Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Unlabelled

Stripe (yellow) rust, caused by Puccinia striiformis f. sp. tritici (Pst), is a serious disease of wheat occurring in most wheat areas with cool and moist weather conditions during the growing season. The basidiomycete fungus is an obligate biotrophic parasite that is difficult to culture on artificial media. Pst is a macrocyclic, heteroecious fungus that requires both primary (wheat or grasses) and alternate (Berberis or Mahonia spp.) host plants to complete its life cycle. Urediniospores have the capacity for wind dispersal over long distances, which may, under high inoculum pressure, extend to thousands of kilometres from the initial infection sites. Stripe rust, which is considered to be the current major rust disease affecting winter cereal production across the world, has been studied intensively for over a century. This review summarizes the current knowledge of the Pst-wheat pathosystem, with emphasis on the life cycle, uredinial infection process, population biology of the pathogen, genes for stripe rust resistance in wheat and molecular perspectives of wheat-Pst interactions.

Taxonomy

The stripe rust pathogen, Puccinia striiformis Westend. (Ps), is classified in kingdom Fungi, phylum Basidiomycota, class Urediniomycetes, order Uredinales, family Pucciniaceae, genus Puccinia. Ps is separated below the species level by host specialization on various grass genera, comprising up to nine formae speciales, of which P. striiformis f. sp. tritici Erikss. (Pst) causes stripe (or yellow) rust on wheat.

Host range

Uredinial/telial hosts: Pst mainly infects common wheat (Triticum aestivum L.), durum wheat (T. turgidum var. durum L.), cultivated emmer wheat (T. dicoccum Schrank), wild emmer wheat (T. dicoccoides Korn) and triticale (Triticosecale). Pst can infect certain cultivated barleys (Hordeum vulgare L.) and rye (Secale cereale L.), but generally does not cause severe epidemics. In addition, Pst may infect naturalized and improved pasture grass species, such as Elymus canadensis L., Leymus secalinus Hochst, Agropyron spp. Garetn, Hordeum spp. L., Phalaris spp. L and Bromus unioloides Kunth. Pycnial/aecial (alternative) hosts: Barberry (Berberis chinensis, B. koreana, B. holstii, B. vulgaris, B. shensiana, B. potaninii, B. dolichobotrys, B. heteropoda, etc.) and Oregon grape (Mahonia aquifolium).

Disease symptoms

Stripe rust appears as a mass of yellow to orange urediniospores erupting from pustules arranged in long, narrow stripes on leaves (usually between veins), leaf sheaths, glumes and awns on susceptible plants. Resistant wheat cultivars are characterized by various infection types from no visual symptoms to small hypersensitive flecks to uredinia surrounded by chlorosis or necrosis with restricted urediniospore production. On seedlings, uredinia produced by the infection of a single urediniospore are not confined by leaf veins, but progressively emerge from the infection site in all directions, potentially covering the entire leaf surface. Individual uredinial pustules are oblong, 0.4-0.7 mm in length and 0.1 mm in width. Urediniospores are broadly ellipsoidal to broadly obovoid, (16-)18-30(-32) × (15-)17-27(-28) μm, with a mean of 24.5 × 21.6 μm, yellow to orange in colour, echinulate, and with 6-18 scattered germ pores. Urediniospores can germinate rapidly when free moisture (rain or dew) occurs on leaf surfaces and when the temperatures range is between 7 and 12 °C. At higher temperatures or during the later growing stages of the host, black telia are often produced, which are pulvinate to oblong, 0.2-0.7 mm in length and 0.1 mm in width. The teliospores are predominantly two-celled, dark brown with thick walls, mostly oblong-clavate, (24-)31-56(-65) × (11-)14-25(-29) μm in length and width, and rounded or flattened at the apex.

Free full text 


Logo of molplantpatholLink to Publisher's site
Mol Plant Pathol. 2014 Jun; 15(5): 433–446.
Published online 2014 May 6. https://doi.org/10.1111/mpp.12116
PMCID: PMC6638732
PMID: 24373199

Wheat stripe (yellow) rust caused by Puccinia striiformis f. sp. tritici

Summary

Stripe (yellow) rust, caused by Puccinia striiformis f. sp. tritici (Pst), is a serious disease of wheat occurring in most wheat areas with cool and moist weather conditions during the growing season. The basidiomycete fungus is an obligate biotrophic parasite that is difficult to culture on artificial media. Pst is a macrocyclic, heteroecious fungus that requires both primary (wheat or grasses) and alternate (Berberis or Mahonia spp.) host plants to complete its life cycle. Urediniospores have the capacity for wind dispersal over long distances, which may, under high inoculum pressure, extend to thousands of kilometres from the initial infection sites. Stripe rust, which is considered to be the current major rust disease affecting winter cereal production across the world, has been studied intensively for over a century. This review summarizes the current knowledge of the Pst–wheat pathosystem, with emphasis on the life cycle, uredinial infection process, population biology of the pathogen, genes for stripe rust resistance in wheat and molecular perspectives of wheat–Pst interactions.

Taxonomy

The stripe rust pathogen, Puccinia striiformisWestend. (Ps), is classified in kingdom Fungi, phylum Basidiomycota, class Urediniomycetes, order Uredinales, family Pucciniaceae, genus Puccinia. Ps is separated below the species level by host specialization on various grass genera, comprising up to nine formae speciales, of which P. striiformis f. sp. triticiErikss. (Pst) causes stripe (or yellow) rust on wheat.

Host range

Uredinial/telial hosts: Pst mainly infects common wheat (Triticum aestivumL.), durum wheat (T. turgidum var. durumL.), cultivated emmer wheat (T. dicoccumSchrank), wild emmer wheat (T. dicoccoidesKorn) and triticale (Triticosecale). Pst can infect certain cultivated barleys (Hordeum vulgareL.) and rye (Secale cerealeL.), but generally does not cause severe epidemics. In addition, Pst may infect naturalized and improved pasture grass species, such as Elymus canadensisL., Leymus secalinusHochst, Agropyron spp. Garetn, Hordeum spp. L., Phalaris spp. L and Bromus unioloidesKunth. Pycnial/aecial (alternative) hosts: Barberry (Berberis chinensis, B. koreana, B. holstii, B. vulgaris, B. shensiana, B. potaninii, B. dolichobotrys, B. heteropoda, etc.) and Oregon grape (Mahonia aquifolium).

Disease symptoms

Stripe rust appears as a mass of yellow to orange urediniospores erupting from pustules arranged in long, narrow stripes on leaves (usually between veins), leaf sheaths, glumes and awns on susceptible plants. Resistant wheat cultivars are characterized by various infection types from no visual symptoms to small hypersensitive flecks to uredinia surrounded by chlorosis or necrosis with restricted urediniospore production. On seedlings, uredinia produced by the infection of a single urediniospore are not confined by leaf veins, but progressively emerge from the infection site in all directions, potentially covering the entire leaf surface. Individual uredinial pustules are oblong, 0.4–0.7 mm in length and 0.1 mm in width. Urediniospores are broadly ellipsoidal to broadly obovoid, (16–)18–30(–32) × (15–)17–27(–28) μm, with a mean of 24.5 × 21.6 μm, yellow to orange in colour, echinulate, and with 6–18 scattered germ pores. Urediniospores can germinate rapidly when free moisture (rain or dew) occurs on leaf surfaces and when the temperatures range is between 7 and 12 °C. At higher temperatures or during the later growing stages of the host, black telia are often produced, which are pulvinate to oblong, 0.2–0.7 mm in length and 0.1 mm in width. The teliospores are predominantly two‐celled, dark brown with thick walls, mostly oblong‐clavate, (24–)31–56(–65) × (11–)14–25(–29) μm in length and width, and rounded or flattened at the apex.

Keywords: genes for resistance, infection procedure, life cycle, molecular aspect, population biology

Introduction

Stripe (yellow) rust, caused by Puccinia striiformis Westend. (Ps), is one of the most widely destructive plant diseases in modern winter cereal production (Wellings, 2011). The specialized form infecting wheat is referred to as P. striiformis f. sp. tritici (Pst). The sexual stage on Berberis spp. and Mahonia spp. has only been described recently (Jin et al., 2010; Wang and Chen, 2013; Zhao et al., 2011, 2013). Mutation, somatic recombination, parasexuality, selection and probably sexual recombination are considered to be the mechanisms that drive pathogen variability (Duan et al., 2010a; Hovmøller et al., 2011; Mboup et al., 2009; Stubbs, 1985). The centre of origin for P. striiformis was earlier assumed to be Transcaucasia, where grasses were the primary host (Hassebrauk, 1965), and from there the pathogen dispersed in all directions, including eastward to East Asia and North America in the 1910s (Carleton, 1915; Fraser and Conners, 1925; Humphrey et al., 1924), and southward into West Asia and East Africa. Geographically isolated wheat‐producing regions were invaded by long‐distance pathogen migration: Australia from Europe in 1979 (O'Brien et al., 1980); Republic of South Africa from the Middle East in 1996 (Boshoff et al., 2002); Western Australia from North America in 2002 (Wellings, 2007). Western China and Central Asia are also speculated as a centre of origin based on a high degree of telial production for isolates sampled in those areas and a high genetic diversity consistent with frequent recombination signatures (Ali et al., 2010b; Mboup et al., 2009).

Stripe rust is considered to be a low‐temperature disease and frequently occurs in temperate areas with cool and moist weather conditions. Recent devastating epidemics have occurred in warmer areas where the disease was previously infrequent or absent (Hovmøller et al., 2010; Mboup et al., 2009). This led to the proposal that populations of Pst had developed adaptation to higher temperature, and supporting evidence was published by Milus et al. (2008). The minimum, optimum and maximum temperatures for urediniospore germination are 0 °C, 7–12 °C and 20–26 °C, respectively (Schröder and Hassebrauk, 1964). Of the three wheat rusts (stripe rust, leaf rust and stem rust), stripe rust appears to be the most sensitive to environmental factors, such as air pollution and UV light, which reduce the germination of urediniospores (Sharp, 1967). Resistance of the host is also influenced by temperature and light, which, in turn, influences disease assessment of the infected plant (Sharp and Volin, 1970). Increasing day length or light intensity lowers the infection type (Bever, 1934; Stubbs, 1967; Wellings et al., 1988).

Yield losses in wheat from Pst infections are usually the result of reduced kernel number per spike, low test weight and reduced kernel quality (Prescott et al., 1986). Evidence from historical epidemics and crop losses, and a contemporary analysis of the incidence and crop loss magnitude experienced in current major world wheat‐growing regions, concluded that Pst was the most serious biotic threat to sustainable international wheat production (Wellings, 2011).

In this review, we summarize the current understanding of Pst biology, including studies of host–pathogen interactions, to develop strategies for resistance breeding which may form the basis for the development of practical control strategies in this economically significant pathosystem.

Life Cycle

In the asexual urediniospore stage, Ps infects a range of grasses within the Pooideae subfamily of the Poaceae. Collections of Ps from certain grasses showed evidence of host specialization, and this led Eriksson (1894) to propose five special forms (Latin: ‘formae speciales’) based on the originating host genus, namely P. striiformis f. sp. tritici (Pst) specialized on wheat, P. striiformis f. sp. hordei (Psh) on barley, P. striiformis f. sp. secalis on rye, P. striiformis f. sp. elymi on Elymus spp. and P. striiformis f. sp. agropyri on Agropyron spp. Additional formae speciales were reported: P. striiformis f. sp. dactylidis (Psd) on orchard grass (Dactylis glomerata L.) (Manners, 1960; Tollenaar, 1967), P. striiformis f. sp. poae (Psp) on Kentucky blue grass (Poa pratensis L.) (Britton and Cummins, 1956; Tollenaar, 1967), P. striiformis f. sp. leymi on Leymus secalinus (Georgi) Tzvel. (Niu et al., 1991) and P. striiformis f. sp. pseudo‐hordei (Psp‐h) on Hordeum spp. in Australia (Wellings, 2007).

Liu and Hambleton (2010) presented molecular [internal transcribed spacer (ITS) and β‐tubulin sequences] and morphological data from a set of 31 isolates to redefine several issues in the biology and taxonomy of Ps. Puccinia series striiformis was proposed to include P. striiformis sensu stricto (host range Aegilops, Elymus, Hordeum, Triticum), P. striiformoides (formerly Psd) infecting D. glomerata, P. pseudostriiformis (formerly Psp) infecting Poa spp. and P. gansensis in a single isolate from Achnatherum inebrians in China. The descriptions of the last three species agree well with the biology of these collections, including very close host affinity and distinct temperature optima for infection. However, the major formae speciales within Ps sensu stricto (i.e. Pst, Psh, Psp‐h) will remain as important taxa that assist in understanding the host biology and comparative potential threat of these forms in commercial agriculture.

Ps was long regarded as an autoecious microcyclic rust pathogen (urediniospores and teliospores forming on grass hosts) until Jin et al. (2010) identified Berberis spp. (B. chinensis, B. holstii, B. koreana, B. vulgaris) as alternative hosts supporting pycniospores and aeciospores of Pst and P. pseudostriiformis. Wang and Chen (2013) demonstrated that Oregon grape (Mahonia aquifolium) is also a host for the sexual phase of Pst.

Ps and P. pseudostriiformis are now classified as heteroecious macrocyclic rust pathogens (Fig. 1). Urediniospores are primarily dikaryotic (n + n) and maintain the dominant asexual stage of the pathogen population on the primary hosts. This phase is responsible for wide‐scale stripe rust epidemics reported on cereal crops.

figure

Life cycle of Puccinia striiformis. (A) Uredinia on wheat leaf containing single‐celled dikaryotic urediniospores (n + n) originating from aeciospores (n + n) or urediniospores. Top inset: echinulate surface of a urediniospore under a scanning electron microscope (SEM) (×4000). Bottom inset: broadly obovoid urediniospores (×1000). (B) Telia typically form beneath the leaf epidermis near the end of the growing season. Top inset: the two‐celled, oblong‐clavate teliospores (2n) (×1000). Bottom inset: the elliptoid basidiospores (n) from the germination of teliospores (×2500). (C) Pycnia produced by basidiospore infection on Berberis chinensis on upper leaf surfaces via inoculation with germinating teliospores of P. striiformis. Top inset: a magnified flask‐shaped pycnia (×400). Middle inset: the oblong‐shaped pycniospores (×4000). Bottom inset: magnified receptive hyphae (×900). (D) Cluster of sunflower‐shaped aecia produced on the lower leaf surface of Berberis shensiana. Top inset: a campanulate aecium (×200). Middle inset: flat spherical‐shaped aeciospores (×3300). Bottom inset: cluster of aeciospores (×250). (E) A wheat seedling that can be infected by aeciospores produced on barberry plants and can produce urediniospores.

As temperatures rise late in the epidemic phase, Pst typically produces thick‐walled, predominantly two‐celled teliospores. Each cell of a mature teliospore contains a diploid (2n) nucleus formed by karyogamy. Pst isolates vary in their ability to produce telia even under similar environmental conditions (Chen XM et al., 2012). Unlike those of P. graminis f. sp. tritici, teliospores of Pst do not have dormancy and readily germinate. Wang and Chen (unpublished Washington State University) have determined that, under free water conditions at 12 °C, teliospores germinate (24 h) to produce a promycelium of four cells. Meiosis then produces a single haploid nucleus that forms a basidiospore (48 h) ready to be ejected from the sterigma (60 h). Basidiospores germinate and infect Berberis spp. or Mahonia spp. (72 h). The authors also determined that a minimum 40‐h dew period was required for the infection of barberry leaves from basidiospores, with the highest infection achieved by keeping inoculated barberry plants in a dew chamber at 10 °C for 93 h. The absence of dormancy in teliospores and the consequent short period of available basidiospores led Rapilly (1979) to conclude that a sexual host, should it ever be discovered, would probably have a small role in pathogen survival. In view of this comment and the prolonged dew periods required for infection, it is perhaps unsurprising that there has been limited confirmation of Pst on Berberis spp. in nature, with just two reports from China (Zhao et al., 2011, 2013). It has been concluded that Berberis spp. do not play a role in stripe rust epidemics under the natural conditions in the US Pacific Northwest (Chen XM et al., 2012).

Uredinial Infection Process

Almost all biochemical and molecular studies have been based on urediniospores and their derived infection structures (Hahn, 2000; Hahn et al., 1997; Ling et al., 2007; Mendgen et al., 2000; Struck et al., 2004; Voegele, 2006; Voegele and Mendgen, 2003; Yin et al., 2009; Zhang et al., 2008). As an obligate biotrophic plant parasite, Pst infects the main hosts (cereal crops, grasses) from urediniospore deposition by wind or raindrops onto the leaf surface. Germination initiates within 3 h of contact with free moisture at a range of temperatures. The cytoplasm of a urediniospore moves into the growing germ tube as it orientates perpendicular to the long axis of epidermal cells during its growth across the leaf surface (Kang, 1996; Kang et al., 1997, 2002) until it reaches a stoma (Fig. 2C,D) (Moldenhauer et al., 2006; Wang et al., 2009). Within 6–8 h post‐inoculation (hpi), an appressorium is formed on the stoma, and at 8–12 hpi, a substomatal vesicle is formed within the stomatal cavity, the penetration hyphae are delimited by a septum (Kapooria and Mendgen, 1985) and the primary infection hypha emerges at 12–18 hpi. On contacting the mesophyll cell, a haustorial mother cell containing two to six nuclei develops (Kang et al., 1994, 2002) and most of the cytoplasm then moves into the haustorial mother cell, leaving earlier structures to become more or less vacuolated. Haustorial mother cells have a thick, multilayered wall that attaches firmly to the host cell wall. A slender neck forms from the haustorial mother cell, which then invaginates the host cell plasma membrane (Heath and Skalamera, 1997) to form a balloon‐shaped feeding structure, known as the haustorium (Fig. 2E; Hovmøller et al., 2011; Kang et al., 1997, 2003). Haustoria draw nutrients from host cells (Hahn and Mendgen, 2001; Mendgen, 1981; Staples, 2001; Voegele and Mendgen, 2003) and have also been shown to be involved in vitamin synthesis (Sohn et al., 2000). Haustoria are typically located in host mesophyll cells and up to 15% of epidermal cells (Hovmøller et al., 2011). Unlike stem rust and leaf rust fungi, multinucleate conditions in hyphae, haustorial mother cells and haustoria are very common (Kang et al., 1994). From 48 to 120 hpi, the primary infection hypha will give rise to many branched hyphae that develop between host mesophyll cells and produce multiple haustorial mother cells and haustoria, resulting in a branching network of fungal mycelium developing inter‐ and intracellularly within the host tissue. As the mycelium develops, a pustule bed becomes established from which a uredinium develops. Symptoms of chlorosis can be observed from 6 to 8 days after infection, whereas sporulation (uredinia appearing on the surface of leaves, leaf sheaths, glumes or awns) commences approximately from 12 to 14 days under favourable conditions. Symptoms of wheat stripe rust in the field and a range of seedling infection types in glasshouse tests are shown in Figs 3 and and4,4, respectively.

figure

Uredinial infection process. (A) Schematic representation of early infection structures by Puccinia striiformis (Ps). (B) Urediniospore under a scanning electron microscope (SEM) (×4000). (C) A urediniospore germ tube (GT) enters the leaf through a stoma (ST). (D) A substomatal vesicle (SSV) is formed, from which an infection hypha (IH) emerges. On contacting with a mesophyll cell, a haustorial mother cell (HMC) is differentiated. (E) A haustorium develops from the haustorial mother cell with a slender neck (Kang, 1996). (F) Ps hyphae spread and form colonies in infected leaf tissue. (G) A uredinium under a SEM (Kang et al., 1997). AP, appressorium; EHM, extrahaustorial matrix; GT, germ tube; HB, haustorial body; HMC, haustorial mother cell; HN, haustorial neck; IH, infection hypha; PP, penetration peg; SSV, substomatal vesicle; U, urediniospore.

figure

Wheat stripe rust symptoms in the field. (A–C) Yellow to orange uredinial pustules on susceptible adult plant leaves. (D) Uredinial pustules on the glumes and awns. (E, F) Yellow to orange uredinial pustules on the seedling wheat leaves. (G) Telial and uredinial pustules together on an adult plant leaf.

figure

Range of stripe rust seedling infection types. From the left, leaves show infection types ranging from immune to completely susceptible reaction (McIntosh et al., 1995).

Pst is noted for enhanced sensitivity to environmental conditions in comparison with other cereal rust pathogens. The most important factors are temperature, moisture, light and air pollution. Temperature is critical for successful germination and infection. In the presence of adequate moisture, in vitro germination studies showed a temperature range of 2–15 °C with an optimum at 7 °C (Sharp, 1965). Similar conclusions were drawn by Rapilly (1979) and Zhang et al. (2008). Newton and Johnson (1936) determined that the optimal temperature range for the latent period (time from infection to the beginning of sporulation) was 13–16 °C. These temperatures are, on average, about 10–15 °C lower in each category than those for P. triticina and P. graminis.

Moisture affects spore deposition, germination, infection and survival. Rapilly and Foucault (1976) observed that the adhesive force between urediniospore and receptive surfaces was significantly greater with high relative humidity, thus increasing the efficiency of spore attachment. Urediniospores require at least 3 h of available moisture on plant surfaces to germinate and infect (Hermansen and Veterinary, 1968; Tu and Hendrix, 1967). Any period of desiccation will irreversibly abort urediniospore germination (Vallavieille‐Pope et al., 1995). Zadoks (1961) and Rapilly and Fournet (1968) concluded that relative humidity must exceed 50% for sporulation to occur, and that urediniospore production increased exponentially with rising relative humidity.

Light intensity influences the host–pathogen interaction (Stubbs, 1967). Low light intensities in the first 4 days post‐infection caused infection type to increase and effectively mask a resistant response (Wellings et al., 1988). Seedling tests should always be conducted at light intensities above 10 000 lx in order to ensure consistent infection types (Stubbs, 1985). Ps has been shown to be more sensitive to air pollution, relative to other cereal rust pathogens, during urediniospore germination (Sharp, 1967). Later work showed that isolates from various regions varied in sensitivity to air pollutants (Melching et al., 1974; Stubbs, 1985).

Population biology of P. striiformis

Like the other cereal rust fungi, Pst has nearly a century of research history investigating detailed aspects of host–pathogen specialization that was fundamental to devising strategies to control the disease in commercial cropping.

Pathogen survey

Pathogenicity surveys for Pst have been historically based on the designation of avirulence or virulence responses of isolates inoculated on host cultivars or genotypes. The assignment of avirulence/virulence in seedling tests has been based on two infection type scales. The 0–9 scale (McNeal et al., 1971) is used by scientists in Europe, North America, Syria and Lebanon (Chen and Line, 1993; Hovmøller and Justesen, 2007a; Line et al., 1970), whereas the 0–4 scale (Gassner and Straib, 1932a, b) is applied in Australia (Wellings and McIntosh, 1990), China (Chen et al., 2009), India (Prashar et al., 2007), South Africa (Boshoff et al., 2002) and Pakistan (Ali et al., 2010a). Designations of avirulence/virulence with respect to specific host resistance genes must be based on the observed deviation from the expected low infection type of that gene, and not on nominal values set within the scale (McIntosh et al., 1995).

The differentiation of isolates into distinct races (or pathotypes) requires a set of host materials and prescribed testing conditions, and this was originally proposed for Ps by Gassner and Straib (1930, 1932a), and later revised by Fuchs (1965). In this approach, isolates with distinct and unique combinations of virulence and avirulence were described as races and given sequential numbers in order of first detection. Similar approaches continue to be used in North America and China. In North America, several sets of differentials have been developed from the original set of 13 proposed by Line et al. (1970) to more than 20 over time (Chen et al., 2002; Line and Qayoum, 1992). Races were given sequential CDL (later changed to PST) numbers, and this has been revised to PSTv designations with a new set of 18 wheat Yr single‐gene lines since 2010 (Wan and Chen, 2012).

In China, pathogenicity surveys of Ps have been conducted since the 1940s (Fang, 1944) and in some cases based on the earlier German differentials (Lu et al., 1956). In the intervening period, differential sets have changed or been modified, together with changes in commercial wheat cultivars, but the race nomenclature system has remained relatively consistent since the 1960s. The current differential set for the identification of races of Pst consists of 20 wheat lines (Chen et al., 2009). The major virulence patterns have been formally designated ‘CYR’ races (Chinese Yellow Rust) with sequential numbers based on their chronological identification. Virulence patterns with low frequencies and limited distribution were temporarily nominated as ‘pathotypes’ using the abbreviations of specific wheat differential genotypes. There have been 33 CYR races and 35 pathotypes of Pst described in China using the current set of wheat differentials (Chen et al., 2009; Liu et al., 2010; Wan et al., 2004).

A substantial revision of Ps pathotype nomenclature was proposed by Johnson et al. (1972) working in Europe, who applied this system to the international studies of Ps conducted by Stubbs and colleagues in the Netherlands. In this approach, pathotypes are designated according to decanary notation using an international and European set of differentials. Supplementary sets with regional significance and international relevance were then devised and/or modified by researchers in India (Prashar et al., 2007), southern Africa (Boshoff et al., 2002) and Australia (Wellings and McIntosh, 1990).

Dynamics of pathogenic variation

Mutation, somatic recombination, parasexuality, selection and sexual recombination are considered to be mechanisms determining the genetic variability of Pst (Duan et al., 2010a; Hovmøller et al., 2011; Mboup et al., 2009; Stubbs, 1985). Single‐step mutation of Pst from avirulence to virulence, and the reverse, were concluded to be the origin of new avirulence/virulence combinations (Chen et al., 2009; Hovmøller and Justesen, 2007b; Wellings, 2007; Wellings and McIntosh, 1990). Gassner and Straib (1932b) were the first to propose the role of mutation in the formation of new races of Pst, and estimated a mutation frequency of 0.8 × 10−6 to 1.6 × 10−6. Similarly, the mutation frequencies in a northwest European population of Pst were estimated to range from 1.4 × 10−6 to 4.1 × 10−6 per locus per generation in individual clonal lineages, as determined by amplified fragment length polymorphism (AFLP) analyses (Hovmøller and Justesen, 2007b). Selection of new mutants within Ps populations is governed by matching resistance genes in commercial cereal cropping. Despite a relatively large number of mutant derivates detected in eastern Australia over a 25‐year period, the Pst population was dominated by relatively few pathotypes in particular time periods because of their specific adaptation to widely grown wheat varieties (Wellings, 2007). In addition to mutation and selection, a factor often underestimated in diversity studies in Ps is the chance event of pathotype survival in the non‐host period. The re‐emergence of Pst in any wheat growing season is a function of local survival events or long‐distance dispersal from regions supporting pathogen populations. The pathotypes surviving in these situations often reflect the dominating pathotypes of the previous season, but the actual pathotypes that survive the non‐host period are essentially random and unpredictable.

Although a majority of regional Pst populations show evidence of clonality with diversity limited to closely related mutational derivatives, high diversity for both virulence and molecular markers was discovered in Pst populations in Gansu, China (Duan, et al., 2010a; Mboup et al., 2009; Zheng et al., 2005) and Pakistan (Bahri et al., 2009), suggesting that processes in addition to single‐step mutation were involved. Somatic recombination and parasexuality events between isolates may represent an alternative means to generate new pathogenic diversity (Manners, 1988; Park and Wellings, 2012). Experimental evidence for these events has involved pairing isolates contrasting in urediniospore colour and virulence, combined with cytological observations demonstrating anastomosis events (Kang et al., 1993; Little and Manners, 1969). Hyphal fusion and nuclear re‐assortment may be plausible explanations for the emergence of recombinant isolates in the absence of sexual reproduction. However, somatic recombination has not been demonstrated in Ps populations under natural conditions. The discovery of Berberis spp. as the alternative host for Pst infecting wheat and P. pseudostriiformis infecting bluegrass (Jin et al., 2010) opened up new possibilities to explain the existence of diverse pathogen populations in certain regions. Some evidence to support this hypothesis in China was reported among collections of aecia from Berberis spp. that yielded several pathotypes of Pst (Zhao et al., 2013). However, it is anticipated that sexuality will be localized in Ps populations because of the limited distribution of alternative hosts, climatic conditions and the short dormancy of basidiospores, which will become rapidly exhausted at the end of the wheat growing season (Wellings, 2011).

The dominance of certain pathotypes that cannot be explained on the basis of avirulence/virulence adaptation has led to studies investigating the basis for aggressiveness in certain populations of Pst. Races CYR32 and CYR33 have become the most predominant races in China in recent years, and this was attributed to greater parasitic fitness of both races on leading commercial cultivars (Tian et al., 2008). Two new strains of Pst with the same virulence loci and different AFLP fragments spread rapidly in North America, Australia and Europe in less than 3 years, and were concluded to be highly aggressive. Both strains produced up to two to three times more urediniospores per day than isolates found in the USA and Europe before 2000, which could provide a basis for increased aggressiveness and accelerate their global spread into areas that were previously considered to be too warm for stripe rust epidemics (Hovmøller et al., 2008; Markell and Milus, 2008; Milus et al., 2008). Although temperature adaptation could not be confirmed in Australian studies (Loladze et al., 2013), spore production has been shown previously to be a feature of adapted Pst pathotypes (Johnson and Taylor, 1976).

Population structure and diversity

Population structure and diversity of Ps were traditionally based on avirulence/virulence analyses, and these studies allowed an appreciation of the broad international population characteristics (Stubbs, 1988). With the development of biotechnology, molecular marker techniques based on polymerase chain reaction (PCR) were successfully applied to population structure and diversity studies of Ps from the 1990s. AFLP and random amplification of polymorphic DNA (RAPD) markers revealed a low correlation between virulence and DNA polymorphism in Pst populations in China, the USA and Australia (Chen et al., 1993; Shan et al., 1995; Wellings, 2007). Comparing AFLP lineages, Justesen et al. (2002) demonstrated that certain Pst populations in Denmark were derived from the aerial dispersal of urediniospores from France and/or Germany, and that frequent migration events occurred between the UK, Germany, France and Denmark (Hovmøller et al., 2002). Hovmøller and Justesen (2007a) observed AFLP diversity among unusual pathotypes in northwestern Europe sampled within a short time period in a small area on very few host cultivars with limited or no selection on the pathogen population, and found three to four times higher diversity than that among isolates sampled from a large number of cultivars with different Yr genes in four different countries over 25 years. This suggests a more frequent and even more exotic incursion of Pst urediniospores than previously anticipated. The population of Pst in Yunnan, southwest China, was considered to be a clonal population compared with that in Gansu, northwest China based on AFLP patterns (Liu et al., 2011). These results suggest geographical isolation and local adaptation in the genetic evolution of the Pst population in this region.

Simple sequence repeat (SSR) markers with advantages of co‐dominance and reliability have been developed and applied to Pst. Twelve SSRs distinguished Mediterranean pathotypes specific to southern France from all the other European pathotypes, and separated Chinese isolates into two other classes (Enjalbert et al., 2002). The northern French population was more closely related to the northwestern European population, and the southern clonal population was most probably related to the Mediterranean population. It was concluded that the two subpopulations were the result of an ancient divergence of two clonal lineages (Enjalbert et al., 2005). Multilocus microsatellite genotyping of worldwide Pst samples defined six genetic groups corresponding to their geographical origin. Clonal population structure was observed in northwest European, Middle Eastern, Mediterranean, East and South African, North and South American, and Asian populations. High genotypic diversity and a recombinant structure in Asian populations near the Himalayan Mountain Chain suggested that the Himalayas may represent a centre of diversity for Ps (Vallavielle‐Pope et al., 2012).

Long‐distance dispersal

Pst is capable of long‐distance dispersal by wind movement and human‐assisted transport (Brown and Hovmøller, 2002). Urediniospores of Pst can be efficiently air dispersed over hundreds and perhaps thousands of kilometres despite their vulnerability to environmental factors, such as ultraviolet light (Zadoks, 1961).

In North America, Pst gradually covered a distance of about 2400 km from northern Mexico and southern Texas to North Dakota within 6 months (Chen et al., 2010; Line and Qayoum, 1992).

In Europe, urediniospores can be transported 800 km and even more than 1200 km from north France to Algeria by wind (Stubbs, 1985). The wheat stripe rust pathogen entered New Zealand in 1982, probably as the result of 2000 km of airborne transport from Australia (Wellings and McIntosh, 1990).

In China, evidence suggests that urediniospores are annually dispersed hundreds or even thousands of kilometres from western over‐summering areas to the main wheat belt in the east and north (Fig. 5; Chen and Xie, 1999; Chen et al., 2007; Xie et al., 1992, 1993; Zeng and Luo, 2006). Upland areas 1500–1800 m and 1900–2500 m above sea level in northwestern and southwestern China, respectively, provide environments for year‐round pathogen survival and development, and hence these regions are major reservoirs of inoculum and pathogenic diversity (Brown and Hovmøller, 2002; Chen et al., 2007, 2009). The pathogen must re‐establish each autumn in northern and eastern China, the main winter wheat‐growing areas, because the pathogen cannot survive the hot dry summers. Ecological management of wheat stripe rust in the areas of inoculum sources has been considered the major strategy for the sustainable control of wheat stripe rust nationwide, resulting in effective containment of wheat yield losses that potentially amount to more than two million tonnes annually (Chen WQ et al., 2012, 2013).

figure

Ecological zones of wheat stripe rust and aerial dispersal of urediniospores in China. 1, The Guanzhong and Huabei winter wheat region is the main area of winter wheat cultivation and the over‐wintering area of Puccinia striiformis f. sp. tritici (Pst) in China. 2, The Chengdu Plain and Jianghan River Basin facultative wheat region is the winter‐increasing area and the major spring inoculum source of Pst. 3, The Northwest winter and spring wheat region is the most important over‐summering area of Pst and the major source of inoculum for wheat infection in autumn. 4, The Yunnan and Guizhou wheat region is one of the over‐summering areas of Pst. 5, The Xinjiang winter and spring wheat region is a relatively independent epidemic zone of Pst. 6, The Tibet highland barley and wheat region is a separate epidemic zone of Pst. 7, The South late‐sowing wheat region does not grow much wheat and stripe rust seldom occurs. 8, The Inner Mongolia and Northeast spring wheat region where wheat stripe rust epidemics occur occasionally. The full black arrows show dispersal of urediniospores from the over‐summering areas to the main wheat‐growing areas in autumn annually. The broken black arrows indicate the main pathway of urediniospore dispersal in spring. The red line indicates the boundary of Pst over‐wintering areas and the red circles show provincial capitals.

In the US Pacific Northwest, where both winter and spring wheat are grown, Pst urediniospores spread readily from winter wheat to spring wheat in the summer and from spring wheat to winter wheat in autumn, in addition to survival on volunteer wheat plants and wild grass species (Chen, 2005; Hendrix et al., 1965; Line, 2002).

In Australia, Pst was first detected in 1979 and the pathogenic features of the initial isolates predicted that it was of European origin. Wellings et al. (1987) provided evidence to support the survival of urediniospores on contaminated travellers’ clothing, and predicted that this was the most likely means of entry of Pst into Australia. The first detection of Pst in remote wheat‐growing regions previously free of the disease in South Africa in 1996 and in Western Australia in 2002 has similarly implicated long‐distance transport assisted by adherent urediniospores on travellers’ clothing (Wellings, 2007).

Molecular Perspectives

Host–pathogen interactions

Ling et al. (2007) published a full‐length cDNA library, consisting of 42 240 clones with an average cDNA insert of 1.9 kb, constructed from urediniospores of US race PST‐78. Initial characterization and gene analysis were conducted by sequencing about 200 randomly selected clones representing various lengths of open reading frames. The first 51 genes with putative functions involving 11 aspects of pathogen cell biology and pathogenicity were identified.

Cantu et al. (2011) accessed the genomic sequence of higher virulent US race PST‐130 using next generation sequencing (NGS), and obtained nearly 80 million high‐quality paired‐end reads that were assembled into 29 178 contigs (64.8 Mb); 22 825 putative coding sequences were identified and tentatively annotated. Cantu et al. (2013) re‐sequenced the genome of four Pst isolates from the US and UK to identify effector candidates and to relate them to their distinct virulence profiles. RNAseq analysis highlighted transcripts encoding secreted proteins that were significantly enriched in haustoria compared with infected tissue. The expression of 22 candidate effector genes was characterized using quantitative reverse transcription‐polymerase chain reaction (qRT‐PCR), revealing distinct temporal expression patterns during infection in wheat. Five polymorphic effector candidates specifically between two UK isolates, which differ in virulence to two wheat cultivars, were identified among 2999 secreted proteins. These allelic variants are now priority for functional validation as virulence/avirulence effectors in the corresponding wheat cultivars. The comparative sequence analysis of the Pst races may provide some candidate genes for the effectors recognized by the stripe rust resistance genes.

A total of 12 282 transcripts of Pst race 104E137A were assembled by means of NGS platforms to compare the germinated urediniospores and haustoria transcriptomes based on Illumina RNAseq data (Garnica et al., 2013). More than 400 genes encoding secreted proteins which constitute candidate effectors were identified from the haustorial transcriptome, with two‐thirds of these up‐regulated in the tissue of wheat relative to germinated spores. RT‐PCR analysis confirmed the expression patterns of 94 effector candidates. The analysis also revealed that urediniospores rely mainly on stored energy reserves for growth and development, whereas haustoria take up host nutrients for massive energy production for the biosynthetic pathway and the ultimate production of spores.

Zhang et al. (2008) isolated 4798 expressed sequence tags (ESTs) derived from a germinated urediniospore library and 267 genes with putative functions were identified from Chinese race CYR32. blastx searches revealed 13 ESTs homologous to known fungal pathogenicity or virulence factors, and six were shown to have high levels of expression in germinated urediniospores. Potential virulence factors were detected among 15 unique transcripts from a cDNA library constructed from haustoria isolated from Pst‐infected wheat leaves (Yin et al., 2009). Despite some success in transient gene expression using the β‐glucuronidase (GUS) reporter and the antibiotic resistance gene hygromycin phosphotransferase (hpt) via particle bombardment (Wang et al., 2006), Yin et al. (2009) were unable to confirm the specific functions of these genes in the infection process. Hypersensitivity‐induced reaction genes (Ta‐hir1, Ta‐hir2, Ta‐hir3, Ta‐hir4) were characterized in wheat–Pst interactions (Yu et al., 2008; Zhang et al., 2009). Studies have revealed an early oxidative burst of reactive oxygen species in the hypersensitive interaction, and have implicated the role of Ca2+ as an intracellular secondary messenger in pathogen defence processes (Long et al., 2010).

Several studies have made insights into the molecular basis of compatible interactions between wheat and Pst. Yu et al. (2008) used suppression subtractive hybridization to identify the genes induced in the fungus during infection, although only two unigenes showed similarity with pathogenic proteins. Ma et al. (2009) constructed an EST library from fully susceptible wheat leaves infected with Pst to explore the pathogen genes expressed during infection. Among the 2743 unisequences, 446 showed homologies to fungal genes from model fungal species, and 15 genes were homologous to other rust fungal genes. Using a complementary DNA‐AFLP, nine transcript‐derived fragments from wheat leaves infected by Pst were shown to be of pathogenic origin (Wang et al., 2009).

Host defence‐related genes isolated from interactions between wheat and Pst corresponded to different stages in the infection process. Coram et al. (2008b) reported that 54 transcripts were induced in both compatible and incompatible interactions, and were considered as basal defence transcripts, whereas 61 transcripts were specific to the incompatible interaction (hypersensitive resistance‐specific transcripts) and 19 were specific to the compatible interaction (biotrophic interaction‐specific transcripts). A fascinating discovery was the quenching of the divergent expression of Pst‐regulated genes in both incompatible and compatible interactions in the middle stages of Pst infection (Wang et al., 2009).

In addition, a number of candidate genes from wheat challenged by Pst, such as a transcription factor gene TabZIP1 (Zhang et al., 2009), a novel wheat NAC gene TaNAC4 (Xia et al., 2010), a wheat HSP70 gene TaHSC70 (Duan et al., 2010b), a wheat β‐1,3‐glucanase gene TaGlu (Liu et al., 2010), a pathogenesis‐related thaumatin‐like protein gene TaPR5 (Wang et al., 2010) and three secreted protein genes PstSP2C7, PstSP11L10 and PstSP11P10 (Dong et al., 2011) have been characterized. Meta‐analysis of the 28 transcripts confirmed the activity of known resistance (R) gene‐mediated pathways in the race‐specific resistance response, including an oxidative burst that probably contributes to hypersensitive resistance, as well as pathogenesis‐related protein expression and activity of the phenylpropanoid pathway (Coram et al., 2010).

Resistance gene cloning

Early successes in resistance gene cloning, such as Yr10, focused on seedling effective genes which had generally been overcome by new pathotypes (Spielmeyer and Lagudah, 2003). The cloned gene sequences from this, and similar resistances to leaf rust caused by P. triticina, showed common molecular motifs that included nucleotide‐binding sites which were rich in leucine repeat elements (NBS‐LRR). The classical NBS‐LRR gene family is considered to have a role in ubiquitous programmed cell death (apoptosis), putative antifungal activities, disease resistance responses, pathogenesis‐related responses and unknown functions providing race‐specific resistance to stripe rust. In contrast, the reputed durable sources of resistance Yr18 and Yr36 showed molecular motifs with distinctly different functions. The former was shown to encode a protein resembling a multidrug ABC transporter (Krattinger et al., 2009), whereas the latter includes a kinase and a putative START lipid‐binding domain (Fu et al., 2009). Functional studies of Yr39 by Coram et al. (2008a) indicated evidence for broad defence responses, including the induction of several R protein homologues, wider induction of the phenylpropanoid pathway and several other putative defence transcripts. These mechanisms are functionally different from those of the NBS‐LRR resistance genes, and hence Yr39 was predicted and proven to be a durable source of resistance to Pst (Lin and Chen, 2007; Coram et al., 2008a).

Conclusions

The availability of Pst sequence data and the integration of genomics, transcriptomics and effector‐directed annotation have led to a more comprehensive understanding of the Pst pathogenesis system, an important step towards the development of more effective surveillance and management strategies. The cloning of durable resistance genes that demonstrate molecular motifs distinct from those of failed race‐specific genes indicates hope for forging substantial advances in determining the molecular nature of durable resistance, and so to provide possibilities for their effective deployment in commercial wheat cultivars.

Genes for Stripe Rust Resistance in Wheat

The genetic characterization of resistance has been an active focus of research since the seminal work of Biffen (1905), who first described the Mendelian nature of resistance to stripe rust in wheat cultivar Rivet. The major types of resistance to Pst in wheat are seedling (or all‐stage) resistance and adult plant resistance (APR). The former is conveniently detected in seedling tests and remains effective throughout all growth stages provided that the same Pst pathotype is used in all assessments. APR, which expresses susceptible infection types in seedling tests, develops varying levels of resistance in post‐seedling stages in either field or glasshouse studies. Major reviews of the genetic basis for resistance to Pst can be found in Chen (2005, 2013), Röbbelen and Sharp (1978) and Wellings et al. (2012).

Seedling resistances are frequently conferred by single genes, or simple combinations of single genes, and have generally become vulnerable to single‐gene changes to increased virulence in Pst populations. APR has been considered to be more robust in terms of resistance in that it has not been as readily overcome by changing Pst populations, although there have been notable severe epidemics arising from the failure of single‐gene APR, for example, the sudden susceptibility of Joss Cambier in the UK in 1969 (Johnson and Taylor, 1972). A major conceptual development in searching for strategies to contain stripe rust epidemics was the description of durable resistance, which was defined as wheat cultivars that remained resistant when cultivated over large areas, for many years and in environments conducive to Pst epidemics (Johnson and Law, 1975).

Seedling resistance

Lupton and Macer (1962) studied the seedling‐expressed resistance in seven wheat cultivars and first assigned Yr symbols to designate stripe rust resistance genes (Yr genes) in wheat. New designations of Yr genes are reviewed by an international consultation group, led by Professor R. A. McIntosh, who publishes annual updates (McIntosh et al., 2007) and complete revisions every 5 years to coincide with the International Wheat Genetics Symposium (McIntosh et al., 2003). To date, more than 50 officially designated and many temporarily designated Yr genes for resistance to stripe rust have been described in wheat, and some have been widely used in different areas of the world (McIntosh et al., 1995, 2003, 2007; Wellings et al., 2012; Xu et al., 2013). Among a total of 55 catalogued Yr genes, 41 confer seedling resistance, and 14 are APR genes. A majority of Yr genes originate from T. aestivum, but several are derived from related genera or species, including Secale cereale (Yr9), Aegilops spp. (Yr8, Yr17, Yr37, Yr38, Yr40 and Yr42), T. spelta (Yr5), T. dicoccoides (Yr15, Yr35 and Yr36), T. turgidum (Yr24/Yr26, Yr53), T. tauschii (Yr28) and Thinopyrum intermedium (Yr50).

Adult plant resistance

APR becomes effective at various stages of post‐seedling development according to genotype (Boyd, 2006). Several resistance genes, including Yr16, Yr18, Yr29, Yr30, Yr39 and Yr52, confer APR, and some, for example Yr18, are regarded as durable (Imtiaz, 2004; Morgounov et al., 2012).

APR that begins early in the growth cycle will be expected to offer greater yield protection than that operating from flag leaf emergence. Other factors governing the expression of APR are crop nutritional status, where high nitrogen leads to more severe disease expression, and temperature. High‐temperature APR (HTAP) is described as quantitatively inherited and pathotype non‐specific, and therefore is concluded to be durable resistance (Chen, 2005, 2013). HTAP, which is triggered in the late stages of plant development when average daily temperatures are typically above 21 °C, causes the initial infection types and severity to decrease, preventing the occurrence of secondary infection events. Genes for HTAP resistance have been genetically characterized, including Yr36 from T. turgidum var. dicoccoides located on chromosome 6BS (Uauy et al., 2005), Yr39 in spring wheat cultivar ‘Alpowa’ located on 7BL (Lin and Chen, 2007), and Yr52 on 7BL in spring wheat line ‘PI 183527’ (Ren et al., 2012).

Quantitative resistance

Quantitative trait loci (QTLs) conferring resistance, which may have individually small effects on reducing disease expression, can contribute collectively in an additive manner to confer high levels of resistance to Pst and segregate according to Mendel's laws (Singh et al., 2000; Singh and Rajaram, 1994; Tanksley, 1993). A large number of QTLs for resistance to Pst and their associated molecular markers have been reviewed by Boyd (2005), Chen (2005, 2013), Singh et al. (2004) and Wellings et al. (2012). QTLs for resistance to Pst have been mapped to all wheat chromosomes except 1D and 3A (Chen, 2005, 2013; Christopher et al., 2013; Imtiaz et al., 2004; Lu et al., 2009; Suenaga et al., 2003; Vazquez et al., 2012).

Linkage relationships

All designated Yr genes, with the exception of pathotype‐specific APRs Yr11 to Yr14, have been mapped to chromosomal positions across the Triticum genome, although chromosome 7A remains currently free of described Yr genes. Gene location provides the basis for linkage relationships that are important to the development of breeding strategies and in determining the potential value or limitation in deploying certain Yr genes. Yr9 on chromosome 1B is linked to leaf and stem rust resistance genes (Lr26 and Sr31, respectively) and to a quality defect in flour used for bread making (McIntosh et al., 1995). Yr18, which locates to chromosome 7DS, is the same gene as Lr34 for resistance to leaf rust and resistance to stem rust and powdery mildew (Krattinger et al., 2009; Spielmeyer et al., 2005; Suenaga et al., 2003). The slow rusting APR gene Yr29 on 1BL is completely linked to gene Lr46, which confers moderate resistance to leaf rust (William et al., 2003). Yr30 on 3BS, involved in APR among several cultivars developed by the International Maize and Wheat Improvement Center, is associated with the durable stem rust resistance gene Sr2 (Singh et al., 2000).

Conclusions

The wide distribution and frequent occurrence of Pst epidemics in world wheat production zones will continue to yield focused effort in the development and deployment of economic control strategies. The increasing use of fungicide management for Pst control over the past decade has been driven by the availability of off‐patent generic products that have reduced chemical costs and thereby increased commercial economic benefit, even in production systems that have relatively low yield expectations. In addition, the emergence of new Pst pathotype lineages across large production regions, either through exotic incursion events or through local evolution to increased virulence, has evidently thwarted the capacity for breeding programmes to quickly respond with varieties that meet grower's demands for resistance, high yield and market quality. Although resistant varieties remain the goal, commercial industry imperatives will mean that alternative approaches, such as integrating fungicides and cultivation measures (including the regulation of plant date, crop nutrition, interplanting of crops and eradication of volunteer seedlings), may assume greater importance in the short to medium term. Specific management strategies in areas which function as pathogen inoculum sources and center of pathogenic diversity are considered to be the highest priority for sustainable stripe rust control.

In order to advance the development and distribution of resistant wheat cultivars to provide cost‐effective and environmentally acceptable disease control, research on Pst will continue to build on the large body of information already assembled for this important cereal pathogen. Molecular studies have assisted the identification of certain resistance genes that are likely to provide durable resistance. Efficient marker systems are available for these genes and should enable breeders to incorporate and select several durable genes in populations to build genetic foundations that stabilize and secure their germplasm against the uncertainties of Pst dynamics. Although gains in durable resistance are being made, it is essential that genetic diversity is encouraged through the identification and characterization of further durable sources for use in practical plant breeding. Seedling resistances will continue to be described and catalogued, but the application of these genes will centre around debates of stewardship that arise from concerns about the potential evolution and selection of new races and the means to minimize these effects using strategies ranging from the development of gene combinations to regional resistance gene deployment. The annotation and functional analysis of genes involved in the biology and pathogenicity of Pst by means of bioinformatics will provide new opportunities to understand the virulence variation in the pathogen and the mechanism of resistance in the host.

In parallel to host resistance, population studies in Pst will continue to focus on mechanisms underpinning pathogenic variability within and between hosts. An important area will be to evaluate the significance of apparent restricted sexual populations of Pst in certain regions and the role that the alternative Berberis hosts may play in generating variability and providing the opportunity for pathogen survival between cereal cropping seasons. Features of pathogen epidemiology, including mechanisms of survival between seasons, and interactions of disease onset and development in relation to variety and environment, remain relevant issues to disease control.

Acknowledgements

Financial support by the Ministry of Science and Technology, China (Grant nos. 2013CB127701 and 20011CB100403) and Ministry of Agriculture, China (Grant nos. 200903035 and CARS‐03‐04B) is gratefully acknowledged. We also thank our colleagues Drs Jing Feng, Bo Liu, Li Gao, Jie Zhao and Xiaojie Wang for their contribution to the survey and literature collection.

References

  • Ali M., Ji W.Q., Hu Y.G., Baloch G.M., Zhong H. and Wang C.Y. (2010a) Molecular implications from SSR markers for stripe rust (Puccinia striiformis f. sp. tritici) resistance gene in bread wheat line N95175. Pakistan J. Botany 42, 383–390. [Google Scholar]
  • Ali S., Leconte M., Walker A.S., Enjalbert J. and de Vallavieille‐Pope C. (2010b) Reduction in the sex ability of worldwide clonal populations of Puccinia striiformis f. sp. tritici . Fungal Genet. Biol. 47, 828–838. [Abstract] [Google Scholar]
  • Bahri B., Leconte M., Ouffroukh A., De Vallavieille‐Pope C. and Enjalbert J. (2009) Geographic limits of a clonal population of wheat yellow rust in the Mediterranean region. Mol. Ecol. 18, 4165–4179. [Abstract] [Google Scholar]
  • Bever W.M. (1934) Physiologic specialization in Puccinia glumarum in the U. S. Phytopathology, 24, 686–688. [Google Scholar]
  • Biffen R.H. (1905) Mendel’ s law of inheritance and wheat breeding. J. Agric. Sci. 1, 4–48. [Google Scholar]
  • Boshoff W.H.P., Pretorius Z. and van Niekerk B.D. (2002) Establishment, distribution, and pathogenicity of Puccinia striiformis f. sp. tritici in South Africa. Plant Dis. 86, 485–492. [Abstract] [Google Scholar]
  • Boyd L.A. (2005) Can Robigus defeat an old enemy?—Yellow rust of wheat. J. Agric. Sci. 143, 233–243. [Google Scholar]
  • Boyd L.A. (2006) Perspective: can the durability of resistance be predicted? J. Sci. Food Agric. 86, 2523–2526. [Google Scholar]
  • Britton M. and Cummins G.B. (1956) The reaction of species of Poa and grasses to Puccinia striiformis . Plant Dis. Rep. 40, 643–645. [Google Scholar]
  • Brown J.K.M. and Hovmøller M. (2002) Aerial dispersal of pathogens on the global and continental scales and its impact on plant disease. Science, 297, 537–541. [Abstract] [Google Scholar]
  • Cantu D., Govindarajulu M., Kozik A., Wang M., Chen X.M., Kojima K.K.K.K., Jurka J., Michelmore R.W. and Dubcovsky J. (2011) Next generation sequencing provides rapid access to the genome of Puccinia striiformis f. sp. tritici, the causal agent of wheat stripe rust. PLoS ONE, 6, e24230. [Europe PMC free article] [Abstract] [Google Scholar]
  • Cantu D., Segovia V., Maclean D., Dayles R., Chen X., Kamoun S., Dubcovsky J., Saunders D.G.O. and Uauy C. (2013) Genome analyses of the wheat yellow (stripe) rust pathogen Puccinia striiformis f. sp. tritici reveal polymorphic and haustorial expressed secreted proteins as candidate effectors. BMC Genomics, 14, 270. [Europe PMC free article] [Abstract] [Google Scholar]
  • Carleton M.A. (1915) A serious new wheat rust in this country. Science, 42, 58–59. [Abstract] [Google Scholar]
  • Chen W.Q. and Xie S.X. (1999) Ecological zones and inter‐regional dispersal of wheat stripe rust in China In: Research Progress in Plant Protection and Plant Nutrition (Hong F.Z., editor; and Li K.X., editor. , eds), pp. 276–277. Beijing: China Agriculture Press. [Google Scholar]
  • Chen W.Q., Xu S.C. and Wu L.R. (2007) Epidemiology and sustainable management of wheat stripe rust caused by Puccinia striiformis West. in China: a historical retrospect and prospect. Sci. Agric. Sin. 40, 177–183. [Google Scholar]
  • Chen W.Q., Wu L.R., Liu T.G., Xu S.C., Jin S.L., Peng Y.L. and Wang B.T. (2009) Pathotype dynamics, diversity and virulence evolution in Puccinia striiformis f. sp. tritici, the causal agent of wheat stripe (yellow) rust in China from 2003 to 2007. Plant Dis. 93, 1093–1101. [Abstract] [Google Scholar]
  • Chen W.Q., Xu S.C., Liu T.G., Jin S.L., Kang Z.S. and Ma Z.H. (2012) Strategies and technologies of integrated management on wheat stripe rust in China In: Disease Risk and Food Security. Proceedings of the 13th International Cereal Rust and Powdery Mildew Conference (Chen W.Q., editor. et al, eds), p. 203 Beijing: China Agricultural Science and Technology ; Press. [Google Scholar]
  • Chen W.Q., Kang Z.S., Ma Z.H., Xu S.C., Jin S.L. and Jiang Y.Y. (2013) Integrated management of wheat stripe rust caused by Puccinia striiformis f. sp. tritici in China. Sci. Agric. Sin. 46, 4254–4262. [Google Scholar]
  • Chen X.M. (2005) Epidemiology and control of stripe rust (Puccinia striiformis f. sp. tritici) on wheat. Can. J. Plant Pathol. 27, 314–337. [Google Scholar]
  • Chen X.M. (2013) High‐temperature adult‐plant resistance, key for sustainable control of stripe rust. Am. J. Plant Sci. 4, 608–627. [Google Scholar]
  • Chen X.M. and Line R.F. (1993) Inheritance of stripe rust (yellow rust) resistance in the wheat cultivar Carstens V. Euphytica, 71, 107–113. [Google Scholar]
  • Chen X.M., Line R.F. and Leung H. (1993) Relationship between virulence variation and DNA polymorphism in Puccinia striiformis . Phytopathology, 83, 1489–1497. [Google Scholar]
  • Chen X.M., Moore M.K., Milus E.A., Long D.L., Line R.F., Marshall D. and Jackson L. (2002) Wheat stripe rust epidemics and races of Puccinia striiformis f. sp. tritici in the United States in 2000. Plant Dis. 86, 39–46. [Abstract] [Google Scholar]
  • Chen X.M., Penman L., Wan A.M. and Cheng P. (2010) Virulence races of Puccinia striiformis f. sp. tritici in 2006 and 2007 and development of wheat stripe rust and distributions, dynamics, and evolutionary relationships of races from 2000 to 2007 in the United States. Can. J. Plant Pathol. 32, 315–333. [Google Scholar]
  • Chen X.M., Wang M.N., Wan A.M., Cheng P. and Cheng J.J. (2012) Sexual or asexual reproduction, which one is more important for stripe rust In: Disease Risk and Food Security. Proceedings of the 13th International Cereal Rust and Powdery Mildew Conference (Chen W.Q., editor. et al, eds), pp. 36–37. Beijing: China Agricultural Science and Technology ; Press. [Google Scholar]
  • Christopher M.D., Liu S.Y., Hall M.D., Marshall D.S., Fountain M.O., Johnson J.W., Milus E.A., Garland‐Campbell K.A., Chen X.M. and Griffey C.A. (2013) Identification and mapping of adult plant stripe rust resistance in soft red winter wheat cultivar ‘USG 3555’. Plant Breed. 132, 53–60. [Google Scholar]
  • Coram T.E., Settles M.L. and Chen X.M. (2008a) Transcriptome analysis of high‐temperature adult‐plant resistance conditioned by Yr39 during the wheat–Puccinia striiformis f. sp. tritici interaction. Mol. Plant Pathol. 9, 479–493. [Europe PMC free article] [Abstract] [Google Scholar]
  • Coram T.E., Wang M.N. and Chen X.M. (2008b) Transcriptome analysis of the wheat–Puccinia striiformis f. sp tritici interaction. Mol. Plant Pathol. 9, 157–169. [Europe PMC free article] [Abstract] [Google Scholar]
  • Coram T.E., Huang X., Zhan G., Settles M.L. and Chen X. (2010) Meta‐analysis of transcripts associated with race‐specific resistance to stripe rust in wheat demonstrates common induction of blue copper‐binding protein, heat‐stress transcription factor, pathogen‐induced WIR1A protein, and ent‐kaurene synthase transcripts. Funct. Integr. Genomics, 10, 383–392. [Abstract] [Google Scholar]
  • Dong Y.L., Yin C.T., Hulbert S., Chen X.M. and Kang Z.S. (2011) Cloning and expression analysis of three secreted protein genes from wheat stripe rust fungus Puccinia striiformis f. sp. tritici . World J. Microbiol. Biotechnol. 27, 1261–1265. 10.1007/s11274-010-0565-6. [CrossRef] [Google Scholar]
  • Duan X.Y., Tellier A., Wan A.M., Leconte M., de Vallavieille‐Pope C. and Enjalbert J. (2010a) Puccinia striiformis f. sp. tritici presents high diversity and recombination in the over‐summering zone of Gansu‐China. Mycologia, 102, 44–53. [Abstract] [Google Scholar]
  • Duan Y.H., Guo J., Ding K., Wang S.J., Zhang H., Dai X.W., Chen Y.X., Govers F., Huang L.L. and Kang Z.S. (2010b) Characterization of a wheat HSP70 gene and its expression in response to stripe rust infection and abiotic stresses. Mol. Biol. Rep. 38, 301–307. 101007/s11033‐010‐0108‐0. [Abstract] [Google Scholar]
  • Enjalbert J., Duan X.Y., Giraud T., Vautrin D., de Vallavielle‐Pope C. and Solignac M. (2002) Isolation of twelve microsatellite loci, using an enrichment protocol, in the phytopathogenic fungus Puccinia striiformis f.sp tritici . Mol. Ecol. Notes, 2, 563–565. [Google Scholar]
  • Enjalbert J., Duan X.Y., Leconte M., Hovmøller M.S. and de Vallavieille‐Pope C. (2005) Genetic evidence of local adaptation of wheat yellow rust (Puccinia striiformis f. sp tritici) within France. Mol. Ecol. 14, 2065–2073. [Abstract] [Google Scholar]
  • Eriksson J. (1894) Uber die Spezialisierung des Parasitismus bei den Getreiderostpilzen. Ber. Dtsch. Bot. Ges. 12, 292–331. [Google Scholar]
  • Fang C.T. (1944) Physiologic specialization of Puccinia glumarum Erikss. and Henn. in China. Phytopathology, 34, 1020–1024. [Google Scholar]
  • Fraser W. and Conners I. (1925) The Uredinales of the prairie provinces of Western Canada. Trans. R. Soc. Can. 19, 275–308. [Google Scholar]
  • Fu D., Uauy C., Distelfeld A., Blechl A., Epstein L., Chen X., Sela H., Fahima T. and Dubcovsky J. (2009) A kinase‐START gene confers temperature‐dependent resistance to wheat stripe rust. Science, 323, 1357–1360. [Europe PMC free article] [Abstract] [Google Scholar]
  • Fuchs E. (1965) Untersuchungen über die physiologische spezialisierung des weizengelbrostes (Puccinia striiformis West. f.sp. tritici Erikks. et. Henn.) in den jahren 1959–1964 und über das anfälligkeitsverhalten einiger weizensorten. Nachrbl. Dtsch. Pflanzenschutzd. 17, 161–176. [Google Scholar]
  • Garnica D.P., Upadhyaya N.M., Dodds P.N. and Rathjen J.P. (2013) Strategies for wheat stripe rust pathogenicity identified by transcriptome sequencing. PLoS One, 8, e67150. [Europe PMC free article] [Abstract] [Google Scholar]
  • Gassner G. and Straib W. (1930) Uber das Auftreten einer neuen Gelbrostform auf Weizen. Züchter, 2, 313–317. [Google Scholar]
  • Gassner G. and Straib W. (1932a) Die bestimmung der biologische rasen des weizengelbrostes (Puccinia glumarum f.sp. tritici (Schmt.) Erikss. und Henn). Arb. Boil. Reichsanst Land Forstw. 20, 141–163. [Google Scholar]
  • Gassner G. and Straib W. (1932b) Uber mutation in einer biologischen Rasse von Puccinia glumarum tritici (Schmidt). Erikss. U. Henn. Z. Indukt. Abstamm. Vererbungsl. 63, 154–160. [Google Scholar]
  • Hahn M. (2000) The rust fungi: cytology, physiology and molecular biology of infection In: Fungal Pathology (Kronstad J., editor. , ed.), pp. 267–306. Dordrecht: Kluwer Academic Publishers. [Google Scholar]
  • Hahn M. and Mendgen K. (2001) Signal and nutrient exchange at biotrophic plant–fungus interfaces. Curr. Opin. Plant Biol. 4, 322–327. [Abstract] [Google Scholar]
  • Hahn M., Deising H., Struck C. and Mendgen K. (1997) Fungal morphogenesis and enzyme secretion during pathogenesis In: Resistance of Crop Plants against Fungi (Hartleb H., editor; , Heitefuss R., editor; and Hoppe H.H., editor. , eds), pp. 33–57. Jena: Fischer. [Google Scholar]
  • Hassebrauk K. (1965) Nomenklatur, geographische Verbreitung und Wirtsbereich des Gelbrostes, Puccinia striiformis West. Mitt. Biol. Bundesanst. Land. 116, 1–75. [Google Scholar]
  • Heath M.C. and Skalamera D. (1997) Cellular interactions between plants and biotrophic fungal parasites. Adv. Bot. Res. 24, 195–225. [Google Scholar]
  • Hendrix J.W., Burleigh J.R. and Tu J.C. (1965) Oversummering of stripe rust at high elevations in the Pacific Northwest—1963. Plant Dis. Rep. 49, 275–278. [Google Scholar]
  • Hermansen J.E. and Veterinary R. (1968) Studies on the Spread and Survival of Cereal Rust and Mildew Diseases in Denmark. Copenhagen: The Royal Veterinary and Agricultural College. [Google Scholar]
  • Hovmøller M.S. and Justesen A.F. (2007a) Appearance of atypical Puccinia striiformis f. sp tritici phenotypes in north‐western Europe. Aust. J. Agric. Res. 58, 518–524. [Google Scholar]
  • Hovmøller M.S. and Justesen A.F. (2007b)) Rates of evolution of avirulence phenotypes and DNA markers in a northwest European population of Puccinia striiformis f. sp tritici . Mol. Ecol. 16, 4637–4647. [Abstract] [Google Scholar]
  • Hovmøller M.S., Justesen A.F. and Brown J.K.M. (2002) Clonality and long‐distance migration of Puccinia striiformis f.sp tritici in north‐west Europe. Plant Pathol. 51, 24–32. [Google Scholar]
  • Hovmøller M.S., Yahyaoui A.H., Milus E.A. and Justesen A.F. (2008) Rapid global spread of two aggressive strains of a wheat rust fungus. Mol. Ecol. 17, 3818–3826. [Abstract] [Google Scholar]
  • Hovmøller M.S., Walter S. and Justesen A.F. (2010) Escalating threat of wheat rusts. Science, 329, 369. [Abstract] [Google Scholar]
  • Hovmøller M.S., Sørensen C.K., Walter S. and Justesen A.F. (2011) Diversity of Puccinia striiformis on cereals and grasses. Annu. Rev. Phytopathol. 49, 20.1–20.21. [Abstract] [Google Scholar]
  • Humphrey H.B., Hungerford C.W. and Johnson A.G. (1924) Stripe rust (Puccinia glumarum) of cereals and grasses in the United States. J. Agric. Res. 29, 209–227. [Google Scholar]
  • Imtiaz M., Ahmad M., Cromey G., Grifin W.B. and Hampton J.G. (2004) Detection of molecular markers linked to the durable adult plant stripe rust resistance gene Yr18 in bread wheat (Triticum aestivum L.). Plant Breed. 123, 401–404. [Google Scholar]
  • Jin Y., Szabo L.J. and Carson M. (2010) Century‐old mystery of Puccinia striiformis life history solved with the identification of Berberis as an alternate host. Phytopathology, 100, 432–435. [Abstract] [Google Scholar]
  • Johnson R. and Law C.N. (1975) Genetic control of durable resistance to stripe rust (Puccinia striiformis) in the wheat cultivar Hybride de Berse. Ann. Appl. Biol. 81, 385–391. [Google Scholar]
  • Johnson R. and Taylor A.J. (1972) Isolates of Puccinia striiformis collected in England from the wheat varieties Maris Beacon and Joss Cambier. Nature, 238, 105–106. [Google Scholar]
  • Johnson R. and Taylor A.J. (1976) Spore yield of pathogens in investigations of the race‐specificity of host resistance. Ann. Rev. Phytopathol. 14, 91–119. [Google Scholar]
  • Johnson R., Stubbs R.W., Fuchs E. and Chamberlain N.H. (1972) Nomenclature for physiological races of Puccinia striiformis infecting wheat. Trans. Br. Mycol. Soc. 58, 475–480. [Google Scholar]
  • Justesen A.F., Ridout C.J. and Hovmøller M.S. (2002) The recent history of Puccinia striiformis f.sp tritici in Denmark as revealed by disease incidence and AFLP markers. Plant Pathol. 51, 13–23. [Google Scholar]
  • Kang Z.S. (1996) Ultrastructure of Plant Pathogenic Fungi. Beijing: China Science & Technology Press. [Google Scholar]
  • Kang Z.S., Li Z.Q., Shang H.S., Chong J. and Rohringer R. (1993) Ultrastructure and cytochemistry of intercellular hyphae of wheat stripe rust. Acta Mycol. Sin. 24, 26–31. [Google Scholar]
  • Kang Z.S., Li Z.Q. and Shang H.S. (1994) Nucleal condition of uredinial stage of wheat stripe rust. Acta Phytopathol. Sin. 24, 26–31. [Google Scholar]
  • Kang Z.S., Huang L.L. and Li J.Y. (1997) Morphology of Plant Pathogenic Fungi under Scanning Electron Microscope. Beijing: China Agriculture Press. [Google Scholar]
  • Kang Z.S., Huang L.L. and Buchenauer H. (2002) Ultrastructural changes and localization of lignin and callose in compatible and incompatible interactions between wheat and Puccinia striiformis . J. Plant Dis. Prot. 109, 25–37. [Google Scholar]
  • Kang Z.S., Wang Y., Huang L.L., Wei G.R. and Zhao J. (2003) Histology and ultrastructure of incompatible combination between Puccinia striiformis and wheat with low reaction type resistance. Agric. Sci. China, 2, 1102–1113. [Google Scholar]
  • Kapooria R.G. and Mendgen K. (1985) Infection structures and their surface changes during differentiation in Uromyces fabae . Phytopathol. Z. 113, 317–323. [Google Scholar]
  • Krattinger S.G., Lagudah E.S., Spielmeyer W., Singh R.P., Huerta‐Espino J., McFadden H., Bossolini E., Selter L.L. and Keller B. (2009) A putative ABC transporter confers durable resistance to multiple fungal pathogens in wheat. Science, 323, 1360–1363. [Abstract] [Google Scholar]
  • Lin F. and Chen X.M. (2007) Genetics and molecular mapping of genes for race‐specific all‐stage resistance and non‐race‐specific high‐temperature adult‐plant resistance to stripe rust in spring wheat cultivar Alpowa. Theor. Appl. Genet. 114, 1277–1287. [Abstract] [Google Scholar]
  • Line R.F. (2002) Stripe rust of wheat and barley in North America: a retrospective historical review. Annu. Rev. Phytopathol. 40, 75–118. [Abstract] [Google Scholar]
  • Line R.F. and Qayoum A. (1992) Virulence, aggressiveness, evolution and distribution of races of Puccinia striiformis (the cause of stripe rust of wheat) in North America, 1968–87. USDA Bull. 1788, 44. [Google Scholar]
  • Line R.F., Sharp E. and Powelson R. (1970) A system for differentiating races of Puccinia striiformis in the United States. Plant Dis. Rep. 54, 992–994. [Google Scholar]
  • Ling P., Wang M.N., Chen X.M. and Campbell K. (2007) Construction and characterization of a full‐length cDNA library for the wheat stripe rust pathogen (Puccinia striiformis f. sp. tritici). BMC Genomics, 8, 145–157. [Europe PMC free article] [Abstract] [Google Scholar]
  • Little R. and Manners J.G. (1969) Somatic recombination in yellow rust of wheat (Puccinia striiformis). II. Germ tube fusions, nuclear number and nuclear size. Trans. Br. Mycol. Soc. 53, 259–267. [Google Scholar]
  • Liu M. and Hambleton S. (2010) Taxonomic study of stripe rust, Puccinia striiformis sensu lato, based on molecular and morphological evidence. Fungal Biol. 114, 881–899. [Abstract] [Google Scholar]
  • Liu T.G., Peng Y.L., Chen W.Q. and Zhang Z.Y. (2010) First detection of virulence in Puccinia striiformis f. sp. tritici in China to resistance genes Yr24/Yr26 present in wheat cultivar Chuanmai 42. Plant Dis. 94, 1163. [Abstract] [Google Scholar]
  • Liu X.F., Huang C., Sun Z.Y., Liang J.M., Luo Y. and Ma Z.H. (2011) Analysis of population structure of Puccinia striiformis in Yunnan Province of China by using AFLP. Eur. J. Plant Pathol. 129, 43–55. [Google Scholar]
  • Loladze S., Druml T. and Wellings C.R. (2013) Temperature adaptation in Australasian populations of Puccinia striiformis f. sp. tritici . Plant Pathol. http://onlinelibrary.wiley.com/doi/10.1111/ppa.12132/pdf. [Google Scholar]
  • Long S.S., Cao Y.L. and Kang Z.S. (2010) Reactive oxygen species and cytosolic free calcium in the process of wheat resistance to stripe rust. Acta Phytopathol. Sin. 40, 265–275. [Google Scholar]
  • Lu S.Y., Fan G.F., Xie S.M., Wu W.Z., Kong X.L. and Yang Z.M. (1956) Studies on stripe rust of wheat I. Physiologic specialization of Puccinia glumarum (Schm.) Erikss. & Henn. Acta Phytopathol. Sin. 2, 153–166. [Google Scholar]
  • Lu Y., Lan C., Liang S., Zhou X., Liu D., Zhou G., Lu Q., Jing J., Wang M., Xia X. and He Z. (2009) QTL mapping for adult‐plant resistance to stripe rust in Italian common wheat cultivars Libellula and Strampelli. Theor. Appl. Genet. 119, 1349–1359. [Abstract] [Google Scholar]
  • Lupton F.G.H. and Macer R.C.F. (1962) Inheritance of resistance to yellow rust (Puccinia glumarum Erikss & Henn) in seven varieties of wheat. Trans. Br. Mycol. Soc. 45, 21–45. [Google Scholar]
  • Ma J.B., Huang X.L., Wang X.J., Chen X.M., Qu Z.P., Huang L.L. and Kang Z.S. (2009) Identification of expressed genes during compatible interaction between stripe rust (Puccinia striiformis) and wheat using a cDNA library. BMC Genomics, 10, 586–597. [Europe PMC free article] [Abstract] [Google Scholar]
  • Manners J.G. (1960) Puccinia striiformis Westend. var. dactylidis var. nov . Trans. Br. Mycol. Soc. 43, 65–68. [Google Scholar]
  • Manners J.G. (1988) Puccinia striiformis, yellow rust (stripe rust) of cereals and grasses. Adv. Plant Pathol. 6, 373–378. [Google Scholar]
  • Markell S.G. and Milus E.A. (2008) Emergence of a novel population of Puccinia striiformis f. sp. tritici in eastern United States. Phytopathology, 98, 632–639. [Abstract] [Google Scholar]
  • Mboup M., Leconte M., Gautier A., Wan A.M., Chen W.Q., de Vallavielle‐Pope C. and Enjalbert J. (2009) Evidence of genetic recombination in wheat yellow rust population of a Chinese over‐summering area. Fungal Genet. Biol. 46, 299–307. [Abstract] [Google Scholar]
  • McIntosh R.A., Wellings C.R. and Park R. (1995) Wheat Rusts: An Atlas of Resistance Genes. Dordrecht: Kluwer Academic Publishers. [Google Scholar]
  • McIntosh R.A., Yamazaki Y., Devos K.M., Dubcovsky J., Rogers W.J. and Appels R. (2003) Catalogue of gene symbols for wheat. In: Proceedings of the 10th International Wheat Genetic Symposium, Roma, Italy (Pogna N.E., editor; , Romanò M., editor; , Pogna E.A., editor; , Galterio G., editor. , eds), Vol. 4, SIMI, Rome, Italy. [Google Scholar]
  • McIntosh R.A., Hart G.E., Devos K.M., Rogers W.J., Morris C.F., Appels R. and Anderson O.D. (2007) Catalogue of gene symbols for wheat: 2007 supplement. Available at <http://www.shigen.nig.ac.jp/wheat/komugi/genes/symbolClassList.js> [accessed on Feb 10, 2014].
  • McNeal F.H., Konzak C.F., Smith E.P., Tate W.S. and Russell T.S. (1971) A uniform system for recording and processing cereal research data. USDA ARS 1–42, 34–121. [Google Scholar]
  • Melching J.S., Stanton J.R. and Koogle D.L. (1974) Deleterious effects of tobacco smoke on germination and infectivity of spores of Puccinia graminis tritici and on germination of spores of Puccinia striiformis, Pyricularia oryzae and an Alternaria species. Phytopathology, 64, 1143–1147. [Google Scholar]
  • Mendgen K. (1981) Nutrient uptake in rust fungi. Phytopathology, 71, 983–989. [Google Scholar]
  • Mendgen K. , Struck C. , Voegele R.T. and Hahn M. (2000) Biotrophy and rust haustoria. Physiol. Mol. Plant Pathol . 56 , 141–145. [Google Scholar]
  • Milus E.A., Kristensen K. and Hovmøller M.S. (2008) Increased aggressiveness of Puccinia striiformis f. sp tritici at least partially explains recent stripe rust epidemics. Phytopathology, 98, 107. [Abstract] [Google Scholar]
  • Moldenhauer J., Moerschbacher B.M. and Van der Westhuizen A.J. (2006) Histological investigation of stripe rust (Puccinia striiformis f. sp. tritici) development in resistant and susceptible wheat cultivars. Plant Pathol. 55, 469–474. [Google Scholar]
  • Morgounov A., Tufan H.A., Sharma R., Akin B., Bagci A., Braun H.J., Kaya Y., Kesser M., Payne T.S., Sonder K. and McIntosh R. (2012) Global incidence of wheat rusts and powdery mildew during 1969–2010 and durability of resistance of winter wheat variety Bezostaya 1. Eur. J. Plant Pathol. 132, 323–340. [Google Scholar]
  • Newton M. and Johnson T. (1936) Stripe rust, Puccinia glumarum in Canada. Can. J. Res. 14, 89–108. [Google Scholar]
  • Niu Y.C., Li Z.Q. and Shang H.S. (1991) Puccinia striiformis West. f. sp. leymi and f. sp. elymi, two new formae speciales. Acta Univ. Agric. Boreali Occident. 19, 58–62. [Google Scholar]
  • O'Brien L., Brown J.S., Young R.M. and Pascoe T. (1980) Occurrence and distribution of wheat stripe rust in Victoria and susceptibility of commercial wheat cultivars. Aust. Plant Pathol. 9, 14. [Google Scholar]
  • Park R.F. and Wellings C.R. (2012) Somatic hybridization in the Uredinales. Annu. Rev. Phytopathol. 50, 219–239. [Abstract] [Google Scholar]
  • Prashar M., Bhardwaj S.C., Jain S.K. and Datta D. (2007) Pathotypic evolution in Puccinia striiformis in India during 1995–2004. Aust. J. Agric. Res. 58, 602–604. [Google Scholar]
  • Prescott J.M., Burnett P.A., Saari E.E., Ransom J., Bowman J., Milliano W.de, Singh R.P. and Bekele G. (1986) Wheat Diseases and Pests: A Guide for Field Identification. Mexico, D. F.: International Maize and Wheat Improvement Center (CIMMYT). [Google Scholar]
  • Rapilly F. (1979) Yellow rust epidemiology. Annu. Rev. Phytopathol. 17, 59–73. [Google Scholar]
  • Rapilly F. and Foucault B. (1976) Premières études sur la rétention de spores fongiques par des épidermes foliaires. Ann. Phytopathol. 8, 31–40. [Google Scholar]
  • Rapilly F. and Fournet J. (1968) Observation sur la dissémination de Puccinia striiformis, en fonction de l'humidité relative, relation avec la structure morphologique des urédospores. Proceedings of the Cereal Rust Conference, Oeiras, Portugal: Tip. Anva rio ComercõÁal de Portugal, pp. 26–29. [Google Scholar]
  • Ren R.S., Wang M.N., Chen X.M. and Zhang Z.J. (2012) Characterization and molecular mapping of Yr52 for high‐temperature adult‐plant resistance to stripe rust in spring wheat germplasm PI 183527. Theor. Appl. Genet. 125, 847–857. [Abstract] [Google Scholar]
  • Röbbelen G. and Sharp E.L. (1978) Mode of inheritance, interaction and application of genes conditioning resistance to yellow rust. Fortschr. Pflanzenzucht. Beih. Z. Pflanzenzucht. 9, 1–88. [Google Scholar]
  • Schröder J. and Hassebrauk K. (1964) Undersuchungen uber die Keimung der Uredosporen des Gelbrostes (Puccinia striiformis West). Zentrab. Bakteriol. Parasitenk. Infektionskrank. Hyg. 118, 622–657. [Google Scholar]
  • Shan W.X., Chen S.Y., Wu L.R. and Li Z.Q. (1995) RAPD analysis of the epidemic races of Puccinia striiformis f. sp. tritici in China. Sci. Agric. Sin. 28, 1–7. [Google Scholar]
  • Sharp E.L. (1965) Prepenetration and postpenetration environment and development of Puccinia striiformis . Phytopathlogy, 55, 198–203. [Google Scholar]
  • Sharp E.L. (1967) Atmospheric ions and germination of urediospores of Puccinia striiformis . Science, 156, 1359–1360. [Abstract] [Google Scholar]
  • Sharp E.L. and Volin R.B. (1970) Additive genes in wheat conditioning resistance to stripe rust. Phytopathology, 60, 1146–1147. [Google Scholar]
  • Singh R.P. and Rajaram S. (1994) Genetics of adult plant resistance to stripe rust in ten spring bread wheats. Euphytica, 72, 1–7. [Google Scholar]
  • Singh R.P., Nelson J.C. and Sorrells M.E. (2000) Mapping Yr28 and other genes for resistance to stripe rust in wheat. Crop Sci. 40, 1148–1155. [Google Scholar]
  • Singh R.P., William H.M., Huerta‐Espino J. and Rosewarne G. (2004) Wheat rust in Asia: meeting the challenges with old and new technologies. In: New Directions for a Diverse Planet. Proceedings of the 4th International Crop Science Congress, Brisbane, Australia (Fischer T., editor; , Turner N., editor; , Angus J., editor; , McIntyre L., editor; , Robertson M., editor; , Borrell A., editor; and Lloyd D., editor. , eds), pp. 1–13. [Google Scholar]
  • Sohn J., Voegele R.T., Mendgen K. and Hahn M. (2000) High level activation of vitamin B1 biosynthesis genes in haustoria of the rust fungus Uromyces fabae . Mol. Plant–Microbe Interact. 13, 629–636. [Abstract] [Google Scholar]
  • Spielmeyer W. and Lagudah S. (2003) Homoeologous set of NBS‐LRR genes located at leaf and stripe rust resistance loci on short arms of chromosome 1 of wheat. Funct. Integr. Genomics, 3, 86–90. [Abstract] [Google Scholar]
  • Spielmeyer W., McIntosh R.A., Kolmer J. and Lagudah E.S. (2005) Powdery mildew resistance and Lr34/Yr18 genes for durable resistance to leaf and stripe rust cosegregate at a locus on the short arm of chromosome 7D of wheat. Theor. Appl. Genet. 111, 731–735. [Abstract] [Google Scholar]
  • Staples R.C. (2001) Nutrients for a rust fungus: the role of haustoria. Trends Plant Sci. 6, 196–198. [Abstract] [Google Scholar]
  • Struck C., Voegele R.T., Hahn M. and Mendgen K. (2004) Rust haustoria as sink in plant tissues or how to survive in leaves In: Biology of Plant–Microbe Interactions (Tikhonovich I., editor; , Lugtenberg B., editor; and Provorov N., editor. , eds), pp. 177–179. St. Paul, MN: International Society for Molecular Plant–Microbe Interactions. [Google Scholar]
  • Stubbs R.W. (1967) Investigations on yellow rust in wheat and barley, 1962–1965. In: Collected Papers on Cereal Disease (I). Sticht. Neth. Graan Centrum. Techn. Bericht. 17, 69–83. [Google Scholar]
  • Stubbs R.W. (1985) Stripe rust In: Cereal Rusts Diseases, Distribution, Epidemiology, and Control, Vol. II (Roelfs A.P., editor; and Bushnell W.R., editor. , eds), pp. 61–101. New York: Academic Press. [Google Scholar]
  • Stubbs R.W. (1988) Pathogenicity analysis of yellow (stripe) rust of wheat and its significance in a global context In: Breeding Strategies for Resistance to the Rusts of Wheat (Simmonds N.W., editor; and Rajaram S., editor. , eds), pp. 23–38. Mexico, D. F.: International Maize and Wheat Improvement Center (CIMMYT). [Google Scholar]
  • Suenaga K., Singh R.P., Huerta‐Espino J. and William H.M. (2003) Microsatellite markers for genes Lr34/Yr18 and other quantitative trait loci for leaf and stripe rust resistance in bread wheat. Phytopathology, 93, 881–890. [Abstract] [Google Scholar]
  • Tanksley S.D. (1993) Mapping polygenes. Annu. Rev. Genet. 27, 205–233. [Abstract] [Google Scholar]
  • Tian H.M., Liu T.G., Gao L. and Chen W.Q. (2008) Parasitic fitness of four major prevalent races of Puccinia striiformis f. sp. tritici in China. Acta Phytopathol. Sin. 38, 599–606. [Google Scholar]
  • Tollenaar H. (1967) A comparison of Puccinia striiformis f. sp. poae on bluegrass with P. striiformis f. sp. tritici and f. sp. dactylidis . Phytopathology, 57, 418–420. [Google Scholar]
  • Tu J.C. and Hendrix J.W. (1967) The summer biology of Puccinia striiformis in southeastern Washington. I. Induction of infection during the summer. Plant Dis. Rep. 51, 911–914. [Google Scholar]
  • Uauy C., Brevis J.C., Chen X.M., Khan I., Jackson L., Chicaiza O., Distelfeld A., Fahima T. and Dubcovsky J. (2005) High‐temperature adult‐plant (HTAP) stripe rust resistance gene Yr36 from Triticum turgidum ssp. dicoccoides is closely linked to the grain protein content locus Gpc‐B1. Theor. Appl. Genet. 112, 97–105. [Abstract] [Google Scholar]
  • Vallavieille‐Pope C.de, Huber L., Leconte M. and Goyeau H. (1995) Comparative effects of temperature and interrupted wet periods on germination, penetration, and infection of Puccinia recondita f. sp. tritici and P. striiformis on wheat seedling. Phytopathology, 85, 409–415. [Google Scholar]
  • Vallavielle‐Pope C.de, Ali S., Enjalbert J., Leconte M., Duan X.Y., Chen W.Q., Justesen A.F. and Hovmøller M. (2012) Diversity of Puccinia striiformis population at global scale In: Disease Risk and Food Security. Proceedings of the 13th International Cereal Rust and Powdery Mildew Conference (Chen W.Q., editor. et al, eds), pp. 3–4. Beijing: China Agricultural Science and Technology Press. [Google Scholar]
  • Vazquez M.D., Peterson C.J., Riera‐Lizarazu O., Chen X.M., Heesacker A., Ammar K., Crossa J. and Mundt C.C. (2012) Genetic analysis of adult plant, quantitative resistance to stripe rust in wheat cultivar Stephens in multi‐environment trials. Theor. Appl. Genet. 124, 1–11. [Abstract] [Google Scholar]
  • Voegele R.T. (2006) Uromyces fabae: development, metabolism, and interactions with its host Vicia faba . FEMS Microbiol. Lett. 259, 165–173. [Abstract] [Google Scholar]
  • Voegele R.T. and Mendgen K. (2003) Rust haustoria: nutrient uptake and beyond. New Phytol. 159, 93–100. [Abstract] [Google Scholar]
  • Wan A.M. and Chen X.M. (2012) Stripe rust epidemics of wheat and barley and races of Puccinia striiformis identified in the United States in 2011. Phytopathology, 102 (S4), 130. [Google Scholar]
  • Wan A.M., Zhao Z.H., Chen X.M., He Z.H., Jin S.L., Jia Q.Z., Yao G., Yang J., Wang B., Li G., Bi Y. and Yuan Z. (2004) Wheat stripe rust epidemic and virulence of Puccinia striiformis f. sp tritici in China in 2002. Plant Dis. 88, 896–904. [Abstract] [Google Scholar]
  • Wang M.N. and Chen X.M. (2013) First report of Oregon grape (Mahonia aquifolium) as an alternate host for the wheat stripe rust pathogen (Puccinia striiformis f. sp. tritici) under artificial inoculation. Plant Dis. 97, 839. [Abstract] [Google Scholar]
  • Wang X.J., Tang C.L., Zhang G., Li Y.C., Wang C.F., Liu B., Qu Z.P., Zhao J., Han Q.M., Huang L.L., Chen X.M. and Kang Z.S. (2009) cDNA‐AFLP analysis reveals differential gene expression in compatible reaction of wheat challenged with Puccinia striiformis f. sp. tritici . BMC Genomics, 10, 289–300. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wang X.J., Tang C.L., Deng L., Cai G.L., Liu X.Y., Liu B., Han Q.M., Buchenauer H., Wei G.R., Han D.J., Huang L.L. and Kang Z.S. (2010) Characterization of a pathogenesis‐related thaumatin‐like protein gene TaPR5 from wheat induced by stripe rust fungus. Physiol. Plant. 139, 27–38. [Abstract] [Google Scholar]
  • Wang Y., Wang M.N., Zhang R.J., Xu C., Feng Q. and Li Z.Q. (2006) Transient expression of transformed genes by particle bombardment. Acta Bot. Boreal Occident Sin. 26, 1115–1118. [Google Scholar]
  • Wellings C.R. (2007) Puccinia striiformis in Australia: a review of the incursion, evolution, and adaptation of stripe rust in the period 1979–2006. Aust. J. Agric. Res. 58, 567–575. [Google Scholar]
  • Wellings C.R. (2011) Global status of stripe rust: a review of historical and current threats. Euphytica, 179, 129–141. [Google Scholar]
  • Wellings C.R. , Boyd L.A. and Chen X.M. (2012) Resistance to stripe rust in wheat: pathogen biology driving resistance breeding In: Disease Resistance in Wheat (Sharma I., editor. , ed.), pp. 63–83. London: CAB International. [Google Scholar]
  • Wellings C.R. and McIntosh R.A. (1990) Puccinia striiformis f. sp. tritici in Australasia: pathogenic changes during the first 10 years. Plant Pathol. 39, 316–325. [Google Scholar]
  • Wellings C.R., McIntosh R.A. and Walker J. (1987) Puccinia striiformis f. sp. tritici in eastern Australia—possible means of entry and implications for plant quarantine. Plant Pathol. 36, 239–241. [Google Scholar]
  • Wellings C.R., McIntosh R.A. and Hussain M. (1988) A new source of resistance to Puccinia striiformis f. sp. tritici in spring wheats (Triticum aestivum). Plant Breed. 100, 88–96. [Google Scholar]
  • William H.M., Singh R.P., Huerta‐Espino J., Ortiz Islas S. and Hoisington D. (2003) Molecular marker mapping of leaf rust resistance gene Lr46 and its association with stripe rust resistance gene Yr29 in wheat. Phytopathology, 93, 153–159. [Abstract] [Google Scholar]
  • Xia N., Zhang G., Liu X.Y., Deng L., Cai G.L., Zhang Y., Wang X.J., Zhao J., Huang L.L. and Kang Z.S. (2010) Characterization of a novel wheat NAC transcription factor gene involved in defense response against stripe rust pathogen infection and abiotic stresses. Mol. Biol. Rep. 37, 3703–3712 [Epub: 7 March 2010]. [Abstract] [Google Scholar]
  • Xie S.X., Chen W.Q., Chen Y.L., Wang K.N., Lin X.L. and Yang Z.H. (1992) Study on the dispersal of wheat stripe rust (Puccinia striiformis West.) in Longnan prefecture and Abazhou. Acta Phytopathol. Sin. 22, 138–143. [Google Scholar]
  • Xie S.X., Wang K.N., Chen Y.L. and Chen W.Q. (1993) Preliminary study on the relationship between transport of wheat stripe rust and upper air current in China. Acta Phytopathol. Sin. 23, 203–209. [Google Scholar]
  • Xu L.S., Wang M.N., Cheng P., Kang Z.S., Hulbert S.H. and Chen X.M. (2013) Molecular mapping of Yr53: a new gene for stripe rust resistance in durum wheat accession PI 480148 and its transfer to common wheat. Theor. Appl. Genet. 126, 523–533. [Abstract] [Google Scholar]
  • Yin C.T., Chen X.M., Wang X.J., Han Q.M., Kang Z.S. and Hulbert S. (2009) Generation and analysis of expression sequence tags from haustoria of the wheat stripe rust fungus Puccinia striiformis f. sp. tritici . BMC Genomics, 10, 626. [Europe PMC free article] [Abstract] [Google Scholar]
  • Yu X.D., Qu Z.P., Guo J., Yu X.M., Huang X.L., Han Q.M., Huang L.L. and Kang Z.S. (2008) Construction and preliminary analysis of the compatible SSH cDNA library between Puccinia striiformis and wheat. Sci. Agric. Sin. 41, 1267–1273. [Google Scholar]
  • Zadoks J.C. (1961) Yellow rust on wheat: studies in epidemiology and physiologic specialization. Tijdschr. Plantenziekten. 67, 69–256. [Google Scholar]
  • Zeng S.M. and Luo Y. (2006) Long‐distance spread and interregional epidemics of wheat stripe rust in China. Plant Dis. 90, 980–988. [Abstract] [Google Scholar]
  • Zhang Y., Zhang G., Xia N., Wang X.J., Huang L.L. and Kang Z.S. (2009) Cloning and characterization of a bZIP transcription factor gene in wheat and its expression in response to stripe rust pathogen infection and abiotic stresses. Physiol. Mol. Plant Pathol. 73, 88–94. [Google Scholar]
  • Zhang Y.H., Qu Z.P., Zheng W.M., Liu B., Wang X.J., Xue X.D., Xu L.S., Huang L.L., Han Q.M., Zhao J. and Kang Z.S. (2008) Stage‐specific gene expression during urediniospore germination in Puccinia striiformis f. sp. tritici . BMC Genomics, 9, 203–212. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhao J., Zhang H., Yao J., Huang L. and Kang Z. (2011) Confirmation of Berberis spp. as alternate hosts of Puccinia striiformis f. sp. tritici on wheat in China. Mygosystema, 30, 895–900. [Google Scholar]
  • Zhao J., Wang L., Wang Z., Chen X., Zhang H., Yao J., Zhan G., Chen W., Huang L. and Kang Z. (2013) Identification of eighteen Berberis species as alternate hosts of Puccinia striiformis f. sp. tritici and virulence variation in the pathogen isolates from natural infection of barberry plants in China. Phytopathology, 103, 927–934. [Abstract] [Google Scholar]
  • Zheng W.M., Chen S.Y., Kang Z.S., Wang Y., Wu L.R. and Li Z.Q. (2005) DNA fingerprinting of natural population of Puccinia striiformis f. sp. tritici in Tianshu area. Mygosystema, 24, 199–206. [Google Scholar]

Articles from Molecular Plant Pathology are provided here courtesy of Wiley-Blackwell

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/2007563
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/2007563

Article citations


Go to all (117) article citations