systems. Agron J. 94(4):955. doi: 10.2134/agronj2002.955a.
  • Krupinsky JM, Halvorson AD, Tanaka DL, Merrill SD. 2007. Nitrogen and tillage effects on wheat leaf spot diseases in the northern great plains. Agron J. 99(2):562–569. doi: 10.2134/agronj2006.0263.
  • Krupinsky JM, Tanaka DL, Lares MT, Merrill SD. 2004. Leaf spot diseases of barley and spring wheat as influenced by preceding crops. Agron J. 96(1):259–266. doi: 10.2134/agronj2004.2590.
  • Leroux P, Walker A. 2011. Multiple mechanisms account for resistance to sterol 14α‐demethylation inhibitors in field isolates of Mycosphaerella graminicola. Pest Manage Sci. 67(1):44–59. doi: 10.1002/ps.2028.
  • Lillemo M, Jalli M, Ficke A, Djurle A, Jørgensen LNS, Latvala E, Teperi Happstadius I. 2011. Effect if wheat debris as source of primary inoculum on the early stages of Septoria leaf blotch epidemics. In: Occurrence of Leaf Blotch Pathogens on Wheat in the Nordic Countries and Potential Role of Toxin-Sensitivity Loci in Determining Susceptibility Und. International Symposium on Mycosphaerella and Stagonospora Diseases of Cereals, Mexico.
  • Linde CC, Zhan J, McDonald BA. 2002. Population structure of Mycosphaerella graminicola: from lesions to continents. Phytopathol. 92(9):946–955. doi: 10.1094/PHYTO.2002.92.9.946.
  • Liu Y, Zhang L, Thompson IA, Goodwin SB, Ohm HW 2013. Molecular mapping re-locates the Stb2 gene for resistance to septoria tritici blotch derived from cultivar veranopolis on wheat chromosome 1BS. Euphytica. 190:145–156. doi: 10.1007/s10681-012-0796-8
  • Lovell DJ, Hunter T, Powers SJ, Parker SR, Vanden BF. 2004. Effect of temperature on latent period of Septoria leaf blotch on winter wheat under outdoor conditions. Plant Pathol. 53(2):170–181. doi: 10.1111/j.0032-0862.2004.00983.x.
  • Lovell DJ, Parker SR, Hunter T, Welham SJ, Nichols AR. 2004. Position of inoculum in the canopy affects the risk of Septoria tritici blotch epidemics in winter wheat. Plant Pathol. 53(1):11–21. doi: 10.1046/j.1365-3059.2003.00939.x.
  • Lucas J. 1998. Plant pathology and plant pathogens Blackwell Sci. UK: Oxford.
  • Lynch KM, Zannini E, Guo J, Axel C, Arendt EK, Kildea S, Coffey A. 2016. Control of Zymoseptoria tritici cause of Septoria tritici blotch of wheat using antifungal Lactobacillus strains. J Appl Microbiol. 121(2):485–494. doi: 10.1111/jam.13171.
  • Mäe A, Fillinger S, Sooväli P, Heick TM, Robert G, Mcgrann D. 2020. Fungicide sensitivity shifting of Zymoseptoria tritici in the Finnish-Baltic region and a novel insertion in the MFS1 Promoter. Front Plant Sci. 11:1–10. doi:10.3389/fpls.2020.00385.
  • Makhdoomi A, Mehrabi R, Khodarahmi M, Abrinbana M. 2015. Efficacy of wheat genotypes and Lb resistance genes against Iranian isolates of Zymoseptoria tritici. J Gen Plant Pathol. 81(1):5–14. doi: 10.1007/s10327-014-0565-8.
  • McDermott JM, McDonald BA, Allard RW, Webster RK. 1989. Genetic variability for pathogenicity, isozyme, ribosomal DNA and colony color variants in populations of Rhynchosporium secalis. Genetics. 122(3):561–565. doi: 10.1093/genetics/122.3.561.
  • McDonald BA, Allard RW, Webster RK. 1988. Responses of two‐, three‐, and four‐component barley mixtures to a variable pathogen population. Crop Sci. 28(3):447–452. doi: 10.2135/cropsci1988.0011183X002800030003x.
  • McDonald BA, Mundt CC. 2016. How knowledge of pathogen population biology informs management of Septoria tritici blotch. Phytopathol. 106(9):948–955. doi: 10.1094/PHYTO-03-16-0131-RVW.
  • Medini M, Hamza S. 2008. Pathotype and molecular characterization of Mycosphaerella graminicola isolates collected from Tunisia, Algeria, and Canada. J Plant Pathol. 9(1):65–73.
  • Mehrabi R, Zwiers LH, de Waard MA, Kema GH. 2006. MgHog1 regulates dimorphism and pathogenicity in the fungal wheat pathogen Mycosphaerella graminicola. Mol Plant-Microbe Interact. 19(11):1262–1269. doi: 10.1094/MPMI-19-1262.
  • Mejri S, Siah A, Coutte F, Robert MM, Randoux B, Tisserant B, Krier F, Jacques P, Reignault P, Halama P, et al. 2018. Bio control of the wheat pathogen Zymoseptoria tritici using cyclic lipopeptides from Bacillus subtilis. Environ Sci Pollut Res. 25:29822–29833. doi: 10.1007/s11356-017-9241-9.
  • Mekonnen T, Haileselassie T, Abayo BG, Tesfaye K. 2020. Virulence variability of Ethiopian Zymoseptoria tritici isolates and efficacy of wheat genotypes and Lb resistance genes against the isolates. Eur J Plant Pathol. 158(4):895–910. doi: 10.1007/s10658-020-02125-3.
  • Mekonnen T, Sneller CH, Haileselassie T, Ziyomo C, Abeyo BG, Goodwin SB, Lule D, Tesfaye K. 2021. Genome-wide association study reveals novel genetic loci for quantitative resistance to Septoria tritici blotch in wheat (Triticum aestivum L.). Front Plant Sci. 12:1–20. doi:10.3389/fpls.2021.671323.
  • Melville SC, Jemmett JL. 1971. The effect of glume blotch on the yield of winter wheat. Plant Pathol. 20(1):14–17. doi: 10.1111/j.1365-3059.1971.tb00499.x.
  • Miedaner T, Risser P, Paillard S, Schnurbusch T, Keller B, Hartl L, Holzapfel J, Korzun V, Ebmeyer E, Utz HF, et al. 2012. Broad-spectrum resistance loci for three quantitatively inherited diseases in two winter wheat populations. Mol Breed. 29(3):731–742. doi:10.1007/s11032-011-9586-6.
  • Miedaner T, Zhao Y, Gowda M, Longin CFH, Korzun V, Ebmeyer E, Kazman E, Reif JC. 2013. Genetic architecture of resistance to Septoria tritici blotch in European wheat. BMC Genom. 14(1):1–8. doi: 10.1186/1471-2164-14-858.
  • Mirdita V, He S, Zhao Y, Korzun V, Bothe R, Ebmeyer E, Reif JC, Jiang Y. 2015. Potential and limits of whole genome prediction of resistance to Fusarium head blight and Septoria tritici blotch in a vast Central European elite winter wheat population. Theor Appl Genet. 128(12):2471–2481. doi: 10.1007/s00122-015-2602-1.
  • Moore JW, Herrera-Foessel S, Lan C, Schnippenkoetter W, Ayliffe M, Huerta-Espino J, Lillemo M, Viccars L, Milne R, Periyannan S. 2015. A recently evolved hexose transporter variant confers resistance to multiple pathogens in wheat. Nat Genet. 47(12):1494–1498. doi: 10.1038/ng.3439.
  • Motteram J, Lovegrove A, Pirie E, Marsh J, Devonshire J, van de Meene A, Hammond-Kosack K, Rudd JJ. 2011. Aberrant protein N-glycosylation impacts upon infection-related growth transitions of the haploid plant-pathogenic fungus Mycosphaerella graminicola. Mol Microbiol. 81(2):415–433. doi: 10.1111/j.1365-2958.2011.07701.x.
  • Mundt CC, Hayes PM, Schön CC. 1994. Influence of barley variety mixtures on severity of scald and net blotch and on yield. Plant Pathol. 43(2):356–361. doi: 10.1111/j.1365-3059.1994.tb02696.x.
  • Muqaddasi QH, Zhao Y, Rodemann B, Plieske J, Ganal MW, Röder MS. 2019. Genome‐wide association mapping and prediction of adult stage Septoria tritici blotch infection in European winter wheat via high‐density marker arrays. Plant Genome. 12(1):180029. doi: 10.3835/plantgenome2018.05.0029.
  • Nedyalkova S, Slavov S, Christova P, Stoyanova Z, Rodeva R 2019. Differentiation of the phytopathogenic fungi involved in septoria leaf spot complex by classical and molecular methods. In: Faculty of Biology Book 4 - Scientific Sessions of the Faculty of Biology. Sofia: International Scientific Conference “Kliment’s Days“, Sofia. Vol. 104. p. 79–89.
  • Newton AC, Begg GS, Swanston JS. 2009. Deployment of diversity for enhanced crop function. Ann Appl Biol. 154(3):309–322. doi: 10.1111/j.1744-7348.2008.00303.x.
  • Newton AC, Ellis RP, Hackett CA, Guy DC. 1997. The effect of component number on Rhynchosporium secalis infection and yield in mixtures of winter barley cultivars. Plant Pathol. 46(6):930–938. doi: 10.1046/j.1365-3059.1997.d01-83.x.
  • Newton AC, Thomas WTB. 1993. The interaction of either an effective or a defeated major gene with non‐specific resistance on mildew infection (Erysiphe graminis f. sp. hordei) and yield in mixtures of barley. J Phytopathol. 139(3):268–274. doi: 10.1111/j.1439-0434.1993.tb01426.x.
  • Odilbekov F, Armoniené R, Koc A, Svensson J, Chawade A. 2019. GWAS-assisted genomic prediction to predict resistance to Septoria tritici blotch in Nordic winter wheat at seedling stage. Front Genet. 10:1224. doi:10.3389/fgene.2019.01224.
  • Olivera PD, Rouse MN, Jin Y. 2018. Identification of new sources of resistance to wheat stem rust in Aegilops spp. in the tertiary genepool of wheat. Front Plant Sci. 9:1719. doi:10.3389/fpls.2018.01719.
  • Orellana-Torrejon C, Vidal T, Boixel AL, Gélisse S, Saint-Jean S, Suffert F. 2022. Annual dynamics of Zymoseptoria tritici populations in wheat cultivar mixtures: A compromise between the efficacy and durability of a recently broken-down resistance gene? Plant Pathol. 71:289–303. doi:10.1111/ppa.13458.
  • Orton ES, Deller S, Brown JKM. 2011. Mycosphaerella graminicola: from genomics to disease control. Mol Plant Pathol. 12(5):413–424. doi: 10.1111/j.1364-3703.2010.00688.x.
  • Ostfeld RS, Keesing F. 2012. Effects of host diversity on infectious disease. Annu Rev Ecol Evol Syst. 43(157):2012. doi: 10.1146/annurev-ecolsys-102710-145022.
  • Palma-Guerrero J, Torriani SF, Zala M, Carter D, Courbot M, Rudd JJ, McDonald BA, Croll D. 2016. Comparative transcriptomic analyses of Zymoseptoria tritici strains show complex lifestyle transitions and intraspecific variability in transcription profiles. Mol Plant Pathol. 17(6):845–859. doi: 10.1111/mpp.12333.
  • Palmer C, Skinner W. 2002. Mycosphaerella graminicola: latent infection, crop devastation and genomics. Mol Plant Pathol. 3(2):63–70. doi: 10.1046/j.1464-6722.2002.00100.x.
  • Pedersen EA. 1992. The effect of crop rotation on development of the septoria disease complex on spring wheat in Saskatchewan. Can J Plant Pathol. 14(2):152–158. doi: 10.1080/07060669209500892.
  • Perelló A, Mónaco C, Cordo C. 1997. Evaluation of Trichoderma harzianum and Gliocladium roseum in controlling leaf blotch of wheat (Septoria tritici) under in vitro and greenhouse conditions/Auswertung von Trichoderma harzianum und Gliocladium roseum bei der Bekämpfung der Blattfleckenkrank. J Plant Dis Protect. 104(6):588–598. https://www.jstor.org/stable/43386529.
  • Perello A, Simon MR, Arambarri AM. 2002. Interactions between foliar pathogens and the saprophytic microflora of the wheat (Triticum aestivum L.) phylloplane. J Phytopathol. 150(4‐5):232–243. doi: 10.1046/j.1439-0434.2002.00747.x.
  • Person C. 1966. Genetic polymorphism in parasitic systems. Nature. 212(5059):266–267. doi: 10.1038/212266a0.
  • Pfleeger TG, Mundt CC. 1998. Wheat leaf rust severity as affected by plant density and species proportion in simple communities of wheat and wild oats. Phytopathol. 88(7):708–714. doi: 10.1094/PHYTO.1998.88.7.708.
  • Pielaat A, Van DBF, Fitt BDL, Jeger MJ. 2002. Simulation of vertical spread of plant diseases in a crop canopy by stem extension and splash dispersal. Ecol Model. 151(2–3):195–212. doi: 10.1016/S0304-3800(01)00484-7.
  • Pietravalle S, Shaw MW, Parker SR, Van DBF. 2003. Modeling of relationships between weather and Septoria tritici epidemics on winter wheat: a critical approach. Phytopathol. 93(10):1329–1339. doi: 10.1094/PHYTO.2003.93.10.1329.
  • Plissonneau C, Hartmann FE, Croll D. 2018. Pangenome analyses of the wheat pathogen Zymoseptoria tritici reveal the structural basis of a highly plastic eukaryotic genome. BMC Biol. 16:5. doi: 10.1186/s12915-017-0457-4.
  • Ponomarenko A, Goodwin SB, Kema GHJ. 2011. Septoria tritici blotch (LB) of wheat. Plant Health Instr. doi: 10.1094/PHI-I-2011.
  • Quaedvlieg W, Kema GHJ, Groenewald JZ, Verkley GJM, Seifbarghi S, Razavi M, Mirzadi GA, Mehrabi R, Crous PW. 2011. Zymoseptoria gen. nov.: A new genus to accommodate Septoria-like species occurring on graminicolous hosts. Persoonia: Molecular Phylogeny And Evolution Of Fungi. 26:57–69. doi:10.3767/003158511X571841.
  • Quaedvlieg W, Verkley GJM, Shin H, Barreto RW, Alfenas AC, Swart WJ, Groenewald JZ, Crous PW. 2004. Sizing up Septoria. Studi Mycol. 75:307–390. doi:10.3114/sim0017.
  • Risser P, Ebmeyer E, Korzun V, Hartl L, Miedaner T. 2011. Quantitative trait loci for adult-plant resistance to Mycosphaerella graminicola in two winter wheat populations. Phytopathol. 101(10):1209–1216. doi: 10.1094/PHYTO-08-10-0203.
  • Rohel EA, Payne AC, Fraaije BA, Hollomon DW. 2001. Exploring infection of wheat and carbohydrate metabolism in Mycosphaerella graminicola transformants with differentially regulated green fluorescent protein expression. Mol Plant Microbe Interact. 14:156–163. doi:10.1094/MPMI.2001.14.2.156.
  • Rudd JJ, Kanyuka K, Hassani-Pak K, Derbyshire M, Andongabo A, Devonshire J, Lysenko A, Saqi M, Desai NM, Powers SJ. 2015. Transcriptome and metabolite profiling of the infection cycle of Zymoseptoria tritici on wheat reveals a biphasic interaction with plant immunity involving differential pathogen chromosomal contributions and a variation on the hemibiotrophic lifestyle definition. Plant Physiol. 167(3):1158–1185. doi: 10.1104/pp.114.255927.
  • Saindon G, Huang HC, Kozub GC. 1995. White-mold avoidance and agronomic attributes of upright common beans grown at multiple planting densities in narrow rows. J Am Soc Hort Sci. 120(5):843–847. doi: 10.21273/JASHS.120.5.843.
  • Saintenac C, Cambon F, Aouini L, Verstappen E, Ghaffary SMT, Poucet T, Marande W, Berges H, Xu S, Jaouannet M, et al. 2021. A wheat cysteine-rich receptor-like kinase confers broad-spectrum resistance against Septoria tritici blotch. Nature Commun. 12(1):433. doi:10.1038/s41467-020-20685-0.
  • Saintenac C, Lee WS, Cambon F, Rudd JJ, King RC, Marande W, Powers SJ, Bergès H, Phillips AL, Uauy C, et al. 2018. Wheat receptor-kinase-like protein Lb6 controls gene-for-gene resistance to fungal pathogen Zymoseptoria tritici. Nat Genet. 50(3):368–374. doi:10.1038/s41588-018-0051-x.
  • Salgado JD, Paul PA. 2016. Leaf blotch diseases of wheat—Septoria tritici blotch, stagonospora nodorum blotch and tan spot. Ohio State University Extention. https://Ohioline.Osu.Edu/Factsheet/Plpath-Cer-07.
  • Sánchez-Vallet A, Fouché S, Fudal I, Hartmann FE, Soyer JL, Tellier A, Croll D. 2018. The genome biology of effector gene evolution in filamentous plant pathogens. Annu Rev Phytopathol. 56(1):21–40. doi: 10.1146/annurev-phyto-080516-035303.
  • Sanderson FR, Scharen AL, Scott PR 1985. Sources and importance of primary infection and identities of associated propagules. In: Septoria of Cereals (Scharen AL, ed.). Proceedings of the Workshop, 2 – 4 August 1983. Bozeman, MT: Montana State University. p. 57–64.
  • Savary S, Willocquet L, Pethybridge SJ, Esker P, McRoberts N, Nelson A. 2019. The global burden of pathogens and pests on major food crops. Nat Ecol Evol. 3(3):430–439. doi: 10.1038/s41559-018-0793-y.
  • Scala V, Chiara P, Valentina F, Marzia B, Slaven Z, Fabrizio Q, Mauro F, Marco Z, Babak M, Massimo R, et al. 2020. Tramesan elicits durum wheat defense against the septoria disease complex. Biomolecules. 10(4):608. doi:10.3390/biom10040608.
  • Schellenberger R, Touchard M, Clément C, Baillieul F, Cordelier S, Crouzet J, Dorey S. 2019. Apoplastic invasion patterns triggering plant immunity: plasma membrane sensing at the frontline. Mol Plant Pathol. 20(11):1602–1616. doi: 10.1111/mpp.12857.
  • Schilly A, Risser P, Ebmeyer E, Hartl L, Reif JC, Würschum T, Miedaner T. 2011. Stability of adult‐plant resistance to Septoria tritici blotch in 24 European winter wheat varieties across nine field environments. J Phytopathol. 159(6):411–416. doi: 10.1111/j.1439-0434.2010.01783.x.
  • Schnürer J, Magnusson J. 2005. Antifungal lactic acid bacteria as biopreservatives. Trends Food Sci Technol. 16(1–3):70–78. doi: 10.1016/j.tifs.2004.02.014.
  • Schwartz HF, Steadman JR, Coyne DP. 1978. Influence of Phaseolus vulgaris blossoming characteristics and canopy structure upon reaction to Sclerotinia sclerotiorum. Phytopathol. 68(3):465–470. doi: 10.1094/Phyto-68-465.
  • Shaner G, Finney RE, Patterson FL. 1975. Expression and effectiveness of resistance in wheat to Septoria leaf blotch. Phytopathol. 65(7):761–766. doi: 10.1094/Phyto-65-761.
  • Shaw MW. 1990. Effects of temperature, leaf wetness and cultivar on the latent period of Mycosphaerella graminicola on winter wheat. Plant Pathol. 39(2):255–268. doi: 10.1111/j.1365-3059.1990.tb02501.x.
  • Shaw MW, Royle DJ. 1993. Factors determining the severity of epidemics of Mycosphaerella graminicola (Septoria tritici) on winter wheat in the UK. Plant Pathol. 42(6):882–899. doi: 10.1111/j.1365-3059.1993.tb02674.x.
  • Shetty NP, Kristensen BK, Newman MA, Møller K, Gregersen PL, Jørgensen HJL. 2003. Association of hydrogen peroxide with restriction of Septoria tritici in resistant wheat. Physiol And Mol Plant Pathol. 62(6):333–346. doi: 10.1016/S0885-5765(03)00079-1.
  • Simó NM, Ayala F, Cordo C, Röder M, Börner A. 2004. Molecular mapping of quantitative trait loci determining resistance to Septoria tritici blotch caused by Mycosphaerella graminicola in wheat. Euphytica. 138(1):41–48. doi: 10.1023/B:EUPH.0000047059.57839.98.
  • Singh RP, Singh PK, Rutkoski J, Hodson DP, He X, Jørgensen LN, Hovmøller MS, Huerta-Espino J. 2016. Disease impact on wheat yield potential and prospects of genetic control. Annu Rev Phytopathol. 54(1):303–322. doi: 10.1146/annurev-phyto-080615-095835.
  • Smithson JB, Lenne JM. 1996. Varietal mixtures: a viable strategy for sustainable productivity in subsistence agriculture. Annu Rev Phytopathol. 128(1):127–158. doi: 10.1111/j.1744-7348.1996.tb07096.x.
  • Somasco OA, Qualset CO, Gilchrist DG. 1996. Single-gene resistance to Septoria tritici blotch in the spring wheat cultivar ‘Tadinia’. Plant Breed. 115:261–267. doi:10.1111/j.1439-0523.1996.tb00914.x.
  • Stammler G, Semar M. 2011. Sensitivity of Mycosphaerella graminicola (anamorph: Septoria tritici) to DMI fungicides across Europe and impact on field performance. Bull OEPP/EPPO Bulletin. 41:149–155. doi:10.1111/j.1365-2338.2011.02454.x.
  • Steinberg G. 2015. Cell biology of Zymoseptoria tritici: Pathogen cell organization and wheat infection. Fungal Genet Biol. 79:17–23. doi:10.1016/j.fgb.2015.04.002.
  • Stephens C, Ölmez F, Blyth H, McDonald M, Bansal A, Turgay EB, Hahn F, Saintenac C, Nekrasov V, Solomon P. 2021. Remarkable recent changes in the genetic diversity of the avirulence gene AvrStb6 in global populations of the wheat pathogen Zymoseptoria tritici. Mol Plant Pathol. 22(9):1121–1133. doi: 10.1111/mpp.13101.
  • Stiles ME. 1996. Biopreservation by lactic acid bacteria. Antonie van Leeuwenhoek. BMC Microbiol. 70(2):331–345. doi: 10.1007/BF00395940.
  • Stiles J, Penkar S, Plocková M, Chumchalova J, Bullerman LB. 2002. Antifungal activity of sodium acetate and Lactobacillus rhamnosus. J Food Prot. 65(7):1188–1191. doi: 10.4315/0362-028X-65.7.1188.
  • Stotz HU, Mitrousia GK, de Wit PJ, Fitt BD. 2014. Effector-triggered defence against apoplastic fungal pathogens. Trends Plant Sci. 19(8):491–500. doi: 10.1016/j.tplants.2014.04.009.
  • Stukenbrock EH, Quaedvlieg W, Javan-Nikhah M, Zala M, Crous PW, McDonald BA. 2012. Zymoseptoria ardabiliae and Z. pseudotritici, two progenitor species of the septoria tritici leaf blotch fungus Z. tritici (synonym: Mycosphaerella graminicola). Mycologia. 104(6):1397–1407. doi:10.3852/11-374.
  • Suffert F, Sache I, Lannou C. 2011. Early stages of Septoria tritici blotch epidemics of winter wheat: build‐up, overseasoning, and release of primary inoculum. Plant Pathol. 60(2):166–177. doi: 10.1111/j.1365-3059.2010.02369.x.
  • Tabib Ghaffary SM, JD F, TL F, RGF V, Van der Lee TAJ, Robert O, Kema GHJ. 2012. New broad-spectrum resistance to Septoria tritici blotch derived from synthetic hexaploid wheat. Theor Appl Genet. 124(1):125–142. doi: 10.1007/s00122-011-1692-7.
  • Takele A, Lencho A, Kassa B, Getaneh W, Hailu E. 2015. Estimated yield loss assessment of bread wheat (Triticumaestivum L.) due to Septoria leaf blotch Septaria tritici (Roberge in Desmaz) on wheat in Holeta Agricultural Research Center, West Shewa, Ethiopia. Res Plant Sci. 3(3):61–67.
  • Teferi TA, Gebreslassie ZS. 2015. Occurrence and intensity of wheat Septoria tritici blotch and host response in Tigray, Ethiopia. Crop Prot. 68:67–71. doi:10.1016/j.cropro.2014.10.020.
  • Thomas MR, Cook RJ, King JE. 1989. Factors affecting development of Septoria tritici in winter wheat and its effect on yield. Plant Pathol. 38(2):246–257. doi: 10.1111/j.1365-3059.1989.tb02140.x.
  • Tidd H, Rudd JJ, Ray RV, Bryant R, Kanyuka K. 2023. A large bioassay identifies Stb resistance genes that provide broad resistance against Septoria tritici blotch disease in the UK. Front Plant Sci. 13:1070986. doi:10.3389/fpls.2022.1070986.
  • Tompkins DK, Fowler DB, Wright AT. 1993. Influence of agronomic practices on canopy microclimate and septoria development in no-till winter wheat produced in the Parkland region of Saskatchewan. Canadian J Plant Sci. 73(1):331–344. doi: 10.4141/cjps93-050.
  • Van den Berg F, Van den Bosch F, Paveley ND. 2013. Optimal fungicide application timings for disease control are also an effective anti-resistance strategy: A case study for Zymoseptoria tritici (Mycosphaerella graminicola) on wheat. Phytopathol. 103(12):1209–1219. doi: 10.1094/PHYTO-03-13-0061-R.
  • Wang S, Kirillova K, Lehto X. 2017. Travelers’ food experience sharing on social network sites. J Travel Tourism Mark. 34(5):680–693. doi: 10.1080/10548408.2016.1224751.
  • Wang Y, Wang Y. 2018. Trick or treat: microbial pathogens evolved apoplastic effectors modulating plant susceptibility to infection. Int Soc For Molecular Plant Microbe Interactions. 31(1):6–12. doi: 10.1094/MPMI-07-17-0177-FI.
  • Wegulo SN, Breathnach JA, Baenziger PS. 2009. Effect of growth stage on the relationship between tan spot and spot blotch severity and yield in winter wheat. Crop Prot. 28(8):696–702. doi: 10.1016/j.cropro.2009.04.003.
  • Wegulo S, Stevens J, Zwingman M, Baenziger PS. 2012. Yield response to foliar fungicide application in winter wheat. In: Dhanasekaran D, editor. Fungicides for plant and animal diseases. Croatia: InTech Europe. 227e244. doi:10.5772/25716.
  • Wirthmueller L, Maqbool A, Banfield MJ. 2013. On the front line: structural insights into plant–pathogen interactions. Nat Rev Microbiol. 11(11):761–776. doi: 10.1038/nrmicro3118.
  • Wolfe MS. 1985. The current status and prospects of multiline cultivars and variety mixtures for disease resistance. Annu Rev Phytopathol. 23(1):251–273. doi: 10.1146/annurev.py.23.090185.001343.
  • Wolfe MS, Finckh MR. 1997. Diversity of host resistance within the crop: effects on host, pathogen and disease. In: Hartleb H, Heitefuss R Hoppe H, editors. Plant Resistance to Fungal Diseases. Jena, Germany: Fischer Verlag; p. 378–400.
  • Wyczling D, Lenc L, Sadowski C. 2010. Comparison of disease occurrence and green leaf area (GLA) of winter wheat depending on the forecrop and differentiated fungicidal protection used. J Plant Prot Res. 50(4):489–495. doi: 10.2478/v10045-010-0081-6.
  • Yang N, McDonald MC, Solomon PS, Milgate AW. 2018. Genetic mapping of Stb19, a new resistance gene to Zymoseptoria tritici in wheat. Theor Appl Genet. 131:2765–2773. doi:10.1007/s00122-018-3189-0.
  • Zhan J, Kema GHJ, Waalwijk C, McDonald BA. 2002. Distribution of mating type alleles in the wheat pathogen Mycosphaerella graminicola over spatial scales from lesions to continents. Fungal Genet Biol. 36(2):128–136. doi: 10.1016/S1087-1845(02)00013-0.
  • Zhan J, Linde CC, Jürgens T, Merz U, Steinebrunner F, McDonald BA. 2005. Variation for neutral markers is correlated with variation for quantitative traits in the plant pathogenic fungus Mycosphaerella graminicola. Mol Ecol. 14(9):2683–2693. doi: 10.1111/j.1365-294X.2005.02638.x.
  • Zhan J, Pettway RE, McDonald BA. 2003. The global genetic structure of the wheat pathogen Mycosphaerella graminicola is characterized by high nuclear diversity, low mitochondrial diversity, regular recombination, and gene flow. Fungal Genet Biol. 38(3):286–297. doi: 10.1016/S1087-1845(02)00538-8.
  • Zhong Z, Marcel TC, Hartmann FF, Ma X, Plissonneau C, Zala M, Ducasse A, Confais J, Compain J, Lapalu N, et al. 2017. A small secreted protein in Zymoseptoria tritici is responsible for avirulence on wheat cultivars carrying the Stb6 resistance gene. New Phytol. 214(2):619–631. doi:10.1111/nph.14434.
  • Zhou B, Benbow HR, Brennan CJ, Arunachalam C, Karki SJ, Mullins E, Feechan A, Burke JI, Doohan FM. 2020. Wheat encodes small, secreted proteins that contribute to resistance to Septoria tritici blotch. Front Genet. 11:469. doi:10.3389/fgene.2020.00469.