Next Article in Journal
Influences of the Culturing Media in the Virulence and Cell Wall of Sporothrix schenckii, Sporothrix brasiliensis, and Sporothrix globosa
Next Article in Special Issue
Bioactive Metabolites from the Endophytic Fungus Aspergillus sp. SPH2
Previous Article in Journal
Amphotericin B and Other Polyenes—Discovery, Clinical Use, Mode of Action and Drug Resistance
Previous Article in Special Issue
Space and Vine Cultivar Interact to Determine the Arbuscular Mycorrhizal Fungal Community Composition
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epichloë Fungal Endophytes—From a Biological Curiosity in Wild Grasses to an Essential Component of Resilient High Performing Ryegrass and Fescue Pastures

1
Grasslanz Technology Ltd., Palmerston North PB11008, New Zealand
2
AgResearch Ltd., Palmerston North PB11008, New Zealand
*
Author to whom correspondence should be addressed.
J. Fungi 2020, 6(4), 322; https://doi.org/10.3390/jof6040322
Submission received: 28 September 2020 / Revised: 13 November 2020 / Accepted: 18 November 2020 / Published: 27 November 2020
(This article belongs to the Special Issue Fungal Endophytes in Agriculture and Ecosystems)

Abstract

:
The relationship between Epichloë endophytes found in a wide range of temperate grasses spans the continuum from antagonistic to mutualistic. The diversity of asexual mutualistic types can be characterised by the types of alkaloids they produce in planta. Some of these are responsible for detrimental health and welfare issues of ruminants when consumed, while others protect the host plant from insect pests and pathogens. In many temperate regions they are an essential component of high producing resilient tall fescue and ryegrass swards. This obligate mutualism between fungus and host is a seed-borne technology that has resulted in several commercial products being used with high uptake rates by end-user farmers, particularly in New Zealand and to a lesser extent Australia and USA. However, this has not happened by chance. It has been reliant on multi-disciplinary research teams undertaking excellent science to understand the taxonomic relationships of these endophytes, their life cycle, symbiosis regulation at both the cellular and molecular level, and the impact of secondary metabolites, including an understanding of their mammalian toxicity and bioactivity against insects and pathogens. Additionally, agronomic trials and seed biology studies of these microbes have all contributed to the delivery of robust and efficacious products. The supply chain from science, through seed companies and retailers to the end-user farmer needs to be well resourced providing convincing information on the efficacy and ensuring effective quality control to result in a strong uptake of these Epichloë endophyte technologies in pastoral agriculture.

1. Introduction

Plants and microbes have long been recognised to co-exist in a symbiotic relationship, and in some cases, they are known to provide benefit to each other in a mutualistic interaction. Some of these microbes have provided technologies that can and have been used as commercial products. This includes rhizobium isolates for improved nitrogen fixation [1], arbuscular mycorrhiza for improved water and nutrient acquisition [2], and Epichloë fungal strains for improved animal health and welfare while ensuring grass plant resistance/tolerance to biotic and abiotic stresses [3,4,5,6]. Indeed, there is a view that microbial endophytes have an important role in maintaining productivity levels in environmentally sustainable agricultural systems [7].
Microorganisms are extremely diverse and can exhibit many different biological behaviours related to their symbiotic lifestyle, which allows some of them to function as effective plant protection agents. These differences relate to the type of symbiotic relationship they form with their hosts (mutualistic vs. commensalistic); in planta colonisation patterns (systemic vs. point infections); their level of host-specificity (low vs. high); means of propagation (horizontal vs. vertical); and endophytic lifestyle (obligate vs. facultative) [8]. The symbioses of the Epichloë fungal species with host grasses, of the family Pooidae [9,10], can span the continuum from antagonistic to commensal or mutualistic [11], but here, the focus will be largely on the asexual mutualistic types.
Asexual Epichloë endophytes exhibit the characteristics of mutualism, systemic infection, high host specificity, vertical (maternal) transfer, and an obligate lifestyle [8] that in many ways make this microbial technology unique and in part explains why as a commercial product, it has been so successful [12]. They are known to produce a large range of secondary metabolites of which the alkaloids are the most well characterised [3,13]. The aim here is to review this mutualistic relationship to determine (1) the origins of Epichloë strain variation, (2) reasons for its importance in many temperate grass pastures, (3) methods of managing its negative and positive characteristics, (4) how effective delivery of commercial Epichloë technologies has been achieved, and (5) how further research opportunities can continue to add value to this economically important relationship, which underpins sustainable pastoral farming practices in managed temperate grasslands.

2. Epichloë Endophytes

2.1. Epichloë Taxonomy

The Epichloë genus contains two major categories of fungal organisms, such that of the 43 documented Epichloë taxa associated with grasses (Table 1) [14], 14 are known to develop sexual structures with viable ascospores, while for the other 29 taxa the sexual state has not been observed [15]. Prior to 2014 the Epichloë genus contained only the sexual forms (teleomorph), but now also contains the asexual forms (anamorph), which had previously been classified as Neotyphodium [15], and prior to that, Acremonium [16]. This change resulted from a requirement that a single genus name is to be used for all stages of the development of a fungal species [17].

2.2. Epichloë Diversity and Origins

Epichloë endophytes have been found in more than 100 grass species, which have evolved in most temperate regions of the world (Table 1) [18,19]. However, it is acknowledged that endophyte infection is rare in grasses endemic to Australasia [20], and sub-Saharan Africa [21] in comparison to the wide range of infection found in the wild, uncultivated grasses of the Northern Hemisphere and in particular Europe [5] and Asia [22]. Indeed, the most important temperate grass species from an economic viewpoint, namely Lolium and Festuca species, have originated in Europe and North Africa [23]. In South America, most collections of grasses containing Epichloë have been made in the Patagonian steppe [24]. While modern cereals are not naturally infected by Epichloë, a range of their progenitor species in the genera Elymus and Hordeum are frequently infected [25]. However, when Epichloë strains from the wild grasses have been inoculated into rye (Secale cereale), individual genetically distinct host genotypes show morphological phenotypes that range from heavily stunted through to some that resemble healthy uninfected plants [26,27]. Epichloë infection has also been achieved into the wheat genome using Chinese spring wheat substitution lines [28].
Table 1. Infection of Epichloë species in temperate grasses by region – for more extensive and detailed listings [14,15,22,29,30,31,32].
Table 1. Infection of Epichloë species in temperate grasses by region – for more extensive and detailed listings [14,15,22,29,30,31,32].
Grass GenusEpichloë SpeciesReference
Europe/North Africa
Lolium canarienseE. typhinum var. canariense[33]
Lolium multiflorumE. occultans
Lolium perenneE. hybrida[34]
Lolium rigidumE. occultans[15]
Agropyron repensE. bromicola[35]
AgrostisE. baconii, E. amarillans[32]
AnthoxanthumE. typhina
BrachyelytrumE. brahyelytri
BrachypodiumE. sylvatica, E. typhina
Dactylis glomerataE. typhina
ElymusE. elymi
Festuca arundinaceaE. coenophialum
Festuca giganteus, Festuca rubraE. festucae
GlyceriaE. glyceriae
HolcusE. clarkii
Leymus, BromusE. bromicola
L. perenneE. festucae var. lolii, E. typhina, E. lolii
PhleumE. typhina
PoaE. typhina
SphenopholisE. amarilians
Festuca pratensisE. uncinatum
E. siegelii[36]
HordelymusE. disjuncta, E. danica, E. hordelymi, E. sylvatica subsp, pollinensisi,[15,37]
Holcus mollisE. mollis[15,38]
Asia
AchnatherumE. ganusuensis, E. sibirica[22]
E. chisosum; E. inebrians[29,39]
E. funkii[15]
Brachypodium, Bromus, Elymus, LeymusE. bromicola[22]
CalamagrostisE. stromatolonga
FestucaE. sinofestucae
Elymus, Elytrigia, Festuca, Hordeum, Poa, Roegneria, StipaE. spp.
PoaE. liyangensis[40]
RoegneriaE. sinica[22]
E. yangzii[41]
North America
AmmophilaE. amarillans[42]
Brachyelytrum erectumE. brachyelytri[11]
Bromus laevipesE. cabralii, E. spp.[43]
Cinna arundinaceaE. schardlii[44]
ElymusE. elymi[11]
Elymus canadensisE. canadensis[15,45]
Festuca arizonicaE. huerfanum, E. tembladerae[29]
Glyceria striataE. glyceriae[11]
Poa alsodesE. alsodes[46]
Poa secunda subsp. junicoliaE. poae[31]
South America
Bromus setifoliusE. typhina var. aonikenhana[47]
E. typhinum[48]
E. tembladerae[15]
Bromus auleticusE. pampeana; E. tembladerae
Festuca argentina, F. hieronymi. Poa huecuE. tembladerae[49]
Hordeum comosumE. tembladerae, E. amarillans, E. typhina hybrids[24]
Melica ciliataE. guerinii[15]
Melica decumbensE. melicicola[29,50]
Phleum alpinumE. cabralii[47]
E. tembladerae[15]
Poa, Briza, Festuca, Melica, PhleumE. tembladerae, E. pampeana[50,51,52]
Australia
Echinopogon spp.E. australiense[50,53]
New Zealand
Echinopogon ovatusE. aotearoa[50]
Dichelachne micranthaE. australiensis[20]
Poa matthewsiiE. novae-zelandiae
Sub-Saharan Africa
Festuca costataE. spp.[21]
Melica spp.E. melicicola[50]
Epichloë strains have been classified as either hybrid (being the result of a cross between two or more species) [54,55] or non-hybrid. While hybrids have interspecific origins, there is one known exception, E. schardlii, which has resulted from intraspecific hybridisation [15,44]. At least half of all known Epichloë species are hybrid types [10,15,29,56] and with one rare exception [40], all hybrid species are asexual [48,57]. However, that does not mean that non-hybrid types are necessarily capable of sexual reproduction [37]. Interspecific hybridisation most likely occurs via somatic cell fusion followed by fusion of nuclei [54,56]. Epichloë is notable for having more interspecific hybrids than any other fungal genus [34]. Whereas horizontally transmissible species have haploid genomes, producing ascospores [58], most of the strictly seedborne mutualists, such as most Epichloë species, are hybrids with heteroploid (aneuploid or polyploid) genomes [29]. Yet even some of these can form epithelial growth that produce conidia [31,59] with the potential to horizontally transmit, but the dominant and more successful form of transmission is still vertical transmission through the host plant seed [10]. Direct infection by germinating conidia has not been documented [60].
At least in some instances, hybridisation came after the strain became seedborne rather than being caused by the seedborne habit suggesting a selective advantage of hybridization for the mutualistic endophytes. Hybrids are likely to contain more genetic variation, which may lead to improved adaptation to biotic and abiotic stresses of their host plants [10,29,30,56,61]. There is also a general hypothesis that interspecific hybridisation provides greater genetic variation and hence, a wider adaptation range in stressful environments than intraspecific hybridisation [56,62]. However, when comparing hybrid and non-hybrid Epichloë strains on controlled environments, there is no evidence of niche expansion of Epichloë hybrid-infected plants [63]. They also showed that non-hybrid endophytes increased seed production of their hosts, whereas hybrid endophytes reduced it, suggesting a fitness advantage for plants hosting non-hybrid endophytes.
Diversity within the Epichloë genus can be characterised by the types of alkaloids they produce in planta [3,64,65]. Four major classes of alkaloids are known to be produced by Epichloë strains. These include lolines (saturated 1-aminopyrrolizidines), indole diterpenes (lolitrems, epoxyjanthitrems), ergot alkaloids (main terminal product is ergovaline), and peramine (a pyrrolopyrazine alkaloid) [30,66]. Naturally occurring strains of Epichloë may produce from none to all four types of these known alkaloids. Additionally, most of the secondary metabolite pathways that result in producing the known chemistry are complex and have many intermediate compounds, some of which have been shown to have bioactivity [3,67]. There is still a considerable amount of unknown bioactivity associated with Epichloë endophytes and conversely, there are known secondary metabolites with undescribed or putative functions. Epichloë strains AR48 and AR47, for example, have been shown to control cutworm moth caterpillar (Agrotis ipsilion), but the alkaloid associated with that control is as yet unknown [68]. Whereas, examples of the latter are the non-alkaloid compounds epichloecyclins, which are cyclic ribosomally synthesized and post translationally modified peptides (RiPPs) with no known function [69] and a hybrid peptide-polyketide named Dahurelmusin A with only putative insecticidal activity [70]. While it is the endophyte strain that carries the genes required for alkaloid expression it is unknown factors associated with the host genetics [71,72,73], including the expression of plant hormones [74], that moderate alkaloid expression. Alkaloid expression levels can be further modified quantitatively by the environment [75,76,77,78,79,80]. These alkaloids are either not expressed or at very low levels when Epichloë is grown in axenic culture, but are highly expressed in planta [81,82,83,84]. The epigenetic regulation of the ergot alkaloids and lolitrems via chromatin remodelling also plays a critical role in the symbiosis-specific expression of these alkaloid pathways [81,85,86].
Distribution of alkaloids can vary within the plant and they are not necessarily correlated with the distribution of fungal hyphae associated with the Epichloë endophyte [87]. In perennial ryegrass, lolitrem B accumulates in older tissues, ergovaline is concentrated in the stem and basal leaf sheath of intermediate age tillers, and peramine is evenly distributed across all leaf tissues [88,89]. For flowering ryegrass plants, the seed component contains about 75% of the total peramine present in the plant [90]. In fescue plants, loline can be found in both the shoot and root tissue [91,92]. In shoot tissue, the highest levels of loline occur in the inflorescence, followed by meristem and then pseudostem [93]. The highest peramine concentrations have been found in young leaves of meadow fescue in early spring and in panicles (spikelets, seeds) and leaf pseudostems during the period of vegetative growth in late summer and autumn [94].

2.3. Epichloë Mutualism

Mutualism occurs when each participant receives a net benefit from the association [95,96]. Epichloë endophytes can form mutualistic symbiotic associations [97,98,99] within the aerial tissues of some temperate cool-season grasses of the subfamily Pooideae [26,55]. Within this subfamily, 50% of the 14 tribes have species that host Epichloë [14,55] (Table 1). Discoveries mostly over the last decade, have revealed dynamic and complex cellular and molecular responses critical for establishing and maintaining mutualistic symbiotic interactions (previously reviewed [99]). These include nutrient related processes such as regulation of apoplastic iron homeostasis [100,101,102], epigenetic regulation [81,86], and signalling pathways such as Nox produced reactive oxygen species (ROS) [103], calcineurin signalling [104], lipid signalling [105], G protein and adenosine 3′, 5′ -cyclic monophosphate (cAMP)/cAMP-dependent protein kinase (PKA) signalling [106,107], stress-activated mitogen-activated protein (MAP) kinase pathway [108], and the cell wall integrity (CWI) mitogen-activated protein kinase (MAPK) pathway [109]. Transcriptomic studies indicate that symbiosis establishment requires significant host reprogramming with genes associated with photosynthesis, stress, plant hormone biosynthesis and perception, cell membrane regulation, and plant defence [110,111,112,113,114].

2.4. Epichloë Systemic Infection

Epichloë systemically infect plant tissues [115,116,117] but are only found in the aerial parts of grass plants. Establishment of infection requires colonisation of the meristematic tissues of the shoot apex, which occurs by extensive hyphal branching [118]. To systemically colonise aerial tissues, hyphae grow between leaf cells and as the leaf extends, hyphae attached to host cell walls commits the hyphae to grow by intercalary expansion (so that hyphal filament length increases as the leaves expand) to avoid breakage in a manner that is highly regulated and synchronised with host leaf expansion [118,119].

2.5. Epichloë Host Specificity

The Epichloë fungus has co-evolved with it host grass over millennia [120] to the point where the genome of Epichloë has genes for improved host compatibility [121]. Moving Epichloë strains across grass taxa has been difficult and largely unsuccessful, suggesting that Epichloë species and even some strains have developed through co-speciation and are essentially host species specific [41,122]. Strong host specificity of Epichloë endophytes is related to both host species and their provenance [123].

2.6. Epichloë Vertical Transmission

Vertical transmission of Epichloë through host seeds [124] is a critical element that allows the transfer of the endophyte to successive generations through seed production processes and delivery to end user pastoral farmers. It has been hypothesised with good evidence that vertical transmission results in enhanced capability of host protection [30]. The success of vertical transmission can depend on the compatibility of the endophyte strain with the host genetics. In seed produced from natural associations, the fungus can be associated in seed at close to 100% [125], however in Europe where Epichloë co-evolved along with ryegrass and tall fescue, rates can be lower [126]. The reduced rate is thought to be due to the endophyte not necessarily being beneficial for the host plants in all environments [127,128] and/or an imperfect spread to all tillers of the plant resulting in the lack of transmission through seed [129], or reduced viability of the endophyte in seed [130,131]. For novel associations created by moving endophyte strains into new host germplasm, the rate of transmission can be much lower [132,133], although it has been possible to use host plant selection to improve the transmission rate, showing the importance of host plant genetics [134] for vertical transmission.
While asexual Epichloë endophytes are obligate with no free living form in nature, they are totally reliant on their host plant for survival and can rapidly lose viability when seed is stored at high temperatures and high humidity [135], and over about 6 months if stored at ambient temperatures [136]. To maintain endophyte viability in seed, storage at low temperatures (<5 °C) and low relative humidity (<60%) is recommended [130].

3. Impact of Epichloë Endophytes in Pastoral Systems

3.1. Animal Health and Welfare

Epichloë endophytes were primarily discovered as a result of animal health and welfare issues caused by alkaloids resulting from the mutualistic association, namely in tall fescue [137,138] and ryegrass [139]. Epichloë in tall fescue was shown to be associated with a condition in the USA known as fescue toxicosis [140], which has been estimated to create production losses of about US$1 billion per year [141]. This was particularly evident in cattle and dairy cows [142,143,144], largely because they were the most commonly used grazing animal in the USA, but it also occurs with sheep [145,146,147], goats [148], horses [149], deer [150], and alpacas [151]. The offending alkaloid causing fescue toxicosis has been identified to be ergovaline [152] which in the rumen, breaks down to lysergic acid [153], but a range of other ergot alkaloids may be implicated [154,155] (Table 2).
For perennial ryegrass the presence of Epichloë was associated with ryegrass staggers in New Zealand [140,156,157] caused by the alkaloid lolitrem B [158], although this condition was recorded many years before that [159]. Lolitrem B, a lipophilic compound, is a neurotoxin that affects muscular coordination resulting in tremors [152,158]. It also impacts on respiratory, cardiovascular, and digestive systems [160]. There are many lolitrems that have been characterised and labelled by a letter (A to N) and differ by the presence or absence of an I ring and the number of hydroxyl and aryl substitutions [161]. The tremorgenic properties of these lolitrem compounds can vary considerably (Table 2).
However, for the Epichloë association with ryegrass the presence of ergovaline can cause increases in body temperature [162,163] and respiration rate [163,164] of sheep and cattle. Comparisons of sheep grazed on endophyte free and endophyte infected ryegrass showed that the impact of Epichloë endophyte was much greater than just causing stagger events [165]. Also evident were reductions in daily liveweight gains and plasma prolactin, and increased presence of daggs, incidence of flystrike, and rectal temperatures (Table 3).
In Australia, the presence of Epichloë endophytes in perennial ryegrass causes a condition termed “perennial ryegrass toxicosis”, which has been attributed to the expression of both ergovaline and lolitrem B [156,166]. A severe perennial ryegrass toxicosis epidemic, which occurred in 2002, resulted in an estimated 100,000 sheep deaths.
While much is known about the toxic effects of ergovaline and lolitrem B less is known about the impact of other alkaloids associated with Epichloë infection [13]. A summary of known impacts of Epichloë-associated alkaloids on animal health and welfare is provided in Table 2. Many alkaloids also accumulate in the seed [88], acting as feeding deterrents for birds and rodents [167].
Lolines [168] and peramine [84,169] alkaloids are considered not toxic to grazing animals (Table 2). Peramine is unique and not known outside of the Epichloë genus [82,170]. For meadow fescue and tall fescue it is possible to identify endophyte isolates inducing the production of zero, low, or high loline concentrations, while for perennial ryegrass, endophytes strains have not been found that express loline [171]. Up to seven types of loline have been shown to be expressed by Epichloë endophytes in fescues, with N-formylloline (NFL) and N-acetylloline (NAL) being the most abundant [172] and along with N-acetyl norloline (NANL) the most bioactive [173]. There has been a report of loline and, in particular NANL causing equine fescue oedema [174], but further more detailed and thorough work has shown this is not the case and that lolines or NANL are unlikely to be the causative agent of this disease [175]. Lolines are extensively metabolised in the digestive tract of sheep prior to absorption and/or in the liver or other tissues following absorption resulting in low levels of excretion in urine and faeces [176].
Table 2. Documented effects of alkaloids expressed by Epichloë on animal health and welfare.
Table 2. Documented effects of alkaloids expressed by Epichloë on animal health and welfare.
AlkaloidAnimal EffectAction and Qualifying InformationReference
Ergot Alkaloids [177]
ChanoclavineNo toxic effects at levels found in grassesMay lower prolactin serum levels at high concentrations[178,179]
DehydroergovalineMay contribute to toxicityPresent only in fescue[13]
ErgineStuporHigh levels in Stipa robusta and Achnatherum inebrians[13,180,181]
ErgocornineFescue toxicosisIntermediate in vasoconstriction between ergovaline and lysergic acid[154]
Ergocristine
Ergocryptine
Ergonovine
ErgonovineFescue toxicosisLowered skin temperature, heart rate, and prolactin and had a higher respiration rate and blood pressure[182]
ErgotamineFescue toxicosisSimilar vasoconstriction effect as ergovaline[183]
ErgotamineFescue toxicosisFever, diarrhoea, weight loss, laboured breathing, salivation, low prolactin[182,184]
Ergosine
Agroclavine
ErgovalineFescue toxicosis/fescue footInability to regulate body temperature; vasoconstrictor; regulates prolactin[143,152,182,185,186,187]
Heat stressIncreased body temperature[146,188]
Lysergic acidFescue toxicosisLysergic acid is a major breakdown compound from ergovaline in rumen[153,189]
1000 times less potent than ergovaline as a vasoconstrictor[183,190]
Indole-Diterpenoids
EpoxyjanthitremsStaggersCan be intense but short lived[191,192]
Lolilline Not tremorgenic [193]
Lolitrems A, B, and FRyegrass staggersNeurotoxin that affects muscular coordination; delayed onset but persistent; marked increases in respiration rate, heart rate, and blood pressure.[84,152,158,160,193,194,195,196,197,198]
31-epi-Lolitrem BNot tremorgenic-[193]
Lolitrem EMinor tremorgenInhibitor of mitotic kinesin (Eg5)[199,200]
LolitriolNot tremorgenic-[201]
PaspalineNot tremorgenic-[198]
PaxillineModerate tremorgenFast acting but short longevity; marked increases in respiration rate, heart rate. and blood pressure.[160,201,202,203,204,205]
Terpendole CTremorgenFast acting, intense but short lived[206]
Terpendole MMild tremorgenShort lived[207]
Pyrrolopyrazine Alkaloid
PeramineNo known mammalian toxicityPossible association with causing diarrhoea, but later proven incorrect[169,208,209]
Pyrrolizidine Alkaloids [175]
N-acetyl loline (NAL)No known mammalian toxicity-[168,175,210]
N-acetylnorloline (NANL)No consistent mammalian toxicity-[168,174,175,210]
N-formyl loline (NFL)No known mammalian toxicity-[168,175,210]
Table 3. The productivity and health of young sheep (30 per treatment) grazing either endophyte-free or endophyte-infected perennial ryegrass during summer and autumn periods between 1992 and 1995 on unirrigated pasture in Canterbury, New Zealand. (Taken from [165]).
Table 3. The productivity and health of young sheep (30 per treatment) grazing either endophyte-free or endophyte-infected perennial ryegrass during summer and autumn periods between 1992 and 1995 on unirrigated pasture in Canterbury, New Zealand. (Taken from [165]).
Animal TraitEndophyte-FreeEndophyte-Infected (Standard Strain)Level of Significant Difference
Daily liveweight gain (g/head/d)5230**
Ryegrass staggers score (0–5 scale)03.3**
Dags score (0–5 Scale)0.32.3**
Flystrike (% affected)215**
Rectal temperature (°C)40.240.5*
Plasma prolactin (ng/mL)19890**
** p < 0.01; * p < 0.05.

3.2. Plant Persistence and Yield

The association between Epichloë endophyte presence that resulted in animal health and welfare issues led to the logical conclusion that Epichloë endophytes were problematic and needed to be removed from grasses. This was easily achieved because it was found that Epichloë strain survival in seed was negatively impacted by high temperatures and humidity [211]. The removal of Epichloë endophytes from sown pasture quickly led to the discovery that Epichloë endophytes were required for grass persistence through providing resistance/tolerance to both biotic and abiotic stresses [212,213,214,215,216]. The presence of Epichloë endophytes in leaf material can also increase the tolerance of grasses to herbivory [217].

3.3. Epichloë Effects on Abiotic Stresses

Epichloë endophytes have been demonstrated to improve drought tolerance in tall fescue [218,219,220,221,222,223,224,225,226,227], perennial ryegrass [228,229], and Agrostis [230]. However, other studies have shown no benefit of endophyte infection on drought tolerance of grasses [224,231]. It has been proposed with good evidence that interactions between plant genotype and fungal endophyte strain may explain inconsistent responses to drought due to endophyte infection [219,232,233,234,235,236,237,238,239]. Other abiotic stresses that influence plant growth and persistence that have been to some extent ameliorated by Epichloë endophytes include salinity [240,241,242], improved phosphorus uptake from insoluble sources [243] or nutrient poor soils [244], and tolerance to heavy metal (nickel and cadmium) stresses [245,246].

3.4. Epichloë Effects on Invertebrates

Epichloë bioactivity against insect pests were reported in the early 1980s [247]. In New Zealand, the major negative impact on ryegrass persistence is caused by a range of insect pests, some native and some introduced [248], and is often compounded by abiotic factors such as drought [249]. Ergot alkaloids, indole diterpenes (e.g., lolitrem B and epoxyjanthitrems), peramine, and the saturated aminopyrrolizidines (lolines) are alkaloids expressed by Epichloë strains that can protect the host plant from a range of insects [250,251] (Table 4) and can also result in anti-herbivore effects [30].
Peramine does not appear to control any pasture insect pests other than Argentine stem weevil [84,247,326].
A number of important pasture pests have to date not been shown to be controlled by specific strains or different species of Epichloë endophytes. These include blackheaded pasture cockchafer (Aphodius tasmaniae) in Australia [262,367], tobacco hornworm (Manduca sexta), tobacco budworm (Heliothis virescens), redlegged grasshoppers (Melanoplus femurrubrum) [368], the aphids Sitobion avenae [326], Metopholophium dirhodum and Sitobion fragariae [325], and the nematodes Helicotylenchus pseudorobustus [356], Paratylenchus, and Tylenchus [369].

3.5. Epichloë Effects on Other Microorganisms

Epichloë endophytes have frequently shown a negative impact on pathogens of grasses in planta [370,371] (Table 5). In vitro testing using dual culture assays have also often shown some antifungal effect from Epichloë [372,373,374,375,376], but these do not necessarily predict in planta effects [373]. Mechanisms for preventing disease in host plants by Epichloë may include (a) expression of volatile organic compounds to prevent insect attack which may transfer pathogens, (b) occupation of similar ecological niches in the plant, (c) enhancing the host plants growth, particularly at establishment, and/or (d) production of antifungal molecules, proteins, antioxidants, alkaloids, phytohormones, and phenolic compounds [371]. Interestingly, it has been shown that the Epichloë symbiosis strongly influences the endophytic fungal community (including pathogens) in the leaves of its host plant (tall fescue) so that the relative abundance of other fungal taxa can be quite different from Epichloë free plants [377]. However, the same study showed that there were only negligible effects of Epichloë on bacterial community structures in plant leaves. Rhizosphere communities are also affected by Epichloë, the presence of which increases species richness, particularly of Firmicutes in colonised tall fescue plants [378]. The diversity of root-associated bacterial and fungal communities was, however, found to decrease with Epichloë gansuensis within its host grass Achnatherum inebrians, but this interaction enhanced the diversity and richness of the rhizosphere soil bacterial community [379,380]. Within the phyllosphere, particular epiphytic bacterial microflora was observed to be selected for in endophyte-infected tall fescue associations [381]. Interestingly, it has been found that an increased population of plant-growth promoting bacteria in infected seed compared to endophyte-free varieties, may provide a non-direct mechanism by which Epichloë could possibly improve reproductive plant processes [382]. These studies demonstrate that microbial keystone species such as Epichloë can impact the host’s microbial community structures, which in turn can affect plant performance and ecosystem functions associated with the plant.

3.6. Epichloë Effects on Plant Growth

Epichloë presence can improve host establishment, growth, survival, tillering, and seed production [156,404]. Using clonal ryegrass genotypes, it has been shown that there can be significant improvements in yield of leaf, pseudostem, and root due to Epichloë endophyte infection compared with uninfected plants [405]. However, often the endophyte will interact with genotype to influence relative growth rate and productivity [406]. From a physiological viewpoint Epichloë endophyte in perennial ryegrass contributed to maintaining the photosynthesis mechanism under zinc stress, although it did not significantly modify net photosynthesis [407].

4. Delivering Epichloë into Managed Pastoral Systems

The impact of Epichloë endophytes has been of greater interest in New World pastures than in Europe driven by enhancing productivity and persistence of the host species [408]. The demonstration and realisation that Epichloë endophytes were important for grass persistence in these temperate pastures led to the creation of novel host plant–endophyte strain combinations that greatly enhance the persistence of the grass but with nil or much reduced (acceptable and manageable) adverse impacts on animals [6,409]. The process to deliver Epichloë endophytes to commerce requires a range of science capability and testing to ensure reliable bioactivity against biotic stresses that enhances plant survival while ensuring good animal health and welfare outcomes [6,410,411,412]. Through this process a number of novel Epichloë strains have been delivered and are now commercially used in New Zealand, USA and South America.

4.1. Case Study—AR1TM for Ryegrass

The animal health and welfare issues created by the expression of ergovaline and lolitrem B led to the search for Epichloë strains that did not express these alkaloids, but were still able to provide the grass plant with resistance to major pasture pests. In New Zealand, during the 1990s, this was Argentine stem weevil and the endophyte released commercially to provide resistance while not causing ryegrass staggers was AR1 [280,413]. AR1 associations produce peramine but do not produce lolitrem B or ergovaline [414,415]. However, while effectively controlling Argentine stem weevil and pasture mealy bug, AR1 has only a moderate effect on African black beetle [282] (Table 6). AR1 can also be more susceptible to root aphid when compared to the same ryegrass germplasm without endophyte [259,416].
Released in 2001, AR1 quickly gained prominence in the market and become an endophyte of choice [12,417,418]. Over a 3-year period cows grazing AR1-infected ryegrass pastures produced 318 kg milk solids per cow per season while cows grazing standard-endophyte-infected pastures produced only 292 kg milk solids per cow, a significant 9% difference [419]. Other dairy grazing trials have demonstrated milk production increases of 6.7% [420] and up to 14% [421]. Mean summer–autumn growth rates were 170, 150, and 102 g/head/d for weaned lambs grazing cultivars with standard endophyte, nil endophyte, and AR1 endophyte, respectively [LSD0.05 = 48 g/head/d] [417]. These increases in production, without any endophyte associated animal health problems, have led to an unprecedented uptake of this technology by New Zealand pastoral farmers [12,422].
Table 6. Effects of AR1 endophyte strain in perennial ryegrass on pasture pests. (Taken from [282]).
Table 6. Effects of AR1 endophyte strain in perennial ryegrass on pasture pests. (Taken from [282]).
Insect PestEndophyte Strain
NilStandardAR1
Argentine Stem Weevil
% tillers with larval damage34 b4 a1 a
African Black Beetle
% tillers damaged by adults—6-month-old plants52 c8 a22 b
% plants damaged by larvae58 b36 a,b28 a
Pasture Mealy Bug
Number per core23 b0.6 a0 a
Root Aphid
Number per core1.4 a3.5 a2.4 a
a,b,c Within a row, means without a common superscript letter differ significantly (p < 0.05).

4.2. Case Study—AR37TM for Ryegrass

Despite the success of AR1 in controlling the impact of Argentine stem weevil on ryegrass persistence, a loss of plants began to occur through the early 2000s and this was due to the presence of other pests that were not controlled by AR1 [248,300]. Notably, these included African black beetle [423], another introduced pest and the two native pests, grass grub and porina [424]. Also impacting persistence were root aphid [259] and pasture mealy bug [271]. The AR37 endophyte was identified in the early 1990s and was shown to not produce any known problematic alkaloid compounds, but did produce a unique set of epoxyjanthitrem compounds [66,425]. These compounds have been linked to staggers in sheep, but they tend to be less frequent and less severe than those caused by lolitrem B [191,417,426]. Ryegrass staggers has not been recorded in dairy cows grazing pastures infected with AR37 endophyte [427].
In New Zealand, AR37 was found to confer a wide range of tolerance to insect pests, including Argentine stem weevil, African black beetle, root aphid, pasture mealy bug, and porina [248,259,260,261,263,271,300,352,353,416,428,429,430,431] (Table 7). The high level of resistance to the ubiquitous root aphid may be one of the factors that give plants infected with AR37 a yield advantage in nation-wide field trials [432]. AR37 also provided increased ryegrass tiller numbers, root mass and depth, persistence, and higher yields at critical times of the year [432]. With these significant benefits provided by AR37, farmers have learnt to manage the potential downside associated with epoxyjanthitrem compounds such that staggers events are rarely reported.
In New Zealand, AR37 provides significant benefits to sheep farmers through providing improved growth during the summer and autumn. During this period, lambs on pure ryegrass pastures, over a 6-year period, averaged 44 g/head/day on standard endophyte, 129 g/day on nil-endophyte and 131 g/day on AR37 infected pastures, representing increases in lamb growth of 198% over standard endophyte [417]. Total milk solids production over three consecutive lactations were not affected by use of AR37 compared with standard endophyte, indicating that AR37 is a choice of novel endophyte for pasture renewal when local insect pest populations are high [433].
In Australia, under dairy management and supplementary feeding regimes common to south-eastern Australia, the novel endophytes AR1 and AR37 had no effect on the milk production compared with the standard endophyte and did not cause ryegrass staggers [262]. They also noted that AR37 gave protection against pasture tunnel moth (Philobota spp.), root aphid, and an unidentified species of mealybug.

4.3. Case Study—Endo5TM and NEA Endophytes for Ryegrass

Another approach to providing efficacious endophyte for improving ryegrass persistence was to identify Epichloë strains that produced little or no lolitrem B and only low levels of ergovaline. This resulted in the identification and subsequent commercialisation of the branded endophytes Endo5 (originally marketed as Endosafe) [430], NEA (which is strain NEA2) [434], NEA2 (mixture of strains NEA2 and NEA6) [435], and NEA4 (mixture of strains NEA2 and NEA3) (dxgh891opzso3.cloudfront.net › files › NEA4 booklet; [435]). The strategy behind these types of endophytes was to identify strains where ergovaline concentrations are high enough to protect against insect attack, but low enough to have minimal impact on grazing animals [436]. While NEA2 endophyte does protect ryegrass against African black beetle and pasture mealybug [248] and Argentine stem weevil [316] it does not protect ryegrass against porina or the mealybug Phenococcus sp. [264]. For protection against Argentine stem weevil, NEA2, which produces peramine has shown some resistance in the diploid cultivar Trojan [437], but little protection when in tetraploid cultivar Bealey [248,431,438]. Endo5 provides good protection against Argentine stem weevil, African black beetle, pasture mealybug [248], and root aphid [264], but not against grass grub [248]. This study also showed that for the NEA type endophytes, even though they express some level of ergovaline, they did not protect the host plant against root aphid.
Some of the NEA branded endophytes, such as NEA2 may also express low levels of ergovaline [434]. This however allows for the potential risk of ergovaline rising to toxic levels in some seasons or in adverse environments [439], which is predicted to occur more frequently due to climate change. It has been concluded that when ambient temperatures are suitable, NEA2-branded endophytes, just like standard endophyte, have the potential to express concentrations of ergovaline sufficient to induce heat stress in grazing sheep [434]. Others have also noted that ryegrasses infected with NEA2/3 (branded NEA4) and NEA2/6 (branded NEA2) endophytes had similar or higher concentrations of ergovaline than standard endophyte-infected ryegrass [440]. The impacts of ergovaline in New Zealand pastures has been well reviewed and found that ergovaline in standard endophyte-infected pastures can reach concentrations sufficient to cause toxicosis when ambient temperatures are suitable [439].

4.4. Case Study—Happe and U2 Both Fescue Epichloë Strains for Use in Ryegrass

Unlike Epichloë endophytes from ryegrass, those found in fescue can express lolines which are animal safe and yet have insecticidal properties against a range of insect pests (Table 4). Moving Epichloë endophytes from fescues into ryegrass through isolation and inoculation has been attempted but has proven challenging. Only two have moved to commercialisation, Happe, a unique endophyte of the species E. siegelii [36], and U2 (E. uncinatum) [302,303,441], both from meadow fescue.
Perennial ryegrass inoculated with Happe have shown reasonably high expression of loline alkaloids [172], which may be sufficient to give protection against major insect pests including the grass grub.
U2 has been inoculated into festulolium hybrids [442] in an attempt to improve seed transmission rates. The principle loline type expressed by U2 in festulolium hybrids was NFL (68% of total lolines), followed by NAL (23%), and NANL (8%) [443]. The endophyte strain U2 has shown to provide good resistance against a range of insect pests, including grass grub [92,321], African black beetle [302], Argentine stem weevil [431,444], and crickets [302].

4.5. Case Study—AR542 and AR548 (MaxQTM, MaxQIITM, and MaxPTM) for Tall Fescue

Fescue toxicosis has been associated with the presence of high ergovaline expressing Epichloë strains in tall fescue [161,445,446]. Replacement with endophyte strains that do not produce ergovaline has been successfully achieved and led to the release of strain AR542 (MaxQ) in the USA in 2000 [447,448,449,450,451]. This was later replaced with AR584 (MaxQII), a strain that provided all the benefits of AR542 but had improved seed borne transmission and storage characteristics [452]. AR542 expresses peramine and the loline compound NANL, while AR584 expresses peramine and the three loline compounds NFL, NAL, and NANL [284].
The MaxQ brand of endophytes has provided agronomically superior tall fescue cultivars that do not cause any fescue toxicosis symptoms [439] and has been described as a “win-win” outcome [411]. In New Zealand, MaxPTM endophyte reduces damage by African black beetle, Argentine stem weevil, pasture mealy bug, grass grub, and root aphid in a range of tall fescue cultivars [209,267,284,300,453]. Other insect pests that these ergot alkaloid free endophytes control include fall armyworm [454], corn flea beetle (Chaetocnema pulicaria) [455], and bird cherry oat aphids [341,456]. Sheep show no difference in preference to grazing MaxPTM endophyte containing tall fescue compared with nil-endophyte tall fescue [457]. Lambs grazing MaxQIITM containing tall fescue gained an average of >139 g d−1, more than twice the 68 g d−1 gained by animals grazing endophyte-infected Kentucky-31 [458].
Brood-balls from the dung beetle Onthophagu taurus preferred dung from cows grazing tall fescue Texoma MaxQ II while dung from cows grazing tall fescue Kentucky31 and BarOptima PLUSE34 were avoided [459]. Both O. taurus and the other beetle species Digitonthophagus gazella preferred dung from Texoma MaxQII compared with endophyte-infected Kentucky31 pasture.

4.6. Case Study—E34 for Tall Fescue

E34 (also known as BE9301A) produces ergovaline but at lower levels (<10% to 50% depending on host germplasm and environment) than standard endophyte Kentucky 31 tall fescue, resulting in a significantly higher average daily gains of steers of 1.93 lb compared with 1.29 lb, respectively [460]. In field trials over two years in two USA states the value of novel endophyte varieties that produce no ergot alkaloids was confirmed, and it was demonstrated that while varieties such as BarOptima Plus E34 express consistently lesser levels of ergot alkaloids than Kentucky 31 [461] (Table 8), they can elevate in some circumstances to levels that are greater than that considered safe for livestock based on previous studies [152,462].
Comparison of BarOptima and MaxQ (AR542) tall fescue endophytes, however, does show that animal performance in terms of average daily weight gain of cattle of both was similar to endophyte free tall fescue and considerably better than on the endophyte-infected Kentucky 31 pasture (Table 9). Grazing days on endophyte free pasture was low due to poor pasture resilience without the endophyte. Interestingly, blood serum prolactin levels were slightly lower for BarOptima than endophyte free and MaxQ (Table 9).

4.7. Case Study—Protek (E647) for Tall Fescue

Protek is an endophyte that does not produce ergovaline or any other ergopeptide alkaloids and in combination with tall fescue increased yields of young seedlings by 20 to 100% and increased resistance to African black beetle, which reduces severely damaged tillers of seedlings by 20% to 45% depending on host germplasm [464]. Average daily weight gain of ewes grazing over three years showed that ewes on Kentucky 31 achieved only 32 mg/head/day while those on tall fescue cultivar Martin E647 achieved 102 mg/head/day which compared favourably with a nil-endophyte Martin which achieved 103 mg/head/day [464].

4.8. Case Study—ArkShield in Tall Fescue

Also known as Strain 4 or ArkPlus, ArkShield is a strain that does not produce ergot alkaloids but does produce the lolines compounds NFL and NAL at about 50% and 100% of the levels expressed in endophyte-infected Kentucky 31 [465] (Table 10). Compared with Kentucky 31, ArkShield improved animal live weight gains and increased blood serum prolactin levels (Table 10).

4.9. Delivery of Commercial Novel Epichloë Endophytes

Effective delivery of these novel endophyte infected cultivars requires care with management of seed crops ensuring appropriate fungicides are used and seed moistures levels are 10% to 12% at seed harvest [209]. When processed the seed must be packaged appropriately and stored at low temperature and humidity until ready to be sown. Quality control systems and monitoring of endophyte viability is required through the retail and distribution chain [130,466,467]. This has been agreed among suppliers of Epichloë endophyte products.
Endophyte viability in seed should be above 70% at the point of sale to ensure farmers are purchasing a quality product [468,469]. Ensuring that the supply chain from science through seed companies and retailers to the end-user farmer are well resourced and consistent is crucial in the uptake and use of endophyte technologies in pastoral agriculture [422,470,471]. This requires using well designed production and quality assurance guidelines to deliver a high-quality endophytic seed technology, giving the farmer confidence that it will provide the promised benefits [466].

5. Future Opportunities

A significant challenge for delivering future Epichloë stains of commercial value for tall fescue and ryegrass is the scarcity of new and novel variation available in natural strains. Considerations to overcome this might include:
  • Genetic modification of Epichloë using traditional gene insertion or deletion [472,473] and the more recent CRISPR (clustered regularly interspaced short palindromic repeats)-Cas9 (CRISPR-related nuclease 9) system [474] to either:
    manipulate existing alkaloid pathways to increase the expression of mammalian safe intermediate pathway compounds, whilst removing toxic end products;
    insert secondary metabolite genes to make new compounds in planta; and/or
    repair non-functional genes (pseudogenes) in secondary metabolite pathways to restore lost bioactivity
  • Using DNA marker information to improve the efficiency of selection for endophyte compatibility in host plants when moving strains across taxa [475];
  • Identify and determine the function of bacteria associated with Epichloë in planta [476]; and
  • Develop an understanding of molecular processes that underpin compatibility between the host and fungal endophyte so that movement of Epichloë strains across widely separated taxa can be achieved successfully, ensuring normal phenotypes and good transmission through seed [475,477,478]. This may require genetic manipulation of genes in both partners to be successful, but on the other hand, the genetic information may simply be used to screen for compatible endophyte and host germplasm that are more likely able to form stable and beneficial symbioses.
Epichloë endophytes are known to produce a large number of secondary metabolites, many in planta [67,479], but some at low amounts in culture [83]. Exploitation of these has not as yet been realised but may result in bioactives that have anthelmintic effects, impacts on methanogenic microbes in ruminants, and pesticidal [480] and antifungal effects [374,401,481,482].

6. Concluding Comment

Epichloë endophytes have been found in a wide range of wild grasses across most temperate regions of the world. Strains of Epichloë are characterised by the range of alkaloids they are capable of producing in planta. These can provide an adaptive advantage to the host grass through reducing herbivory of ruminants, providing resistance to some pests and pathogens, and improving tolerances to some abiotic stresses. In some temperate regions, namely New Zealand, Australia, and USA, it has been demonstrated that ryegrass and tall fescue pastures require plants to be infected with Epichloë for them to yield well and persist. However, for Epichloë strains to be effectively commercialised, their characterisation is required to ensure that the expression of specific alkaloids while providing an advantage to the plant do not also result in animal health and welfare concerns. This has been achieved, with several different Epichloë strains being successfully commercialised and widely used by pastoral farmers.

Author Contributions

J.R.C. conceptualized the review scope and undertook much of the detailed reviewing. L.J.J. added considerable value in the sections pertinent to her knowledge base and understanding. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

John Caradus is employed by Grasslanz Technology Ltd. who is a part owner of the intellectual property associated with Epichloë strains AR1TM, AR37TM, Endo5TM, MaxQTM, MaxQIITM, MaxPTM, Happe and Protek.

References

  1. Brockwell, J.; Bottomley, P.J.; Thies, J.E. Manipulation of rhizobia microflora for improving legume productivity and soil fertility: A critical assessment. Plant Soil 1995, 174, 143–180. [Google Scholar] [CrossRef]
  2. Rillig, M.C.; Sosa-Hernández, M.A.; Roy, J.; Aguilar-Trigueros, C.A.; Vályi, K.; Lehmann, A. Towards an Integrated Mycorrhizal Technology: Harnessing Mycorrhiza for Sustainable Intensification in Agriculture. Front. Plant Sci. 2016, 7, 1625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Johnson, L.J.; de Bonth, A.C.M.; Briggs, L.R.; Caradus, J.R.; Finch, S.C.; Fleetwood, D.J.; Fletcher, L.R.; Hume, D.E.; Johnson, R.D.; Popay, A.J.; et al. The exploitation of epichloae endophytes for agricultural benefit. Fungal Divers. 2013, 60, 171–188. [Google Scholar] [CrossRef]
  4. Hume, D.E.; Ryan, G.D.; Gibert, A.; Helander, M.; Mirlohi, A.; Sabzalian, M.R. Epichloë fungal endophytes for grassland ecosystems. In Sustainable Agriculture Reviews; Lichtfouse, E., Ed.; Springer International Publishing: Cham, Switzerland, 2016; pp. 233–305. [Google Scholar]
  5. Hume, D.E.; Stewart, A.V.; Simpson, W.R.; Johnson, R.D. Epichloë fungal endophytes play a fundamental role in New Zealand grasslands. J. R. Soc. N. Z. 2020, 50, 279–298. [Google Scholar] [CrossRef]
  6. Johnson, L.; Caradus, J. The Science Required to Deliver Epichloë Endophytes to Commerce. In Endophytes for a Growing World; Hodkinson, T., Doohan, F., Saunders, M., Murphy, B., Eds.; Cambridge University Press: Cambridge, UK, 2019; pp. 343–370. [Google Scholar] [CrossRef]
  7. Le Cocq, K.; Gurr, S.J.; Hirsch, P.R.; Mauchline, T.H. Exploitation of endophytes for sustainable agricultural intensification. Mol. Plant Pathol. 2017, 18, 469–473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Card, S.; Johnson, L.; Teasdale, S.; Caradus, J. Deciphering endophyte behaviour—The link between endophyte biology and efficacious biological control agents. FEMS Microbiol. Ecol. 2016. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. White, J.F., Jr. Endophyte-host associations in grasses. XIX. A systematic study of some sympatric species of Epichloë in England. Mycologia 1993, 85, 444–455. [Google Scholar]
  10. Saikkonen, K.; Young, C.A.; Helander, M.; Schardl, C.L. Endophytic Epichloë species and their grass hosts: From evolution to applications. Plant Mol. Biol. 2016, 90, 65–675. [Google Scholar] [CrossRef] [Green Version]
  11. Schardl, C.L.; Leuchtmann, A. Three new species of Epichloë symbiotic with North American grasses. Mycologia 1999, 91, 95–107. [Google Scholar] [CrossRef]
  12. Caradus, J.R.; Lovatt, S.; Belgrave, B. Adoption of forage technologies. Proc. N. Z. Grassl. Assoc. 2013, 75, 39–44. [Google Scholar] [CrossRef]
  13. Lane, G.A. Chemistry of Endophytes: Patterns and Diversity. In Proceedings of the Ryegrass Endophyte: An Essential New Zealand Symbiosis; Napier, New Zealand, 8 October 1999, Woodfield, D.R., Matthew, C., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 7, pp. 85–94.
  14. Tadych, M.; Bergen, M.S.; White Jr, J.F. Epichloë spp. associated with grasses: New insights on life cycles, dissemination and evolution. Mycologia 2014, 106, 181–201. [Google Scholar] [CrossRef] [PubMed]
  15. Leuchtmann, A.; Bacon, C.W.; Schardl, C.L.; White, J.F.; Tadych, M. Nomenclatural realignment of Neotyphodium species with genus Epichloë. Mycologia 2014, 106, 202–215. [Google Scholar] [CrossRef] [PubMed]
  16. Morgan-Jones, G.; Gams, W. Notes on hyphomycetes. XLI. An endophyte of Festuca arundinacea and the anamorph of Epichloë typhina, new taxa in one of two new sections of Acremonium. Mycotaxon 1982, 15, 311–318. [Google Scholar]
  17. McNeill, J. (Ed.) International Code of Nomenclature for Algae, Fungi, and Plants (Melbourne Code); Eighteenth International Botanical Congress: Melbourne, Australia, 2012. [Google Scholar]
  18. Schardl, C.L.; Phillips, T.D. Protective grass endophytes. Where are they from and where are they going? Plant Dis. 1997, 81, 430–438. [Google Scholar] [CrossRef] [Green Version]
  19. Faeth, S.H. Are endophytic fungi defensive plant mutualists? Oikos 2002, 98, 25–36. [Google Scholar] [CrossRef] [Green Version]
  20. Leuchtmann, A.; Young, C.A.; Stewart, A.V.; Simpson, W.R.; Hume, D.E.; Scott, B. Epichloë novaezelandiae, a new endophyte from the endemic New Zealand grass Poa matthewsii. N. Z. J. Bot. 2019, 57, 271–288. [Google Scholar] [CrossRef]
  21. McGranahan, D.A.; Burgdorf, R.; Kirkman, K.P. Epichloae infection in a native South African grass, Festuca costata Nees. Plant Biol. 2015, 17, 914–921. [Google Scholar] [CrossRef]
  22. Song, H.; Nan, Z.; Song, Q.; Xia, C.; Li, X.; Yao, X.; Xu, W.; Kuang, Y.; Tian, P.; Zhang, Q. Advances in research on Epichloë endophytes in Chinese native grasses. Front. Microbiol. 2016, 7, 1399. [Google Scholar] [CrossRef] [Green Version]
  23. Zeven, A.C.; de Wet, J.M.J. Dictionary of Cultivated Plants and Their Regions of Diversity; Pudoc.: Wageningen, The Netherlands, 1982. [Google Scholar]
  24. Iannone, L.J.; Irisarri, J.G.N.; Mc Cargo, P.D.; Perez, L.I.; Gundel, P.E. Occurrence of Epichloë fungal endophytes in the sheep-preferred grass Hordeum comosum from Patagonia. J. Arid Environ. 2015, 115, 19–26. [Google Scholar] [CrossRef]
  25. Card, S.D.; Faville, M.J.; Simpson, W.R.; Johnson, R.D.; Voisey, C.R.; de Bonth, A.C.M.; Hume, D.E. Mutualistic fungal endophytes in the Triticeae—Survey and description. FEMS Microbiol. Ecol. 2014, 88, 94–106. [Google Scholar] [CrossRef] [Green Version]
  26. Simpson, W.R.; Faville, M.J.; Moraga, R.A.; Williams, W.M.; McManus, M.T.; Johnson, R.D. Epichloë fungal endophytes and the formation of synthetic symbioses in Hordeeae (= Triticeae) grasses. J. Syst. Evol. 2014, 52, 794–806. [Google Scholar] [CrossRef]
  27. Simpson, W.R.; Popay, A.J.; Mace, W.J.; Hume, D.E.; Johnson, R.D. Creating synthetic symbioses between Epichloë and rye (Secale cereale) to improve crop performance. In Proceedings of the International Symposium on Fungal Endophyte of Grasses, Salamanca, Spain, 18–21 June 2018; p. 96. [Google Scholar]
  28. Simpson, W.R.; Tsujimoto, H.; Johnson, R.D. Endophyte Screening. New Zealand Patent 740,055, 21 February 2018. [Google Scholar]
  29. Moon, C.D.; Craven, K.D.; Leuchtmann, A.; Clement, S.L.; Schardl, C.L. Prevalence of interspecific hybrids among asexual fungal endophytes of grasses. Mol. Ecol. 2004, 13, 1455–1467. [Google Scholar] [CrossRef] [PubMed]
  30. Schardl, C.L.; Young, C.A.; Faulkner, J.R.; Florea, S.; Pan, S.J. Chemotypic diversity of epichloae, fungal symbionts of grasses. Fungal Ecol. 2012, 5, 331–344. [Google Scholar] [CrossRef]
  31. Tadych, M.; Ambrose, K.V.; Bergen, M.S.; Belanger, F.C.; White, J.F., Jr. Taxonomic placement of Epichloë poae sp. nov. and horizontal dissemination to seedlings via conidia. Fungal Divers. 2012, 54, 117–131. [Google Scholar] [CrossRef]
  32. Cagnano, G.; Roulund, N.; Jensen, C.S.; Forte, F.P.; Asp, T.; Leuchtmann, A. Large Scale Screening of Epichloë Endophytes Infecting Schedonorus pratensis and Other Forage Grasses Reveals a Relation Between Microsatellite-Based Haplotypes and Loline Alkaloid Levels. Front. Plant Sci. 2019, 10, 765. [Google Scholar] [CrossRef]
  33. Moon, C.D.; Scott, B.; Schardl, C.L.; Christensen, M.J. Evolutionary origins of Epichloë endophytes from annual ryegrasses. Mycologia 2000, 92, 1103–1118. [Google Scholar]
  34. Campbell, M.A.; Tapper, B.A.; Simpson, W.R.; Johnson, R.D.; Mace, W.J.; Ram, A.; Lukito, Y.; Dupont, P.-Y.; Johnson, L.J.; Scott, D.B.; et al. Epichloë hybrida, sp. nov., an emerging model system for investigating fungal allopolyploidy. Mycologia 2017, 109, 715–729. [Google Scholar] [CrossRef] [Green Version]
  35. Lembicz, M.; Górzyńska, K.; Leuchtmann, A. Choke disease caused by Epichloë bromicola in the grass Agropyron repens in Poland. Plant Dis. 2010, 94, 1372. [Google Scholar] [CrossRef]
  36. Craven, K.D.; Blankenship, J.D.; Leuchtmann, A.; Highnight, K.; Schardl, C.L. Hybrid fungal endophytes symbiotic with the grass Lolium pratense. Sydowia 2001, 53, 44–73. [Google Scholar]
  37. Oberhofer, M.; Leuchtmann, A. Genetic diversity in epichloid endophytes of Hordelymus europaeus suggests repeated host jumps and interspecific hybridizations. Mol. Ecol. 2012, 21, 2713–2726. [Google Scholar] [CrossRef]
  38. Clay, K.; Brown, V.K. Infection of Holcus lanatus and H. mollis by Epichloë in experimental grasslands. Oikos 1997, 79, 363–370. [Google Scholar] [CrossRef]
  39. Chen, L.; Li, X.Z.; Li, C.J.; Swoboda, G.A.; Young, C.A.; Sugawara, K.; Leuchtmann, A.; Schardl, C.L. Two distinct Epichloë species symbiotic with Achnatherum inebrians, drunken horse grass. Mycologia 2015, 107, 863–873. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Kang, Y.; Ji, Y.L.; Zhu, K.; Wang, H.; Miao, H.; Wang, Z.W. A new Epichloë species with interspecific hybrid origins from Poa pratensis ssp. pratensis in Liyang, China. Mycologia 2011, 103, 1341–1350. [Google Scholar] [CrossRef] [Green Version]
  41. Li, W.; Ji, Y.-L.; Yu, H.-S.; Wang, Z.-W. A new species of Epichloë symbiotic with Chinese grasses. Mycologia 2006, 98, 560–570. [Google Scholar] [CrossRef] [PubMed]
  42. Drake, I.; White, J.F., Jr.; Belanger, F.C. Identification of the fungal endophyte of Ammophila breviligulata (American beachgrass) as Epichloë amarillans. PeerJ 2018, 6, e4300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Charlton, N.D.; Craven, K.D.; Afkhami, M.E.; Hall, B.A.; Ghimire, S.R.; Young, C.A. Interspecific hybridization and bioactive alkaloid variation increases diversity in endophytic Epichloë species of Bromus laevipes. FEMS Microbiol. Ecol. 2014, 90, 276–289. [Google Scholar] [CrossRef] [Green Version]
  44. Ghimire, S.R.; Rudgers, J.A.; Charlton, N.D.; Young, C.; Craven, K.D. Prevalence of an intraspecific Neotyphodium hybrid in natural populations of stout wood reed (Cinna arundinacea L.) from eastern North America. Mycologia 2011, 103, 75–84. [Google Scholar] [CrossRef] [Green Version]
  45. Burr, K.; Mittai, S.; Hopkins, A.; Young, C. Characterisation of fungal endophytes present in Elymus canadensis (Canada rye). In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 473–476.
  46. Shymanovich, T.; Charlton, N.D.; Musso, A.M.; Scheerer, J.; Cech, N.B.; Faeth, S.H.; Young, C.A. Interspecific and intraspecific hybrid Epichloë species symbiotic with the North American native grass Poa alsodes. Mycologia 2017, 109, 459–474. [Google Scholar] [CrossRef] [Green Version]
  47. McCargo, P.D.; Iannone, L.J.; Vignale, M.V.; Schardl, C.L.; Rossi, M.S. Species diversity of Epichloë symbiotic with two grasses from southern Argentinean Patagonia. Mycologia 2014, 106, 339–352. [Google Scholar] [CrossRef]
  48. Gentile, A.; Rossi, M.S.; Cabral, D.; Craven, K.D.; Schardl, C.L. Origin, divergence and phylogeny of Epichloë endophytes of native Argentine grasses. Mol. Phylogenet Evol. 2005, 35, 196–208. [Google Scholar] [CrossRef]
  49. Cabral, D.; Iannone, L.J.; Stewart, A.; Novas, M.V. The distribution and incidence of Neotyphodium endophytes in native grasses from Argentina and its association with environmental factors. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses, Christchurch, New Zealand, 25–28 March 2007; Popay, A.J., Thom, E.R., Eds.; Volume 13, pp. 79–82. [Google Scholar]
  50. Moon, C.D.; Miles, C.O.; Järlfors, U.; Schardl, C.L. The evolutionary origins of three new Neotyphodium endophyte species from grasses indigenous to the Southern Hemisphere. Mycologia 2002, 94, 694–711. [Google Scholar] [CrossRef] [PubMed]
  51. Iannone, L.J.; Novas, M.V.; Young, C.A.; De Battista, J.P.; Schardl, C.L. Endophytes of native grasses from South America: Diversity and ecology. Fungal Ecol. 2012, 5, 357–363. [Google Scholar] [CrossRef]
  52. Iannone, L.J.; McCargo, P.D.; Giussani, L.M.; Schardl, C.L. Geographic distribution patterns of vertically transmitted endophytes in two native grasses in Argentina. Symbiosis 2013, 59, 99–110. [Google Scholar] [CrossRef]
  53. Miles, C.O.; di Menna, M.; Jacobs, S.W.L.; Garthwaite, I.; Lane, G.A.; Prestidge, R.A.; Marshall, S.L.; Wilkinson, H.H.; Schardl, C.L.; Ball, O.-J.P.; et al. Endophytic fungi in indigenous Australasian grasses associated with toxicity to livestock. Appl. Environ. Microbiol. 1998, 64, 601–606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Selosse, M.A.; Schardl, C.L. Fungal endophytes of grasses: Hybrids rescued by vertical transmission? An evolutionary perspective. New Phytol. 2007, 173, 452–458. [Google Scholar] [CrossRef] [PubMed]
  55. Schardl, C.L. The epichloae, symbionts of the grass subfamily Poideae. Ann. Mo. Bot. Gard. 2010, 97, 646–665. [Google Scholar] [CrossRef]
  56. Schardl, C.L.; Craven, K.D. Interspecific hybridization in plant-associated fungi and oomycetes: A review. Mol. Ecol. 2003, 12, 2861–2873. [Google Scholar] [CrossRef]
  57. Kuldau, G.A.; Tsai, H.-F.; Schardl, C.L. Genome sizes of Epichloë species and anamorphic hybrids. Mycologia 1999, 91, 776–782. [Google Scholar] [CrossRef]
  58. Chung, K.-R.; Schardl, C.L. Sexual cycle and horizontal transmission of the grass symbiont. Epichloë typhina. Mycol. Res. 1997, 101, 295–301. [Google Scholar] [CrossRef]
  59. White, J.; Martin, T.I.; Cabral, D. Endophyte-host associations in grasses. XXII. Conidia formation by Acremonium endophytes on the phylloplanes of Agrostis hiemalis and Poa rigidifolia. Mycologia 1996, 88, 174–178. [Google Scholar] [CrossRef]
  60. Schardl, C.L.; Young, C.; Moore, N.; Krom, N.; Dupont, P.-Y.; Pan, J.; Florea, S.; Webb, J.S.; Jaromczyk, J.; Jaromczyk, J.W.; et al. Genomes of plant-associated Clavicipitaceae. Adv. Bot. Res. 2014, 70, 291–327. [Google Scholar]
  61. Saari, S.; Faeth, S.H. Hybridization of Neotyphodium endophytes enhances competitive ability of the host grass. New Phytol. 2012, 195, 231–236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Shymanovich, T.; Faeth, S.H. Environmental factors affect the distribution of two Epichloë fungal endophyte species inhabiting a common host grove bluegrass (Poa alsodes). Ecol. Evol. 2019, 9, 6624–6642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Oberhofer, M.; Gusewell, S.; Leuchtmann, A. Effects of natural hybrid and non-hybrid Epichloë endophytes on the response of Hordelymus europaeus to drought stress. New Phytol. 2014, 201, 242–253. [Google Scholar] [CrossRef] [PubMed]
  64. Schardl, C.L.; Florea, S.; Pan, J.; Nagabhyru, P.; Bec, S.; Calie, P.J. The epichloae: Alkaloid diversity and roles in symbiosis with grasses. Curr. Opin. Plant Biol. 2013, 16, 80–488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Panaccione, D.G.; Beaulieu, W.T.; Cook, D. Bioactive alkaloids in vertically transmitted fungal endophytes. Funct. Ecol. 2014, 28, 299–314. [Google Scholar] [CrossRef] [Green Version]
  66. Finch, S.C.; Prinsep, M.R.; Popay, A.J.; Wilkins, A.L.; Webb, N.G.; Bhattarai, S.; Jensen, J.G.; Hawkes, A.D.; Babu, J.V.; Tapper, B.A.; et al. Identification and structure elucidation of epoxyjanthitrems from Lolium perenne infected with the endophytic fungus Epichloë festucae var. lolii and determination of the tremorgenic and anti-insect activity of epoxyjanthitrem I. Toxins 2020, 12, 526. [Google Scholar] [CrossRef]
  67. Schardl, C.L.; Young, C.A.; Hesse, U.; Amyotte, S.G.; Andreeva, K.; Calie, P.J.; Fleetwood, D.J.; Haws, D.C.; Moore, N.; Oeser, B.; et al. Plant symbiotic fungi as chemical engineers: Multi-genome analysis of the Clavicipitaceae reveals dynamics of alkaloid loci. PLoS Genet. 2013, 9, e1003323. [Google Scholar] [CrossRef] [Green Version]
  68. Miller, T.A. Insect Bioactive Capabilities of Epichloë Festucae var lolii AR48 Infected Lolium perenne. Ph.D. Thesis, Massey University, Palmerston North, New Zealand, 2018. [Google Scholar]
  69. Johnson, R.D.; Lane, G.A.; Koulman, A.; Cao, M.; Fraser, K.; Fleetwood, D.J.; Voisey, C.; Dyer, J.M.; Pratt, J.; Christensen, M.; et al. A novel family of cyclic oligopeptides derived from ribosomal peptide synthesis of an in planta-induced gene, gigA, in Epichloë endophytes of grasses. Fungal Genet Biol. 2015, 85, 14–24. [Google Scholar] [CrossRef]
  70. Song, Q.; Yu, H.; Zhang, X.; Nan, Z.; Kun Gao, K. Dahurelmusin A, a hybrid peptide–polyketide from Elymus dahuricus Infected by the Epichloë bromicola endophyte. Org. Lett. 2017, 19, 298–300. [Google Scholar] [CrossRef]
  71. Lane, G.A.; Christensen, M.J.; Miles, C.O. Coevolution of fungal endophytes with grasses: The significance of secondary metabolites. In Microbial Endophytes; Bacon, C.W., White, J.F., Eds.; Marcel Dekker: New York, NY, USA, 2000; pp. 341–388. [Google Scholar]
  72. Easton, H.S.; Latch, G.; Tapper, B.; Ball, J. Ryegrass host genetic control of concentrations of endophyte-derived alkaloids. Crop Sci. 2002, 42, 51–57. [Google Scholar] [PubMed]
  73. Faeth, S.H.; Bush, L.P.; Sullivan, T. Peramine alkaloid variation in Neotyphodium-infected Arizona fescue: Effects of endophyte and host genotype and environment. J. Chem. Ecol. 2002, 28, 1511–1526. [Google Scholar] [CrossRef] [PubMed]
  74. Bastias, D.A.; Martinez-Ghersa, M.A.; Ballare, C.L.; Gundel, P.E. Epichloë fungal endophytes and plant defenses: Not just alkaloids. Trends Plant Sci. 2017. [Google Scholar] [CrossRef] [PubMed]
  75. Lyons, P.C.; Plattner, R.D.; Bacon, C.W. Occurrence of peptide and clavine ergot alkaloids in tall fescue. Science 1986, 232, 487–489. [Google Scholar] [CrossRef] [Green Version]
  76. Agee, C.; Hill, N. Ergovaline variability in Acremonium-infected tall fescue due to environment and plant genotype. Crop Sci. 1994, 34, 221–226. [Google Scholar] [CrossRef]
  77. Rasmussen, S.; Parsons, A.J.; Bassett, S.; Christensen, M.J.; Hume, D.E.; Johnson, L.J.; Johnson, R.D.; Simpson, W.R.; Stacke, C.; Voisey, C.R.; et al. High nitrogen supply and carbohydrate content reduce fungal endophyte and alkaloid concentration in Lolium perenne. New Phytol. 2007, 173, 787–797. [Google Scholar] [CrossRef]
  78. Brosi, G.; McCulley, R.; Bush, L.; Nelson, J.; Classen, A.; Norby, R. Effects of multiple climate change factors on the tall fescue-fungal endophyte symbiosis: Infection frequency and tissue chemistry. New Phytol. 2011, 189, 797–805. [Google Scholar] [CrossRef]
  79. Repussard, C.; Zbib, N.; Tardieu, D.; Guerre, P. Ergovaline and lolitrem B concentrations in perennial ryegrass in field culture in southern France: Distribution in the plant and impact of climatic factors. J. Agric. Food Chem. 2014, 62, 12707–12712. [Google Scholar] [CrossRef]
  80. Hennessy, L.M.; Popay, A.J.; Finch, S.C.; Clearwater, M.J.; Cave, V.M. Temperature and plant genotype alter alkaloid concentrations in ryegrass infected with an Epichloë endophyte and this affects an insect herbivore. Front. Plant Sci. 2016, 7, 1097. [Google Scholar] [CrossRef]
  81. Chujo, T.; Lukito, Y.; Eaton, C.J.; Dupont, P.Y.; Johnson, L.J.; Winter, D.; Cox, M.P.; Scott, B. Complex epigenetic regulation of alkaloid biosynthesis and host interaction by heterochromatin protein I in a fungal endophyte-plant symbiosis. Fungal Genet. Biol. 2019, 125, 71–83. [Google Scholar] [CrossRef]
  82. Tanaka, A.; Tapper, B.A.; Popay, A.; Parker, E.J.; Scott, B. A symbiosis expressed non-ribosomal peptide synthetase from a mutualistic fungal endophyte of perennial ryegrass confers protection to the symbiotum from insect herbivory. Mol. Microbiol. 2005, 57, 1036–1050. [Google Scholar] [CrossRef]
  83. Blankenship, J.D.; Spiering, M.J.; Wilkinson, H.H.; Fannin, F.F.; Bush, L.P.; Schardl, C.L. Production of loline alkaloids by the grass endophyte, Neotyphodium uncinatum, in defined media. Phytochemistry 2001, 58, 395–401. [Google Scholar] [CrossRef]
  84. Rowan, D.D. Lolitrems, peramine and paxilline: Mycotoxins of the ryegrass/endophyte interaction. Agric. Ecosyst. Environ. 1993, 44, 103–122. [Google Scholar] [CrossRef]
  85. Lukito, Y.; Chujo, T.; Hale, T.K.; Mace, W.; Johnson, L.J.; Scott, B. Regulation of subtelomeric fungal secondary metabolite genes by H3K4me3 regulators CclA and KdmB. Mol. Biol. 2019, 112, 837–853. [Google Scholar] [CrossRef]
  86. Chujo, T.; Scott, B. Histone H3K9 and H3K27 methylation regulates fungal alkaloid biosynthesis in a fungal endophyte-plant symbiosis. Mol. Microbiol. 2014, 92, 413–434. [Google Scholar] [CrossRef]
  87. Spiering, M.J.; Lane, G.A.; Christensen, M.J.; Schmid, J. Distribution of the fungal endophyte Neotyphodium lolii is not a major determinant of the distribution of fungal alkaloids in Lolium perenne plants. Phytochemistry 2005, 66, 195–202. [Google Scholar] [CrossRef]
  88. Ball, O.J.-P.; Barker, G.M.; Prestidge, R.A.; Lauren, D.R. Distribution and accumulation of the alkaloid peramine in Neotyphodium lolii-infected perennial ryegrass. J. Chem. Ecol. 1997, 23, 1419–1434. [Google Scholar] [CrossRef]
  89. Koulman, A.; Lane, G.A.; Christensen, M.J.; Fraser, K.; Tapper, B.A. Peramine and other fungal alkaloids are exuded in the guttation fluid of endophyte-infected grasses. Phytochemistry 2007, 68, 355–360. [Google Scholar] [CrossRef]
  90. Ball, O.J.-P.; Bernard, E.C.; Gwinn, K.D. Effect of selected Neotyphodium lolii isolates on root-knot nematode (Meloidogyne marylandi) numbers in perennial ryegrass. In Proceedings of the 50th New Zealand Plant Protection Conference, Canterbury, New Zealand, 18–21 August 1997; pp. 65–68. [Google Scholar]
  91. Patchett, B.J.; Chapman, R.B.; Fletcher, L.R.; Gooneratne, S.R. Root loline concentration in endophyte infected meadow fescue (Festuca pratensis) is increased by grass grub (Costelytra zealandica) attack. In Proceedings of the 61st NZ Plant Protection Conference, Paihia, New Zealand, 12–14 August 2008; pp. 210–214. [Google Scholar]
  92. Patchett, B.J.; Gooneratne, S.; Chapman, R.B.; Fletcher, L.R. Effects of loline-producing endophyte-infected meadow fescue ecotypes on New Zealand grass grub (Costelytra zealandica). N. Z. J. Agric. Res. 2011, 54, 303–313. [Google Scholar] [CrossRef] [Green Version]
  93. Moore, J.R.; Pratley, J.E.; Mace, E.J.; Weston, L.A. Variation in alkaloid production from genetically diverse Lolium accessions infected with Epichloë species. J. Agric. Food Chem. 2015, 63, 10355–10365. [Google Scholar] [CrossRef]
  94. Justus, M.; Witte, L.; Hartmann, T. Levels and tissue distribution of loline alkaloids in endophyte-infected Festuca pratensis. Phytochemistry 1996, 44, 51–57. [Google Scholar] [CrossRef]
  95. DeAngelis, D.L.; Post, W.M.; Travis, C.C. Mutualistic and Competitive Systems. In Positive Feedback in Natural Systems, Biomathematics; Springer: Berlin, Germany, 1986; p. 15. [Google Scholar]
  96. Thompson, J.N. The Coevolutionary Process; University of Chicago Press: Chicago, IL, USA, 1994; p. 383. [Google Scholar]
  97. Wilkinson, H.H.; Schardl, C.L. The evolution of mutualism in grass-endophyte associations. In Neotyphodium/Grass Interactions; Bacon, C.W., Hill, N.S., Eds.; Plenum Press: New York, NY, USA, 1997; pp. 13–25. [Google Scholar]
  98. Schardl, C.L. Epichloë festucae and related mutualistic symbionts of grasses. Fungal Genet. Biol. 2001, 33, 69–82. [Google Scholar] [CrossRef]
  99. Scott, B.; Green, K.; Berry, D. The fine balance between mutualism and antagonism in the Epichloë festucae-grass symbiotic interaction. Curr. Opin. Plant Biol. 2018, 44, 32–38. [Google Scholar] [CrossRef]
  100. Forester, N.T.; Lane, G.A.; Steringa, M.; Lamont, I.L.; Johnson, L.J. Contrasting roles of fungal siderophores in maintaining iron homeostasis in Epichloë festucae. Fungal Genet. Biol. 2018, 111, 60–72. [Google Scholar] [CrossRef]
  101. Forester, N.T.; Lane, G.A.; McKenzie, C.M.; Lamont, I.L.; Johnson, L.J. The Role of SreA-mediated iron regulation in maintaining Epichloë festucaeLolium perenne symbioses. Mol. Plant Microbe Interact. 2019, 32, 1324–1335. [Google Scholar] [CrossRef]
  102. Johnson, L.J.; Koulman, A.; Christensen, M.; Lane, G.A.; Fraser, K.; Forester, N.; Johnson, R.D.; Bryan, G.T.; Rasmussen, S. An extracellular siderophore is required to maintain the mutualistic interaction of Epichloë festucae with Lolium perenne. PLoS Pathog. 2013, 9, e1003332. [Google Scholar] [CrossRef] [Green Version]
  103. Tanaka, A.; Christensen, M.J.; Takemoto, D.; Park, P.; Scott, B. Reactive oxygen species play a role in regulating a fungus-perennial ryegrass mutualistic association. Plant Cell 2006, 18, 1052–1066. [Google Scholar] [CrossRef] [Green Version]
  104. Mitic, M.; Berry, D.; Brasell, E.; Green, K.; Young, C.A.; Saikia, S.; Rakonjac, J.; Scott, B. Disruption of calcineurin catalytic subunit (cnaA) in Epichloë festucae induces symbiotic defects and intrahyphal hyphae formation. Mol. Plant Pathol. 2018, 19, 1414–1426. [Google Scholar] [CrossRef] [Green Version]
  105. Hassing, B.; Eaton, C.J.; Winter, D.; Green, K.A.; Brandt, U.; Savoian, M.S.; Mesarich, C.H.; Fleissner, A.; Scott, B. Phosphatidic acid produced by phospholipase D is required for hyphal cell-cell fusion and fungal-plant symbiosis. bioRxiv 2020. [Google Scholar] [CrossRef]
  106. Bisson, A. The Role of the G Protein and cAMP/PKA Signalling Pathway in Establishment and Maintenance of the Mutualistic Epichloë festucae—Ryegrass Association. Ph.D. Thesis, Massey University, Palmerston North, New Zealand, 2017. [Google Scholar]
  107. Voisey, C.R.; Christensen, M.T.; Johnson, L.J.; Forester, N.T.; Gagic, M.; Bryan, G.T.; Simpson, W.R.; Fleetwood, D.J.; Card, S.D.; Koolaard, J.P.; et al. cAMP signaling regulates synchronised growth of symbiotic Epichloë fungi with the host grass Lolium perenne. Front. Plant Sci. 2016, 7, 1546. [Google Scholar] [CrossRef] [Green Version]
  108. Eaton, C.J.; Cox, M.P.; Ambrose, B.; Becker, M.; Hesse, U.; Schardl, C.L.; Scott, B. Disruption of signaling in a fungal-grass symbiosis leads to pathogenesis. Plant Physiol. 2010, 153, 1780–1794. [Google Scholar] [CrossRef] [Green Version]
  109. Becker, Y.; Eaton, C.J.; Brasell, E.; May, K.J.; Becker, M.; Hassing, B.; Cartwright, G.M.; Reinhold, L.; Barry, S.B. The fungal cell-wall integrity MAPK cascade is crucial for hyphal network formation and maintenance of restrictive growth of Epichloë festucae in symbiosis with Lolium perenne. MPMI 2015, 28, 69–85. [Google Scholar] [CrossRef] [Green Version]
  110. Ambrose, K.V.; Belanger, F.C. SOLiD-SAGE of endophyte-infected red fescue reveals numerous effects on host transcriptome and an abundance of highly expressed fungal secreted proteins. PLoS ONE 2012, 7, e53214. [Google Scholar] [CrossRef] [Green Version]
  111. Dinkins, R.D.; Nagabhyru, P.; Graham, M.A.; Boykin, D.; Schardl, C.L. Transcriptome response of Lolium arundinaceum to its fungal endophyte Epichloë coenophiala. New Phytol. 2017, 213, 324–337. [Google Scholar] [CrossRef]
  112. Dupont, P.-Y.; Eaton, C.J.; Wargent, J.J.; Fechtner, S.; Solomon, P.; Schmid, J.; Day, R.C.; Scott, B.; Cox, M.P. Fungal endophyte infection of ryegrass reprograms host metabolism and alters development. New Phytol. 2015, 208, 1227–1240. [Google Scholar] [CrossRef]
  113. Schmid, J.; Day, R.; Zhang, N.; Dupont, P.Y.; Cox, M.P.; Schardl, C.L.; Minards, N.; Truglio, M.; Moore, N.; Harris, D.R.; et al. Host tissue environment directs activities of an Epichloë endophyte, while it induces systemic hormone and defense responses in its native perennial ryegrass host. Mol. Plant Microbe Interact. 2017, 30, 138–149. [Google Scholar] [CrossRef] [Green Version]
  114. Nagabhyru, P.; Dinkins, R.D.; Schardl, C.L. Transcriptomics of Epichloë-Grass symbioses in host vegetative and reproductive stages. Mol. Plant Microbe Interact. 2019, 32, 194–207. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Sampson, K. The systemic infection of grasses by Epichloë typhina (Pers.) Tul. Trans. Br. Mycol. Soc. 1933, 18, 30–47. [Google Scholar] [CrossRef]
  116. Stone, J.K.; Bacon, C.W.; White, J. An Overview of Endophytic Microbes: Endophytism Defined. In Microbial Endophytes; Bacon, C.W., White, J.F., Eds.; Marcel Dekker: New York, NY, USA, 2000; pp. 29–33. [Google Scholar]
  117. Schardl, C.L.; Leuchtmann, A.; Spiering, M.J. Symbioses of grasses with seedborne fungal endophytes. Ann. Rev. Plant Biol. 2004, 55, 315–340. [Google Scholar] [CrossRef]
  118. Voisey, C.R. Intercalary growth in hyphae of filamentous fungi. Fungal Biol. Rev. 2010, 24, 123–131. [Google Scholar] [CrossRef]
  119. Christensen, M.J.; Bennett, R.J.; Ansari, H.A.; Koga, H.; Johnson, R.D.; Bryan, G.T.; Simpson, W.R.; Koolaard, J.P.; Nickless, E.M.; Voisey, D.R. Epichloë endophytes grow by intercalary hyphal extension in elongating grass leaves. Fungal Genet. Biol. 2008, 45, 84–93. [Google Scholar] [CrossRef] [PubMed]
  120. Schardl, C.L.; Craven, K.D.; Speakman, S.; Stromberg, A.; Lindstrom, A.; Yoshida, R. A novel test for host-symbiont codivergence indicates ancient origin of fungal endophytes in grasses. Syst. Biol. 2008, 5, 483–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Schirrmann, M.K.; Zoller, S.; Croll, D.; Stukenbrock, E.H.; Leuchtmann, A.; Fior, S. Genomewide signatures of selection in Epichloë reveal candidate genes for host specialization. Mol. Ecol. 2018, 27, 3070–3086. [Google Scholar] [CrossRef] [PubMed]
  122. Schirrmann, M.K.; Leuchtmann, A. The role of host-specificity in the reproductive isolation of Epichloë endophytes revealed by reciprocal infections. Fungal Ecol. 2015, 15, 29–38. [Google Scholar] [CrossRef]
  123. Karimi, S.; Mirlohi, A.; Sabzalain, M.R.; Tabatabaei, B.E.S.; Sharifnabi, B. Molecular evidence for Neotyphodium fungal endophyte variation and specificity within host grass species. Mycologia 2012, 104, 1281–1290. [Google Scholar] [CrossRef]
  124. Philipson, M.N. A symptomless endophyte of ryegrass (Lolium perenne) that spores on its host—a light microscope study. N. Z. J. Bot. 1989, 27, 513–519. [Google Scholar] [CrossRef] [Green Version]
  125. Clement, S.L.; Elberson, L.R.; Kynaston, M. High Neotyphodium Infection Frequencies in Tillers and Seed of Infected Wild Tall Fescue Plants. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; Grassland Association Grassland Research and Practice Series; Volume 13, pp. 49–52.
  126. Lewis, G.C.; Ravel, C.; Naffaa, W.; Astier, C.; Charmet, G. Occurrence of Acremonium—Endophytes in wild populations of Lolium sp. in European countries and a relationship between level of infection and climate in France. Ann. Appl. Biol. 1997, 130, 27–38. [Google Scholar] [CrossRef]
  127. Cheplick, G.P.; Clay, K.; Marks, S. Interactions between infection by endophytic fungi and nutrient limitation in the grasses Lolium perenne and Festuca arundinacea. New Phytol. 1989, 111, 89–97. [Google Scholar] [CrossRef]
  128. Saikkonen, K.; Faeth, S.H.; Helander, M.L.; Sullivan, T.J. Fungal endophytes: A continuum of interactions with host plant. Ann. Rev. Ecol. Syst. 1998, 29, 319–343. [Google Scholar] [CrossRef]
  129. Ravel, C.; Michalakis, Y.; Charmet, G. The effect of imperfect transmission on the frequency of mutualistic seed-borne endophytes in natural populations of grasses. Oikos 1997, 80, 18–24. [Google Scholar] [CrossRef]
  130. Rolston, M.P.; Hare, M.D.; Moore, K.K.; Christensen, M.J. Viability of Lolium endophyte fungus in seed stored at different moisture contents and temperatures. N. Z. J. Exp. Agric. 1986, 14, 297–300. [Google Scholar]
  131. Welty, R.E.; Azevedo, M.D.; Cooper, T.M. Influence of moisture content, temperature, and length of storage on seed germination and survival of endophytic fungi in seeds of tall fescue and perennial ryegrass. Phytopathology 1987, 77, 893–900. [Google Scholar] [CrossRef]
  132. Gundel, P.E.; Omacini, M.; Sadras, V.O.; Ghersa, C.M. The interplay between the effectiveness of the grass-endophyte mutualism and the genetic variability of the host plant. Evol. Appl. 2010, 3, 538–546. [Google Scholar] [CrossRef] [PubMed]
  133. Gundel, P.E.; Martinez-Ghersa, M.A.; Omacini, M.; Cuyeu, R.; Pagano, E.; Rios, R.; Ghersa, C.M. Mutualism effectiveness and vertical transmission of symbiotic fungal endophytes in response to host genetic background. Evol. Appl. 2012, 5, 838–849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Gagic, M.; Faville, M.J.; Zhang, W.; Forester, N.T.; Rolston, M.P.; Johnson, R.D.; Ganesh, S.; Koolaard, J.P.; Easton, H.S.; Hudson, D.; et al. Seed transmission of Epichloë endophytes in Lolium perenne is heavily influenced by host genetics. Front. Plant Sci. 2018, 9, 1580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Tian, P.; Le, T.-N.; Smith, K.F.; Forster, J.W.; Guthridge, K.M.; Spangenberg, G.C. Stability and viability of novel perennial ryegrass host–Neotyphodium endophyte associations. Crop Pasture Sci. 2013, 64, 39–50. [Google Scholar] [CrossRef]
  136. Hume, D.E.; Schmid, J.; Rolston, M.P.; Vijayan, P.; Hickey, M.J. Effect of climatic conditions on endophyte and seed viability in stored ryegrass seed. Seed Sci. Technol. 2011, 39, 481–489. [Google Scholar] [CrossRef]
  137. Bacon, C.W.; Porter, J.K.; Robbins, J.D.; Luttrell, E.S. Epichloë typhina from toxic tall fescue grasses. Appl. Environ. Microbiol. 1977, 34, 576–581. [Google Scholar] [CrossRef] [Green Version]
  138. Hoveland, C.S. Endophyte-research and impact. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium, Soest, Germany, 27–29 September 2000; Paul, V.H., Dapprich, P.D., Eds.; pp. 1–8. [Google Scholar]
  139. Fletcher, L.R.; Harvey, I.C. An association of a Lolium endophyte with ryegrass staggers. N. Z. Vet. J. 1981, 29, 185–186. [Google Scholar] [CrossRef]
  140. Siegel, M.R.; Latch, G.C.M.; Johnson, M.C. Acremonium fungal endophytes of tall fescue and perennial ryegrass—Significance and control. Plant Dis. 1985, 69, 179–183. [Google Scholar]
  141. Comis, D. The grass farmers love to hate. Agricl. Res. 2000, 48, 4–7. [Google Scholar]
  142. Schmidt, S.P.; Hoveland, C.S.; Clark, E.M.; Davis, N.D.; Smith, L.A.; Grimes, H.W.; Holliman, J.L. Association of an endophytic fungus with fescue toxicity in steers fed Kentucky 31 tall fescue seed or hay. J. Anim. Sci. 1982, 55, 1259–1263. [Google Scholar] [CrossRef] [Green Version]
  143. Hemken, R.W.; Jackson, J.A.; Boling, J.A. Toxic factors in tall fescue. J. Anim. Sci. 1984, 58, 1011–1016. [Google Scholar] [CrossRef] [PubMed]
  144. Bacon, C.W. Toxic endophyte-infected tall fescue and range grasses: Historic perspectives. J Anim. Sci. 1995, 73, 861–870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Hemken, R.W.; Bull, L.S.; Boling, J.; Kane, E.; Bush, L.P.; Buckner, R.C. Summer fescue toxicosis in lactating dairy cows and sheep fed experimental strains of ryegrass-tall fescue hybrids. J. Anim. Sci. 1979, 49, 641–646. [Google Scholar] [CrossRef] [PubMed]
  146. Gadberry, M.S.; Denard, T.M.; Spiers, D.E.; Piper, E.L. Effects of feeding ergovaline on lamb performance in a heat stress environment. J. Anim. Sci. 2003, 81, 1538–1545. [Google Scholar] [CrossRef]
  147. Zbib, N.; Repussard, C.; Tardieu, D.; Priymenko, N.; Domange, C.; Guerre, P. Ergovaline in tall fescue and its effect on health, milk quality, biochemical parameters, oxidative status, and drug metabolizing enzymes of lactating ewes. J. Anim. Sci. 2014, 92, 5112–5123. [Google Scholar] [CrossRef] [Green Version]
  148. Browning, R.; Donnelly, B., Jr.; Payton, T.; Pandya, P.; Byars, M. Body weight gain and voluntary intake in meat goat does fed endophyte-infected and endophyte-free tall fescue seed. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 427–429.
  149. Cross, D.L. Fescue Toxicosis in Horses. In Neotyphodium/Grass Interactions; Bacon, C.W., Hill, N.S., Eds.; Springer: Boston, MA, USA, 1997; pp. 289–309. [Google Scholar]
  150. Wolfe, B.A.; Bush, M.; Monfort, S.L.; Mumford, S.L.; Pessier, A.; Montali, R.J. Abdominal lipomatosis attributed to tall fescue toxicosis in deer. J. Am. Vet. Med. Assoc. 1998, 213, 783–1754. [Google Scholar]
  151. Sampaio, N.; Gishen, M.; Reed, K.; Brown, M.; Gregory, D.; Munyard, K. The occurrence and severity of grass toxicoses in Australian alpaca (Vicugna pacos) herds. Aust. J. Exp. Agric. 2008, 48, 1099–1104. [Google Scholar] [CrossRef]
  152. Tor-Agbidye, J.; Blythe, L.L.; Craig, A.M. Correlation of endophyte toxins (ergovaline and lolitrem B) with clinical disease: Fescue foot and perennial ryegrass staggers. Vet. Human Toxicol. 2011, 43, 140–146. [Google Scholar]
  153. Ayers, A.W.; Hill, N.S.; Rottinghaus, G.E.; Stuedemann, J.A.; Thompson, F.N.; Purinton, P.T.; Seman, D.H.; Dawe, D.L.; Parks, A.H.; Ensley, D. Ruminal metabolism and transport of tall fescue ergot alkaloids. Crop Sci. 2009, 49, 2309–2316. [Google Scholar] [CrossRef] [Green Version]
  154. Klotz, J.L.; Kirch, B.H.; Aiken, G.E.; Bush, L.P.; Strickland, J.R. Contractile response of fescue-näive bovine lateral saphenous veins to increasing concentrations of tall fescue alkaloids. J. Anim. Sci. 2010, 88, 408–415. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Strickland, J.R.; Looper, M.L.; Matthews, J.C.; Rosenkrans, C.F., Jr.; Flythe, M.D.; Brown, K.R. BOARD-INVITED REVIEW: St. Anthony’s Fire in livestock: Causes, mechanisms, and potential solutions. J. Anim. Sci. 2011, 89, 1603–1626. [Google Scholar] [CrossRef] [PubMed]
  156. Van Heeswijck, R.; McDonald, G. Acremonium endophytes in perennial ryegrass and other pasture grasses in Australia and New Zealand. Aust. J. Agric. Res. 1992, 43, 1683–1709. [Google Scholar] [CrossRef]
  157. Di Menna, M.E.; Finch, S.C.; Popay, A.J.; Smith, B.L. A review of the Neotyphodium lolii/Lolium perenne symbiosis and its associated effects on animal and plant health, with particular emphasis on ryegrass staggers. N. Z. Vet. J. 2012, 20, 315–328. [Google Scholar] [CrossRef] [PubMed]
  158. Gallagher, R.T.; White, E.P.; Mortimer, P.H. Ryegrass staggers: Isolation of potent neurotoxins lolitrem A and lolitrem B from staggers producing pastures. N. Z. Vet. J. 1981, 29, 189–190. [Google Scholar] [CrossRef]
  159. Gilruth, J.A. Menings-encephalitis (stomach staggers) of horses cattle and sheep. Annu. Rep. N. Z. Depart. Gric. 1906, 14, 293–297. [Google Scholar]
  160. McLeay, L.M.; Smith, B.L. Effects of the mycotoxins lolitrem B and paxilline on gastrointestinal smooth muscle, the cardiovascular and respiratory systems, and temperature in sheep. In Proceedings of the Ryegrass Endophyte: An Essential New Zealand Symbiosis; Napier, New Zealand, 8 October 1999, Woodfield, D.R., Matthew, C., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 7, pp. 69–75.
  161. Guerre, P. Ergot alkaloids produced by endophytic fungi of the genus Epichloë. Toxins 2015, 7, 773–790. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Fletcher, L.R. Heat stress in lambs grazing ryegrass with different endophytes. In Proceedings of the 2nd International Symposium on Acremonium/Grass Interactions, Palmerston North, New Zealand, 2–12 February 1993; AgResearch: Hamilton, New Zealand, 1993; pp. 114–118. [Google Scholar]
  163. Fletcher, L.R.; Markham, L.J.; White, S.R. Endophytes and heat tolerance in lambs grazing perennial ryegrass. Proc. N. Z. Grassl. Assoc. 1994, 56, 265–270. [Google Scholar] [CrossRef]
  164. Stuedemann, J.A.; Thompson, F.N. Management strategies and potential opportunities to reduce the effects of endophyte-infested tall fescue on animal performance. In Proceedings of the 2nd International Symposium Acremonium/Grass Interactions, Palmerston North, New Zealand, 2–12 February 1993; Hume, D.E., Latch, G.C.M., Easton, H.S., Eds.; Plenary Papers. pp. 103–114. [Google Scholar]
  165. Fletcher, L.R.; Sutherland, B.L.; Fletcher, G.C. The impact of endophyte on the health and productivity of sheep grazing ryegrass-based pasture. In Ryegrass Endophyte: An Essential New Zealand Symbiosis; NZ Grassland Association Grassland Research and Practice Series; Woodfield, D.R., Matthew, C., Eds.; NZ Grassland Association: Napier, New Zealand, 1999; Volume 7, pp. 11–17. [Google Scholar]
  166. Reed, K.F.M. Pasture ryegrass toxins in Australian pasture. In Proceedings of the Symposium, Meat and Livestock Australia, Attwood, Australia, 18 March 2005; Perennial Ryegrass Toxicosis in Australia. pp. 11–17. [Google Scholar]
  167. Pennell, C.G.L.; Popay, A.J.; Rolston, M.P.; Townsend, R.J.; Lloyd-West, C.M.; Card, S.D. Avanex unique endophyte technology: Reduced insect food source at airports. Environ. Entomol. 2016, 45, 101–108. [Google Scholar] [CrossRef]
  168. Finch, S.C.; Munday, J.S.; Munday, R.; Kerby, J.W.F. Short term toxicity studies of loline alkaloids in mice. Food Chem. Toxicol. 2016, 94, 243–249. [Google Scholar] [CrossRef]
  169. Pownall, D.B.; Familton, A.S.; Field, R.J.; Fletcher, L.R.; Lane, G.A. The effect of peramine ingestion in pen-fed lambs. Proc. N. Z. Soc. Anim. Producti. 1995, 55, 186. [Google Scholar]
  170. Clay, K.; Schardl, C. Evolutionary origins and ecological consequences of endophyte symbiosis with grasses. Am. Nat. 2002, 160, S99–S127. [Google Scholar] [CrossRef] [PubMed]
  171. Ball, O.J.-P.; Tapper, B.A. The production of loline alkaloids in artificial and natural grass/endophyte associations. In Proceedings of the 52nd New Zealand Plant Protection Conference, Auckland Airport Royal-Centra, Auckland, New Zealand, 9–12 August 1999; pp. 264–269. [Google Scholar]
  172. Adhikari, K.B.; Boelt, B.; Fomsgaard, I.S. Identification and quantification of loline-type alkaloids in endophyte-infected grasses by LC-MS/MS. J. Agric. Food Chem. 2016, 64, 6212–6218. [Google Scholar] [CrossRef]
  173. Patchett, B.J. Loline alkaloids: Analysis and Effect on Sheep and Pasture Insects. Ph.D. Thesis, Lincoln University, Lincoln, New Zealand, 2007. [Google Scholar]
  174. Bourke, C.A.; Hunt, E.; Watson, R. Fescue-associated oedema of horses grazing on endophyte inoculated tall fescue grass (Festuca arundinacea) pastures. Aust. Vet. J. 2009, 87, 492–498. [Google Scholar] [CrossRef] [PubMed]
  175. Finch, S.C.; Munday, J.S.; Sutherland, B.L.; Vlaming, J.B.; Fletcher, L.R. Further investigation of equine fescue oedema induced by Mediterranean tall fescue (Lolium arundinaceum) infected with selected fungal endophytes (Epichloë coenophiala). N. Z. Vet. J. 2017, 65, 322–326. [Google Scholar] [CrossRef]
  176. Gooneratne, S.R.; Patchett, B.J.; Wellby, M.; Fletcher, L.R. Excretion of loline alkaloids in urine and faeces of sheep dosed with meadow fescue (Festuca pratensis) seed containing high concentrations of loline alkaloids. N. Z. Vet. J. 2012, 60, 176–182. [Google Scholar] [CrossRef]
  177. Schardl, C.; Panaccione, D.; Tudzynski, P. Ergot alkaloids—Biology and molecular biology. Alkaloids Chem. Biol. 2007, 63, 45–86. [Google Scholar]
  178. Piper, E.; Denard, T.; Johnson, Z.; Flieger, M. Effect of chanoclavine on in vitro prolactin release. In Proceedings of the 4th International Neoptyphodium/Grass interactions Symposium, Soest, Germany, 27–29 September 2000; Paul, V.H., Dapprich, P.D., Eds.; pp. 531–534. [Google Scholar]
  179. Finch, S.C.; Munday, J.S.; Sprosen, J.M.; Bhattarai, S. Toxicity studies of chanoclavine in mice. Toxins 2019, 11, 249. [Google Scholar] [CrossRef] [Green Version]
  180. Petroski, R.J.; Powell, R.G.; Clay, K. Alkaloids of Stipa robusta (sleepygrass) infected with an Acremonium endophyte. Nat. Toxins 1992, 1, 84–88. [Google Scholar] [CrossRef]
  181. Miles, C.O.; Lane, G.A.; di Menna, M.E.; Garthwaite, I.; Piper, E.L.; Ball, O.J.-P.; Latch, G.C.M.; Allen, J.M.; Hunt, M.B.; Bush, L.P.; et al. High levels of ergonovine and lysergic acid amide in toxic Achnatherum inebrians accompany infection by an Acremonium-like endophytic fungus. J. Agric. Food Chem. 1996, 44, 1285–1290. [Google Scholar] [CrossRef]
  182. Browning, R.; Leite-Browning, M.L. Effect of ergotamine and ergonovine on thermal regulation and cardiovascular function in cattle. J. Anim. Sci. 1997, 75, 176–181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Klotz, J.L.; Bush, L.P.; Smith, D.L.; Schafer, W.D.; Smith, L.L.; Arrington, B.C.; Strickland, J.R. Ergovaline-induced vasoconstriction in an isolated bovine lateral saphenous vein bioassay. J. Anim. Sci. 2007, 85, 2330–2336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Coulfal-Majewski, S.; Stanford, K.; McAllister, T.; Blakley, B.; McKinnon, J.; Chaves, A.V.; Wang, Y. Impacts of cereal ergot in food animal production. Front. Vet. Sci. 2016, 3, 15. [Google Scholar] [CrossRef] [Green Version]
  185. Solomons, R.N.; Oliver, J.W.; Linnabary, R.D. Reactivity of the dorsal pedal vein of cattle to selected alkaloids associated with Acremonium coenophialum-infected fescue grass. Am. J. Vet. Res. 1989, 50, 235–238. [Google Scholar]
  186. Porter, J.K. Chemical constituents of grass endophytes. In Biotechnology of Endophytic Fungi of Grasses; Bacon, C.W., White, J.F., Jr., Eds.; CRC Press: Boca Raton, FL, USA, 1994; pp. 103–123. [Google Scholar]
  187. Strickland, J.R.; Cross, D.L.; Birrenkott, G.P.; Grimes, L.W. Effect of ergovaline, loline, and dopamine antagonists on rat pituitary cell prolactin-release in-vitro. Am. J. Vet. Res. 1994, 55, 716–721. [Google Scholar]
  188. Easton, H.S.; Lane, G.A.; Tapper, B.A.; Keogh, R.G.; Cooper, B.M.; Blackwell, M.; Anderson, M.; Fletcher, L.R. Ryegrass endophyte-related heat stress in cattle. Proc. N. Z. Grassl. Assoc. 1996, 57, 37–41. [Google Scholar] [CrossRef]
  189. De Lorme, M.J.M.; Lodge-Ivey, S.L.; Craig, A.M. Physiological and digestive effects of Neotyphodium coenophialum infected tall fescue fed to lambs. J. Anim. Sci. 2007, 85, 1199–1206. [Google Scholar] [CrossRef] [Green Version]
  190. Klotz, J.L.; Bush, L.P.; Smith, D.L.; Schafer, W.D.; Smith, L.L.; Vevoda, A.O.; Craig, A.M.; Arrington, B.C.; Strickland, J.R. Assessment of vasoconstrictive potential of d-lysergic acid using an isolated bovine lateral saphenous vein bioassay. J. Anim. Sci. 2006, 84, 3167–3175. [Google Scholar] [CrossRef] [Green Version]
  191. Fletcher, L.R.; Sutherland, B.L. Sheep responses to grazing ryegrass with AR37 endophyte. Proc. N. Z. Grassl. Assoc. 2009, 71, 127–132. [Google Scholar] [CrossRef]
  192. Babu, J.V.; Popay, A.J.; Miles, C.O.; Wilkins, A.L.; di Menna, M.E.; Finch, S.C. Identification and structure elucidation of janthitrems A and D from Penicillium janthinellum and determination of the tremorgenic and anti-insect activity of janthitrems A and B. J. Agric. Food Chem. 2018, 66, 13116–13125. [Google Scholar] [CrossRef] [PubMed]
  193. Munday-Finch, S.C.; Wilkins, A.L.; Miles, C.O.; Ede, R.M.; Thomson, R.A. Structure elucidation of lolitrem F, a naturally-occurring stereoisomer of the tremorgenic mycotoxin lolitrem B, isolated from Lolium perenne infected with Acremonium lolii. J. Agric. Food Chem. 1996, 44, 2782–2788. [Google Scholar] [CrossRef]
  194. Gallagher, R.T.; Campbell, A.G.; Hawkes, A.D.; Holland, P.T.; McGaveston, D.A.; Pansier, E.A. Ryegrass staggers: The presence of lolitrem neurotoxins in perennial ryegrass seed. N. Z. Vet. J. 1982, 30, 183–184. [Google Scholar] [CrossRef] [PubMed]
  195. Gallagher, R.T.; Hawkes, A.D. Estimation of neurotoxin levels in perennial ryegrass by mouse bioassay. N. Z. J. Agric. Res. 1985, 28, 427–431. [Google Scholar] [CrossRef]
  196. Munday-Finch, S.C.; Miles, C.O.; Wilkins, A.L.; Hawkes, A.D. Isolation and structure elucidation of lolitrem A, a tremorgenic mycotoxin from perennial ryegrass infected with Acremonium lolii. J. Agric. Food Chem. 1995, 43, 1283–1288. [Google Scholar] [CrossRef]
  197. Blythe, L.; Estill, C.; Males, J.; Craig, A.M. Determination of the Toxic Threshold of Lolitrem B in Cattle Eating Endophyte-Infected Perennial Ryegrass. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 399–402.
  198. Guerre, P. Lolitrem B and indole diterpene alkaloids produced by endophytic fungi of the genus Epichloë and their toxic effects in livestock. Toxins 2016, 8, 47. [Google Scholar] [CrossRef] [Green Version]
  199. Miles, C.O.; Munday, S.C.; Wilkins, A.L.; Ede, R.M.; Towers, N.R. Large-scale isolation of lolitrem B and structure determination of lolitrem E. J. Agric. Food Chem. 1994, 42, 1488–1492. [Google Scholar] [CrossRef]
  200. Nakazawa, J.; Yajima, J.; Usui, T.; Ueki, M.; Takatsuki, A.; Imoto, M.; Toyoshima, Y.Y.; Osada, H. A novel action of terpendole E on the motor activity of mitotic Kinesin Eg5. Chem. Biol. 2003, 10, 131–137. [Google Scholar] [CrossRef] [Green Version]
  201. Miles, C.O.; Wilkins, A.L.; Gallagher, R.T.; Hawkes, A.D.; Munday, S.C.; Towers, N.R. Synthesis and tremorgenicity of paxitriols and lolitriol: Possible biosynthetic precursors of lolitrem B. J. Agric. Food Chem. 1992, 40, 234–238. [Google Scholar] [CrossRef]
  202. Cole, R.J.; Kirksey, J.W.; Wells, J.M. A new tremorgenic metabolite from Penicillium paxilli. Can. J. Microb. 1974, 20, 1159–1162. [Google Scholar] [CrossRef]
  203. Gallagher, R.T.; Hawkes, A.D. The potent tremorgenic neurotoxins lolitrem B and aflatrem: A comparison of the tremor response in mice. Experientia 1986, 42, 823–825. [Google Scholar] [CrossRef] [PubMed]
  204. Fletcher, L.R.; Popay, A.J.; Tapper, B.A. Evaluation of several lolitrem-free endophyte/perennial ryegrass combinations. Proc. N. Z. Grassl. Assoc. 1991, 53, 215–219. [Google Scholar] [CrossRef]
  205. Fletcher, L.R.; Garthwaite, I.; Towers, N. Ryegrass staggers in the absence of lolitrem B. In Proceedings of the 2nd International Symposium Acremonium/Grass Interactions, Palmerston North, New Zealand, 2–12 February 1993; Hume, D.E., Ed.; pp. 119–121. [Google Scholar]
  206. Munday-Finch, S.C.; Wilkins, A.L.; Miles, C.O.; Tomoda, H.; Omura, S. Isolation and structure elucidation of lolilline, a possible biosynthetic precursor of the lolitrem family of tremorgenic mycotoxins. J. Agric. Food Chem. 1997, 45, 199–204. [Google Scholar] [CrossRef]
  207. Gatenby, W.A.; Munday-Finch, S.C.; Wilkins, A.L.; Miles, C.O. Terpendole M, a novel indole-diterpenoid isolated from Lolium perenne infected with the endophytic fungus Neotyphodium lolii. J. Agric. Food Chem. 1999, 47, 1092–1097. [Google Scholar] [CrossRef] [PubMed]
  208. Pownall, D.B.; Lucas, R.J.; Familton, A.S.; Love, B.G.; Hines, S.E.; Fletcher, L.R. The relationship between staggers and diarrhoea in lambs grazing different components of endophyte-infected ryegrass. Proc. N. Z. Soc. Anim. Product. 1993, 53, 19–22. [Google Scholar]
  209. Fletcher, L.R.; Easton, H.S. The evaluation of a Lolium endophyte with ryegrass staggers. N. Z. Vet. J. 1997, 29, 185–186. [Google Scholar] [CrossRef]
  210. Schardl, C.L.; Grossman, R.B.; Nagabhyru, P.; Faulkner, J.R.; Mallik, U.P. Loline alkaloids: Currencies of mutualism. Phytochemist 2007, 68, 980–996. [Google Scholar] [CrossRef]
  211. Hume, D.E.; Card, S.D.; Rolston, M.P. Effects of storage conditions on endophyte and seed viability in pasture grasses. In Proceedings of the 22nd International Grassland Congress, Sydney, Australia, 15–19 September 2013; pp. 405–408. [Google Scholar]
  212. Latch, G.C.M. Physiological interactions of endophytic fungi and their hosts. Biotic stress tolerance imparted to grasses by endophytes. Agric. Ecosyst. Environ. 1993, 44, 143–156. [Google Scholar] [CrossRef]
  213. Joost, R.E. Acremonium in fescue and ryegrass: Boon or bane? A Review. J. Anim. Sci. 1995, 73, 881–888. [Google Scholar] [CrossRef]
  214. Kuldau, G.; Bacon, C.W. Clavicipitaceous endophytes: Their ability to enhance resistance of grasses to multiple stresses. Biol. Control. 2008, 46, 57–71. [Google Scholar] [CrossRef]
  215. Hume, D.E.; Sewell, J.C. Agronomic advantages conferred by endophyte infection of perennial ryegrass (Lolium perenne L.) and tall fescue (Festuca arundinacea Schreb.) in Australia. Crop Pasture Sci. 2014, 65, 747–757. [Google Scholar] [CrossRef]
  216. Thom, E.R.; Popay, A.J.; Waugh, C.D.; Minneé, E.M.K. Impact of novel endophytes in perennial ryegrass on herbage production and insect pests from pastures under dairy cow grazing in northern New Zealand. Grass Forage Sci. 2014, 69, 191–204. [Google Scholar] [CrossRef]
  217. Crutchfield, B.A.; Potter, D.A. Damage Relationships of Japanese Beetle and Southern Masked Chafer (Coleoptera: Scarabaeidae) Grubs in Cool-Season Turfgrasses. J. Econ. Entomol. 1995, 88, 1049–1056. [Google Scholar] [CrossRef]
  218. Arachevaleta, M.; Bacon, C.W.; Plattner, R.D.; Hoveland, C.S.; Radcliffe, D.E. Accumulation of ergopeptide alkaloids in symbiotic tall fescue grown under deficits of soil water and nitrogen fertilizer. Appl. Environ. Microbiol. 1992, 58, 857–861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Bacon, C.W. Abiotic stress tolerances (moisture, nutrients) and photosynthesis in endophyte-infected tall fescue. Agric. Ecosyst. Environ. 1993, 44, 123–141. [Google Scholar] [CrossRef]
  220. West, C.P. Physiology and drought tolerance of endophyte-infected grasses. In Biotechnology of endophytic fungi of grasses. In Biotechnology of Endophytic Fungi of Grasses; Bacon, C.W., White, J.F., Jr., Eds.; CRC Press: Boca Raton, FL, USA, 1994; pp. 87–99. [Google Scholar]
  221. Elbersen, H.W.; West, C.P. Growth and water relations of field-grown tall fescue as influenced by drought and endophyte. Grass Forage Sci. 1996, 51, 333–342. [Google Scholar] [CrossRef]
  222. Buck, G.W.; West, C.P.; Elbersen, H.W. Endophyte effect on drought tolerance in diverse Festuca species. In Neotyphodium/Grass Interactions; Bacon, C., Hill, N., Eds.; Springer: New York, NY, USA, 1997; pp. 141–143. [Google Scholar] [CrossRef]
  223. Elmi, A.A.; West, C.P. Endophyte infection effects on stomatal conductance, osmotic adjustment and drought recovery of tall fescue. New Phytol. 1995, 131, 61–67. [Google Scholar] [CrossRef]
  224. Cheplick, G.P.; Perera, A.; Koulouris, K. Effect of drought on the growth of Lolium perenne genotypes with and without fungal endophytes. Funct. Ecol. 2000, 14, 657–667. [Google Scholar] [CrossRef]
  225. Malinowski, D.P.; Belesky, D.P. Adaptations of endophyte-infected cool-season grasses to environmental stresses: Mechanisms of drought and mineral stress tolerance. Crop Sci. 2000, 40, 923–940. [Google Scholar] [CrossRef]
  226. Nagabhyru, P.; Dinkins, R.D.; Wood, C.L.; Bacon, C.W.; Schardl, C.L. Tall fescue endophyte effects on tolerance to water-deficit stress. BMC Plant Biol. 2013, 13, 127. [Google Scholar] [CrossRef] [Green Version]
  227. Xu, L.; Li, X.; Han, L.; Li, D.; Song, G. Epichloë endophyte infection improved drought and heat tolerance of tall fescue through altered antioxidant enzyme activity. Eur. J. Hortic. Sci. 2017, 82, 90–97. [Google Scholar] [CrossRef]
  228. Barker, D.J.; Davies, E.; Lane, G.A.; Latch, G.C.M.; Nott, H.M.; Tapper, B.A. Effect of water deficit on alkaloid concentrations in perennial ryegrass endophyte associations. In Proceedings of the International Symposium on Acremonium Interaction, Palmerston North, New Zealand, 2–12 February 1993; pp. 67–71. [Google Scholar]
  229. Hahn, H.; McManus, M.T.; Warnstorff, K.; Monahan, B.J.; Young, C.A.; Davies, E.; Tapper, B.A.; Scott, B. Neotyphodium fungal endophytes confer physiological protection to perennial ryegrass (Lolium perenne L.) subjected to a water deficit. Environ. Exp. Bot. 2008, 63, 183–199. [Google Scholar] [CrossRef]
  230. Davitt, A.; Chen, C.; Rudgers, J.A. Understanding context-dependency in plant–microbe symbiosis: The influence of abiotic and biotic contexts on host fitness and the rate of symbiont transmission. Environ. Exp. 2011, 71, 137–145. [Google Scholar] [CrossRef]
  231. Barker, D.J.; Hume, D.E.; Quigley, P.E. Negligible physiological responses to water deficit in endophyte-infected and uninfected perennial ryegrass. In Proceedings of the 3rd International Symposium on Neotyphodium/Grass Interactions, Athens, GA, USA, 28–31 May 1997; pp. 137–139. [Google Scholar]
  232. Belesky, D.P.; Stringer, W.C.; Hill, N.S. Influence of endophyte and water regime upon tall fescue accessions. 1. Growth characteristics. Ann. Bot. 1989, 63, 495–503. [Google Scholar]
  233. White, R.H.; Engelke, M.C.; Morton, S.J.; Johnson-Cicalese, J.M.; Ruemmele, B.A. Acremonium endophyte effects on tall fescue drought tolerance. Crop Sci. 1992, 32, 1392–1396. [Google Scholar] [CrossRef]
  234. Hill, N.S.; Pachon, J.G.; Bacon, C.W. Acremonium coenophialum-mediated short- and long-term drought acclimation in tall fescue. Crop Sci. 1996, 36, 665–672. [Google Scholar] [CrossRef]
  235. Marks, S.; Clay, K. Physiological responses of Festuca arundinacea to fungal endophyte infection. New Phytol. 1996, 133, 727–733. [Google Scholar] [CrossRef]
  236. Assuero, S.G.; Matthew, C.; Kemp, P.D.; Latch, G.C.M.; Barker, D.J.; Haslett, S.J. Morphological and physiological effects of water deficit and endophyte infection on contrasting tall fescue cultivars. N. Z. J. Agric. Res. 2000, 43, 49–61. [Google Scholar] [CrossRef] [Green Version]
  237. Miranda, M.I.; Omacini, M.; Chaneton, E.J. Environmental context of endophyte symbioses: Interacting effects of water stress and insect herbivory. Int. J. Plant Sci. 2011, 172, 499–508. [Google Scholar] [CrossRef] [Green Version]
  238. Hall, S.L.; McCulley, R.L.; Barney, R.J.; Phillips, T.D. Does fungal endophyte infection improve tall fescue’s growth response to fire and water limitation? PLoS ONE 2014, 9, e86904. [Google Scholar] [CrossRef]
  239. Tian, Z.; Huang, B.; Belanger, F.C. Effects of Epichloë festucae fungal endophyte infection on drought and heat stress responses of strong creeping Red Fescue. J. Am. Soc. Hortic. Sci. 2015, 140, 257–264. [Google Scholar] [CrossRef]
  240. Reza Sabzalian, M.; Mirlohi, A. Neotyphodium endophytes trigger salt resistance in tall and meadow fescues. J. Plant Nutr. Soil Sci. 2010, 173, 952–957. [Google Scholar] [CrossRef]
  241. Yin, L.; Ren, A.; Wei, M.; Wu, L.; Zhou, Y.; Li, X.; Gao, Y. Neotyphodium coenophialum-infected tall fescue and its potential application in the phytoremediation of saline soils. Int. J. Phytoremediat 2014, 16, 235–246. [Google Scholar] [CrossRef]
  242. Wang, J.; Tian, P.; Christensen, M.J.; Zhang, X.; Li, C.; Nan, Z. Effect of Epichloë gansuensis endophyte on the activity of enzymes of nitrogen metabolism, nitrogen use efficiency and photosynthetic ability of Achnatherum inebrians under various NaCl concentrations. Plant Soil 2018. [Google Scholar] [CrossRef]
  243. Malinowski, D.P.; Belesky, D.P. Neotyphodium coenophialum-endophyte infection affects the ability of tall fescue to use sparingly available phosphorus. J. Plant Nutr. 1999, 22, 835–853. [Google Scholar] [CrossRef]
  244. Malinowski, D.P.; Alloush, G.A.; Belesky, D.P. Leaf endophyte Neotyphodium coenophialum modifies mineral uptake in tall fescue. Plant Soil 2000, 227, 115–126. [Google Scholar] [CrossRef]
  245. Ren, A.Z.; Li, C.A.; Gao, Y.B. Endophytic fungus improves growth and metal uptake of Lolium arundinaceum Darbyshire Ex. Schreb. Int. J. Phytoremediat 2011, 13, 233–243. [Google Scholar] [CrossRef] [PubMed]
  246. Mirzahossini, Z.; Shabani, L.; Sabzalian, M.R.; Sharifi-Tehrani, M. ABC transporter and metallothionein expression affected by NI and Epichloë endophyte infection in tall fescue. Ecotox. Environ. Safe 2015, 120, 13–19. [Google Scholar] [CrossRef] [PubMed]
  247. Gaynor, D.L.; Rowan, D.D. Insect resistance, animal toxicity and endophyte-infected grass. Proc. N. Z. Grassl. Assoc. 1986, 47, 115–120. [Google Scholar] [CrossRef]
  248. Popay, A.J.; Hume, D.E. Endophytes improve ryegrass persistence by controlling insects. In New Zealand Grassland Association Pasture Persistence; Mercer, C.F., Ed.; NZ Grassland Association: Napier, New Zealand, 2011; Volume 15, pp. 149–156. [Google Scholar]
  249. Zydenbos, S.M.; Barratt, B.I.P.; Bell, N.L.; Ferguson, C.M.; Gerard, P.J.; McNeill, M.R.; Phillips, C.B.; Townsend, R.J.; Jackson, T.A. The impact of invertebrate pests on pasture persistence and their interrelationship with biotic and abiotic factors. Pasture Persistence N. Z. Grassl. Res. Pract. Ser. 2011, 15, 109–118. [Google Scholar]
  250. Popay, A.J.; Rowan, D.D. Endophytic fungi as mediators of plant insect interactions. In Insect-Plant Interactions; Bernays, E.A., Ed.; CRC Press: Boca Raton, FL, USA, 1994; Volume 5, pp. 83–103. [Google Scholar]
  251. Bush, L.P.; Wilkinson, H.H.; Schardl, C.L. Bioprotective alkaloids of grass-fungal endophyte symbioses. Plant Physiol. 1997, 114, 1–7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Breen, J.P. Acremonium endophyte interactions with enhanced plant resistance to insects. Ann. Rev. Entomol. 1994, 39, 401–423. [Google Scholar] [CrossRef]
  253. Ahmad, S.; Govindarajan, S.; Funk, C.R.; Johnson-Cicalese, J.M. Fatality of house crickets on perennial ryegrass infected with a fungal endophyte. Entomol. Exp. Appl. 1985, 39, 183–190. [Google Scholar] [CrossRef]
  254. Bryant, R.H.; Cameron, N.E.; Edwards, G.R. Response of black beetle and red-headed pasture cockchafer larvae to loline alkaloids in meadow fescue roots. Proc. N. Z. Plant Protect. 2010, 63, 219–223. [Google Scholar]
  255. Muegge, M.A.; Quisenberry, S.S.; Bates, G.E.; Joost, R.E. Influence of Acremonium infection and pesticide use on seasonal abundance of leafhoppers and froghoppers (Homoptera: Cicadellidae; Cercopidae) in tall fescue. Environm. Entomol. 1991, 20, 1531–1536. [Google Scholar] [CrossRef]
  256. Potter, D.A.; Stokes, J.T.; Redmond, C.T.; Schardl, C.L.; Panaccione, D.G. Contribution of ergot alkaloids to suppression of a grass-feeding caterpillar assessed with gene-knockout endophytes in perennial ryegrass. Entomol. Exp. Appl. 2007, 126, 138–147. [Google Scholar] [CrossRef]
  257. Baldauf, M.W.; Mace, W.J.; Richmond, D.S. Endophyte mediated resistance to black cutworm as a function of plant cultivar and endophyte strain in tall fescue. Environ. Entomol. 2011, 40, 639–647. [Google Scholar] [CrossRef] [Green Version]
  258. Pennell, C.G.L.; Hume, D.E.; Ball, O.J.-P.; Easton, H.S.; Lyons, T.B. Effects of Neotyphodium lolii infection in ryegrass on root aphid and pasture mealy bug. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium, Soest, Germany, 27–29 September 2000; Paul, V.H., Dapprich, P.D., Eds.; pp. 465–469. [Google Scholar]
  259. Popay, A.J.; Silvester, W.B.; Gerard, P.J. New endophyte isolate suppresses root aphid, Aploneura lentisci, in perennial ryegrass. In Proceedings of the 5th International Symposium on Neotyphodium/Grass Interactions, Fayetteville, AR, USA, 23–26 May 2004; Kallenbach, R., Ed.; p. 317. [Google Scholar]
  260. Popay, A.J.; Hickey, M.J.; Stewart, A.V.; Hume, D.E. Potential use of a selected strain of fungal endophyte for insect control in turf ryegrass. In Proceedings of the 1st European Turfgrass Society Conference, Pisa, Italy, 19–20 May 2008; pp. 151–152. [Google Scholar]
  261. Popay, A.J.; Gerard, P.J. Cultivar and endophyte effects on a root aphid, Aploneura lentisci, in perennial ryegrass. N. Z. Plant Protect. 2007, 60, 223–227. [Google Scholar] [CrossRef] [Green Version]
  262. Moate, P.J.; Williams, S.R.O.; Grainger, C.; Hannah, M.C.; Mapleson, D.; Auldist, M.J.; Greenwood, J.S.; Popay, A.J.; Hume, D.E.; Mace, W.J.; et al. Effects of wild-type, AR1 and AR37 endophyte-infected perennial ryegrass on dairy production in Victoria, Australia. Anim. Prod. Sci. 2012, 52, 1117–1130. [Google Scholar] [CrossRef]
  263. Popay, A.J.; Cox, N.R. Aploneura lentisci (Homoptera: Aphididae) and its interactions with fungal endophytes in perennial ryegrass (Lolium perenne). Front. Plant Sci. 2016, 7, 1395. [Google Scholar] [CrossRef] [Green Version]
  264. Popay, A.J.; Hume, D.E.; Mace, W.J.; Faville, M.J.; Finch, S.C.; Cave, V. A root aphid, Aploneura lentisci is affected by Epichloë endophyte strain and impacts perennial ryegrass growth in the field. Crop Pasture Sci. 2021, in press. [Google Scholar]
  265. Schmidt, D. Effects of Acremonium uncinatum and a Phialophora-like endophyte on vigour, insect and disease resistance of meadow fescue. In Proceedings of the 2nd International Symposium on Acremonium/Grass Interactions, Palmerston Nth, New Zealand, 2–12 February 1993; pp. 185–188. [Google Scholar]
  266. Schmidt, D.; Guy, P.L. Effects of the presence of the endophyte Acremonium uncinatum and of an insecticidal treatment on seed production of meadow fescue. Rev. Suisse Agric. 1997, 29, 97–99. [Google Scholar]
  267. Jensen, J.G.; Popay, A.J. Reduction in root aphid populations by non-toxic endophyte strains in tall fescue. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Research and Practice Series; Volume 13, pp. 341–344.
  268. Popay, A.J.; Easton, H.S. Interactions between host plant genotype and Neotyphodium fungal endophytes affects insects. In Proceedings of the 13th Australasian Plant Breeding Conference, Christchurch, New Zealand, 18–21 April 2006; Mercer, C.F., Ed.; Breeding for Success: Diversity in Action. pp. 561–567. [Google Scholar]
  269. Pearson, W.D. The pasture mealy bug, Balanococcus poae (Maskell), in Canterbury: A preliminary report. In Proceedings of the 5th Australasian Conference of Grassland Invertebrate Ecology, Victoria, Australia, 5–19 August 1988; pp. 297–303. [Google Scholar]
  270. Popay, A.J.; Baltus, J.G.; Pennell, C.G.L. Insect resistance in perennial ryegrass infected with toxin-free Neotyphodium endophytes. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium, Soest, Germany, 27–29 September 2000; Paul, V.H., Dapprich, P.D., Eds.; pp. 187–193. [Google Scholar]
  271. Pennell, C.G.L.; Popay, A.J.; Ball, O.-P.; Hume, D.E.; Baird, D.B. Occurrence and impact of pasture mealybug (Balanococcus poae) and root aphid (Aploneura lentisci) on ryegrass (Lolium spp.) with and without infection by Neotyphodium fungal endophytes. N. Z. J. Agric.Res. 2005, 48, 329–337. [Google Scholar] [CrossRef] [Green Version]
  272. Pennell, C.G.L.; Ball, O. J-P. The effects of Neotyphodium endophytes in tall fescue on pasture mealy bug (Balanococcus poae). Proc. N. Z. Plant Protect. Conf. 1999, 52, 259–263. [Google Scholar] [CrossRef] [Green Version]
  273. Saha, D.C.; Johnson-Cicalese, J.M.; Halisky, P.M.; van Heemstra, M.I.; Funk, C.R. Occurrence and significance of endophytic fungi in the fine fescues. Plant Dis. 1987, 71, 1021–1024. [Google Scholar] [CrossRef]
  274. Mathias, J.K.; Ratcliffe, R.H.; Hellman, J.L. Association of an endophytic fungus in perennial ryegrass and resistance to the hairy chinch bug (Hemiptera: Lygaeidae). J. Econ. Entomol. 1990, 83, 1640–1646. [Google Scholar] [CrossRef]
  275. Carrie`re, Y.; Bouchard, A.; Bourassa, S.; Brodeur, J. Effect of endophyte incidence in perennial ryegrass on distribution, host-choice, and performance of the hairy chinch bug (Hemiptera: Lygaeidae). J. Econ. Entomol. 1998, 91, 324–328. [Google Scholar] [CrossRef]
  276. Richmond, D.S.; Shetlar, D.J. Hairy chinch bug (Hemiptera: Lygaeidae) damage, population, density, and movement in relation to the incidence of perennial ryegrass infected by Neotyphodium endophytes. J. Econ. Entomol. 2000, 93, 1167–1172. [Google Scholar] [CrossRef]
  277. Yue, Q.; Johnson-Cicalese, J.; Gianfagna, T.J.; Meyer, W.A. Alkaloid production and chinch bug resistance in endophyte-inoculated chewings and strong creeping red fescues. J. Chem. Ecol. 2000, 26, 279–292. [Google Scholar] [CrossRef]
  278. Anderson, W.G.; Heng-Moss, T.M.; Baxendale, F.P. Evaluation of cool- and warm season grasses for resistance to multiple chinch bug (Hemiptera: Blissidae) species. J. Econ. Entomol. 2006, 99, 203–211. [Google Scholar] [CrossRef]
  279. Popay, A.J.; Latch, G.C.M. Prospects for utilising endophytes for grass resistance to insect pests in New Zealand. In Proceedings of the 6th Australasian Grassland Invertebrate Ecology Conference, Hamilton, New Zealand, 17–19 February 1993; pp. 129–155. [Google Scholar]
  280. Popay, A.J.; Ball, O.J.-P. The development of fungal endophytes as a pest management tool for New Zealand grasslands. In Proceedings of the 6th Australasian Applied Entomological Research Conference, Brisbane, Queensland, Australia, 29 September–2 October 1998; pp. 373–381. [Google Scholar]
  281. Fletcher, L.R.; Popay, A.J.; Stewart, A.V.; Tapper, B.A. Herbage and sheep production from meadow fescue with and without the endophyte Neotyphodium uncinatum. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium, Soest, Germany, 27–29 September 2000; Volker, P.H., Dapprich, P.D., Eds.; pp. 447–453. [Google Scholar]
  282. Popay, A.J.; Lane, G.A. The effect of crude extracts containing loline alkaloids on two New Zealand insect pests. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium, Soest, Germany, 27–29 September 2000; Paul, V.H., Dapprich, P.D., Eds.; pp. 471–475. [Google Scholar]
  283. Popay, A.J.; Townsend, R.J.; Fletcher, L.R. The effect of endophyte (Neotyphodium uncinatum) in meadow fescue on grass grub larvae. N. Z. Plant Protect. 2003, 5, 123–128. [Google Scholar] [CrossRef] [Green Version]
  284. Popay, A.J.; Tapper, B.A. Endophyte effects on consumption of seed and germinated seedlings of ryegrass and fescue by grass grub (Costelytra zealandica) larvae. Endophyte Symposium. N. Z. Grassl. Res. Pract. Ser. 2007, 13, 353–355. [Google Scholar]
  285. Popay, A.J.; Jensen, J.G.; Mace, W.J. Root herbivory: Grass species, Epichloë endophytes and moisture status make a difference. Microorganisms 2020, 8, 997. [Google Scholar] [CrossRef] [PubMed]
  286. Popay, A.J.; Mainland, R.A.; Sanders, A.J. The effects of endophytes in fescue grass on growth and survival of third instar grass grub larvae. In Proceedings of the 2nd International Symposium on Acremonium/Grass Interactions, Palmerston North, New Zealand, 2–12 February 1993; pp. 174–177. [Google Scholar]
  287. Jensen, J.G.; Miller, T.A.; Cave, V.M.; Johnson, R.D.; Scott, B.; Popay, A.J. Two novel Epichloë festucae var. lolii endophytes reduce larval populations of the wheat sheath miner Cerodontha australis (Diptera: Agromyzidae). J. Pest Sci. 2021, in press. [Google Scholar]
  288. Funk, C.R.; Halisky, P.M.; Johnson, M.C.; Siegel, M.R.; Stewart, A.V.; Ahmed, S.; Hurley, R.H.; Harvey, I.C. An endophytic fungus and resistance to sod webworms: Association in Lolium perenne L. Bio/Technology 1983, 1, 189–191. [Google Scholar] [CrossRef]
  289. Bush, L.P.; Fannin, F.F.; Siegel, M.R.; Dahlman, D.L.; Burton, H.R. Chemistry, occurrence and biological effects of saturated pyrrolizidine alkaloids associated with endophyte-grass interactions. Agric. Ecosyst. Environ. 1993, 44, 81–102. [Google Scholar] [CrossRef]
  290. Clement, S.L.; Pike, K.S.; Kaiser, W.J.; Wilson, A.D. Resistance of endophyte-infected plants of tall fescue and perennial ryegrass to the Russian wheat aphid (Homoptera: Aphididae). J. Kan. Entomol. Soc. 1990, 63, 646–654. [Google Scholar]
  291. Kindler, S.D.; Breen, J.P.; Springer, T.L. Reproduction and damage by Russian wheat aphid (Homoptera: Aphididae) as influenced by fungal endophytes and cool-season turfgrasses. J. Econ. Entomol. 1991, 84, 685–692. [Google Scholar] [CrossRef]
  292. Kirfman, G.W.; Brandenburg, R.L.; Garner, G.B. Relationship between insect abundance and endophyte infestation level in tall fescue in Missouri. J. Kansas Entomol. Soc. 1986, 59, 552–554. [Google Scholar]
  293. Cole, A.M.; Pless, C.D.; Gwinn, K.D. Survival of Drosophila melanogaster (Diptera: Drosophiliae) on Diets Containing Roots or Leaves of Acremonium-Infected or Non-Infected Tall Fescue. In Proceedings of the 1st International Symposium on Acremonium/Grass Interactions, Baton Rouge, LA, USA, 9 October–14 November 1990; Quisenberry, S.S., Joost, R.E., Eds.; p. 128. [Google Scholar]
  294. Koppenhofer, A.M.; Cowles, R.S.; Fuzy, E.M. Effects of turfgrass endophytes (Clavicipitaceae: Ascomycetes) on white grub (Coleoptera: Scarabaeidae) larval development and field populations. Environ. Entomol. 2003, 32, 895–906. [Google Scholar] [CrossRef]
  295. Dymock, J.J.; Rowan, D.D.; McGee, I.R. Effects of endophyte-produced mycotoxins on Argentine stem weevil and the cutworm Graphania mutans. In Proceedings of the 5th Australasian Conference Grassland Invertebrate Ecology, Melbourne, Australia, 15–19 August 1988; Stahle, P.P., Ed.; Vie D & D Printing Pty.: Victoria, Australia, 1988; pp. 35–43. [Google Scholar]
  296. Ball, O.J.-P.; Prestidge, R.A. The effect of the endophytic fungus Acremonium lolii on adult black beetle (Heteronychus arator) feeding. Proc. N. Z. Plant Prot. Conf. 1992, 45, 201–204. [Google Scholar] [CrossRef] [Green Version]
  297. Ball, O.J.-P.; Prestidge, R.A. The use of the endophytic fungus Acremonium lolii as a biological control agent of black beetle, Heteronychus arator (Coleoptera: Scarabaeidae). In Proceedings of the 6th Australasian Conference on Grassland Invertebrate Ecology, Hamilton, New Zealand, 17–19 February 1993; Prestidge, R.A., Ed.; pp. 283–289. [Google Scholar]
  298. Ball, O.J.-P.; Miles, C.O.; Prestidge, R.A. Ergopeptine alkaloids and Neotyphodium lolii mediated resistance in perennial ryegrass against adult Heteronychus arator (Coleoptera:Scarabaeidae). J. Econ. Entomol. 1997, 90, 1382–1391. [Google Scholar] [CrossRef]
  299. Popay, A.J.; Hume, D.E.; Baltus, J.G.; Latch, G.C.M.; Tapper, B.A.; Lyons, T.B.; Cooper, B.M.; Pennell, C.G.; Eerens, J.P.J.; Marshall, S.L. Field performance of perennial ryegrass (Lolium perenne) infected with toxin-free fungal endophytes (Neotyphodium spp.). In Proceedings of the Ryegrass endophyte: An Essential New Zealand Symbiosis; Napier, New Zealand, 8 October 1999, Woodfield, D.R., Matthew, C., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 7, pp. 113–122.
  300. Popay, A.J.; Bonos, S.A. Biotic responses in endophytic grasses. In Neotyphodium in Cool-Season Grasses; Roberts, C.A., West, C.P., Spiers, D.E., Eds.; Blackwell Publishing: Ames, IA, USA, 2005; pp. 163–185. [Google Scholar]
  301. Ball, O.J.-P.; Christensen, M.J.; Prestidge, R.A. Effect of selected isolates of Acremonium endophytes on adult black beetle (Heteronychus arator) feeding. In Proceedings of the 47th New Zealand Plant Protection Conference, Auckland, New Zealand, 9–11 August 1994; pp. 227–231. [Google Scholar]
  302. Barker, G.M.; Patchett, B.J.; Cameron, N.E. Epichloë uncinata infection and loline content afford Festulolium grasses protection from black beetle (Heteronychus arator). N. Z. J. Agric. Res. 2015, 58, 35–56. [Google Scholar] [CrossRef]
  303. Barker, G.M.; Patchett, B.J.; Cameron, N.E. Epichloë uncinata infection and loline content protect Festulolium grasses from crickets (Orthoptera: Gryllidae). J. Econ. Entomol. 2015, 108, 789–797. [Google Scholar] [CrossRef] [PubMed]
  304. Prestidge, R.A.; Pottinger, R.P.; Barker, G.M. An association of Lolium endophyte with ryegrass resistance to Argentine stem weevil. In Proceedings of the 35th NZ Weed Pest Control Conference, Hamilton, New Zealand, 9–12 August 1982; pp. 119–122. [Google Scholar]
  305. Prestidge, R.A.; Lauren, D.R.; van der Zujpp, S.G.; di Menna, M.E. Isolation of feeding deterrents to Argentine stem weevil in cultures of endophytes of perennial ryegrass and tall fescue. N. Z. J. Agric. Res. 1985, 28, 87–92. [Google Scholar] [CrossRef] [Green Version]
  306. Gaynor, D.L.; Hunt, W.F. The relationship between nitrogen supply, endophytic fungus, and Argentine stem weevil resistance in ryegrasses. Proc. N. Z. Grassl. Assoc. 1983, 44, 257–263. [Google Scholar] [CrossRef]
  307. Barker, G.M.; Pottinger, R.P.; Addison, P.J. Effect of tall fescue and ryegrass endophytes on Argentine stem weevil. Proc. N.Z. Weed Pest Control Conf. 1983, 36, 216–219. [Google Scholar] [CrossRef]
  308. Barker, G.M.; Pottinger, R.P.; Addison, P.J.; Prestidge, R.A. Effect of Lolium endophyte fungus infection on behaviour of adult Argentine stem weevil. N. Z. J. Agric. Res. 1984, 27, 271–277. [Google Scholar] [CrossRef]
  309. Gaynor, D.L.; Rowan, D.D. Peramine–An Argentine stem weevil feeding deterrent from endophytic-infected ryegrass. In Proceedings of the 4th Australasian Conference on Grassland Invertebrate Ecology, Canterbury, New Zealand, 13–17 May 1985; pp. 338–343. [Google Scholar]
  310. Rowan, D.D.; Hunt, M.B.; Gaynor, D.L. Peramine, a novel insect feeding deterrent from ryegrass infected with the endophyte Acremonium loliae. J. Chem. Soc. Chem. Commun. 1986, 12, 935–936. [Google Scholar] [CrossRef]
  311. Rowan, D.D.; Dymock, J.J.; Brimble, M.A. Effect of fungal metabolite peramine and analogs on feeding and development of argentine stem weevil (Listronotus bonariensis). J. Chem. Ecol. 1990, 16, 1683–1695. [Google Scholar] [CrossRef]
  312. Dymock, J.J.; Latch, G.C.M.; Tapper, B.A. Novel combinations of endophytes in ryegrasses and fescues and their effects on Argentine stem weevil (Listronotus bonariensis) feeding. In Proceedings of the 5th Australasian Conference Grassland Invertebrate Ecology, Melbourne, Australia, 15–19 August 1988; Stahle, P.P., Ed.; Vie D & D Printing Pty.: Victoria, Australia, 1988; pp. 28–34. [Google Scholar]
  313. Popay, A.J.; Mainland, R.A. Seasonal damage by Argentine stem weevil to perennial ryegrass pastures with different levels of Acremonium lolii. In Proceedings of the 44th New Zealand Weed and Pest Control Conference, Tauranga, New Zealand, 13–15 August 1991; pp. 171–175. [Google Scholar]
  314. Popay, A.J.; Hume, D.E.; Mainland, R.A.; Saunders, C.J. Field resistance to Argentine stem weevil (Listronotus bonariensis) in different ryegrass cultivars infected with an endophyte deficient in lolitrem B. N. Z. J. Agric. Res. 1995, 38, 519–528. [Google Scholar] [CrossRef] [Green Version]
  315. Popay, A.J.; Hume, D.E.; Davis, K.L.; Tapper, B.A. Interactions between endophyte (Neotyphodium spp.) and ploidy in hybrid and perennial ryegrass cultivars and their effects on Argentine stem weevil (Listronotus bonariensis). N. Z. J. Agric. Res. 2003, 46, 311–319. [Google Scholar] [CrossRef]
  316. Popay, A.J.; Mace, W.J.; Finch, S.C.; Faville, M.J.; Jensen, J.G.; Cave, V.M. Epichloë fungal endophyte strains and their ryegrass (Lolium spp.) hosts affects resistance to Listronotus bonariensis (Coleoptera: Curculionidae). N. Z. J. Agric. Res. 2021, in press. [Google Scholar]
  317. Rowan, D.D.; Gaynor, D.L. Isolation of feeding deterrents against Argentine stem weevil from ryegrass infected with the endophyte Acremonium loliae. J. Chem. Ecol. 1986, 12, 647–658. [Google Scholar] [CrossRef] [PubMed]
  318. Prestidge, R.A.; Gallagher, R.T. Lolitrem B–A stem weevil toxin isolated from Acremonium-infected ryegrass. In Proceedings of the 38th NZ Weed and Pest Control Conference, Rotorua, New Zealand, 13–15 August 1985; pp. 38–40. [Google Scholar]
  319. Prestidge, R.A.; Gallagher, R.T. Endophyte fungus confers resistance to ryegrass: Argentine stem weevil studies. Ecol. Entomol. 1988, 13, 429–435. [Google Scholar] [CrossRef]
  320. Dymock, J.J.; Prestidge, R.A.; Rowan, D.D. The effects of lolitrem B on Argentine stem weevil larvae. In Proceedings of the 42nd NZ Weed and Pest Conference, Taranaki, New Zealand, 13–15 August 1989; pp. 73–75. [Google Scholar]
  321. Patchett, B.J.; Chapman, R.B.; Fletcher, L.R.; Gooneratne, S.R. Endophyte-infected Festuca pratensis containing loline alkaloids deter feeding by Listronotus bonariensis. N. Z. Plant Protect. 2008, 61, 5–209. [Google Scholar]
  322. Jensen, J.; Popay, A.; Tapper, B. Argentine stem weevil adults are affected by meadow fescue endophyte and its loline alkaloids. N. Z. Plant Prot. 2009, 62, 12–18. [Google Scholar] [CrossRef] [Green Version]
  323. Popay, A.J.; Tapper, B.A.; Podmore, C. Endophyte-infected meadow fescue and loline alkaloids affect argentine stem weevil larvae. N. Z. Plant Protect. 2009, 62, 19–27. [Google Scholar]
  324. Popay, A.J.; Prestidge, R.A.; Rowan, D.D.; Dymock, J.J. The role of Acremonium lolii mycotoxins in the insect resistance in perennial ryegrass (Lolium perenne). In Proceedings of the International Symposium on Acremonium/Grass Interactions, Baton Rouge, LA, USA, 9 October–14November 1990; Quisenberry, S.S., Joost, R.R., Eds.; pp. 44–47. [Google Scholar]
  325. Latch, G.C.M.; Christensen, M.J.; Gaynor, D.L. Aphid detection of endophyte infection in tall fescue. N. Z. J. Agric. Res. 1985, 28, 129–132. [Google Scholar] [CrossRef]
  326. Johnson, M.C.; Dahlman, D.L.; Siegel, M.R.; Bush, L.P.; Latch, G.C.M.; Potter, D.A.; Varney, D.R. Insect feeding deterrents in endophyte-infected tall fescue. Appl. Environ. Microbiol. 1985, 49, 568–571. [Google Scholar] [CrossRef] [Green Version]
  327. Riedell, W.E.; Kieckhefer, R.E.; Petroski, R.J.; Powell, R.G. Naturally occurring and synthetic loline alkaloid derivatives: Insect feeding behavior modification and toxicity. J. Entomol. Sci. 1991, 26, 122–129. [Google Scholar] [CrossRef]
  328. Kanda, K.; Hirai, H.; Koga, H.; Hasegawa, K. Endophyte enhanced resistance in perennial ryegrass [Lolium perenne] and tall fescue [Festuca arundinacea] to bluegrass webworm, Parapediasia teterrella. Jpn. Appl. Entomol. Zool. 1994, 3, 141–145. [Google Scholar] [CrossRef] [Green Version]
  329. Kanda, K.; Koga, H.; Hirai, Y.; Kasegawa, K.; Uematu, T.; Tsukiboshi, T. Resistance of Acremonium endophyte-infected perennial ryegrass and tall fescue to bluegrass webworm, Parapediasia teterella. Abstract. Proc. Phytopathol. Soc. Jpn. 1992, 58, 587. [Google Scholar]
  330. Koga, H.; Hirai, Y.; Kanda, K.I.; Tsukiboshi, T.; Uematsu, T. Successive transmission of resistance to bluegrass webworm to perennial ryegrass and tall fescue plants by artificial inoculation with Acremonium endophytes. Jpn. Agric. Res. Q. 1997, 31, 109–115. [Google Scholar]
  331. Sabzalian, M.R.; Hatami, B.; Mirlohi, A. Mealybug, Phenococcus solani, and barley aphid, Sipha maydis, response to endophyte-infected tall and meadow fescue. Entomol. Exp. Appl. 2004, 113, 205–209. [Google Scholar] [CrossRef]
  332. Patterson, C.G.; Potter, D.A.; Fannin, F.F. Feeding deterrency of alkaloids from endophyte-infected grasses to Japanese beetle grubs. Entomol. Exp. Appl. 1991, 61, 285–289. [Google Scholar] [CrossRef]
  333. Murphy, J.A.; Sun, S.; Betts, L.L. Endophyte-enhanced resistance to billbug (Coleoptera: Curculionidae), sod webworm (Lepidoptera: Pyralidae) and white grub (Coleoptera: Scarabaeidae) in tall fescue. Environ. Entomol. 1993, 22, 699–703. [Google Scholar] [CrossRef]
  334. Oliver, J.B.; Pless, C.D.; Gwinn, K.D. Effect of endophyte, Acremonium coenophialum in ‘Kentucky 31′ tall fescue, Festuca arundinacae, on survival of Popillia japonica. In Proceedings of the Internation Symposium Acremonium/Grass Interactions, Louisiana Agric. Expt Station, Baton Rouge, LA, USA, 9 October–14 November 1990; pp. 173–175. [Google Scholar]
  335. Potter, D.A.; Patterson, C.G.; Redmond, C.T. Influence of turfgrass species and tall fescue endophyte on feeding ecology of Japanese beetle and southern masked chafer grubs (Coleoptera: Scarabaeidae). J. Econ. Entomol. 1992, 8, 900–909. [Google Scholar] [CrossRef]
  336. Davidson, A.W.; Potter, D.A. Response of plant-feeding, predatory, and soil inhabiting invertebrates to Acremonium endophyte and nitrogen fertilisation in tall fescue turf. J. Econ. Entomol. 1995, 88, 376–379. [Google Scholar] [CrossRef]
  337. Richmond, D.S.; Grewal, P.S.; Cardina, J. Influence of Japanese beetle Popillia japonica larvae and fungal endophytes on competition between turfgrasses and dandelion. Crop Sci. 2004, 44, 600–606. [Google Scholar] [CrossRef]
  338. Siegel, M.R.; Latch, G.C.M.; Bush, L.P.; Fannin, F.F.; Rowan, D.D.; Tapper, B.A.; Bacon, C.W.; Johnson, M.C. Fungal endophyte-infected grasses: Alkaloid accumulation and aphid response. J. Chem. Ecol. 1990, 16, 3301–3315. [Google Scholar] [CrossRef] [PubMed]
  339. Eichenseer, H.; Dahlman, D.L.; Bush, L.P. Influence of endophyte infection, plant age and harvest interval on Rhopalosiphum padi survival and its relation to quantity of N-formyl and N-acetyl loline in tall fescue. Entomol. Exp. Appl. 1991, 60, 29–38. [Google Scholar] [CrossRef]
  340. Wilkinson, H.H.; Siegel, M.R.; Blankenship, J.D.; Mallory, A.C.; Bush, L.P.; Schardl, C.L. Contribution of fungal loline alkaloids to protection from aphids in a grass-endophyte mutualism. Mol. Plant Microbe Interact. 2000, 13, 1027–1033. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  341. Bultman, T.L.; Pulas, C.; Grant, L.; Bell, G.; Sullivan, T.J. Effects of fungal endophyte isolate on performance and preference of bird cherry oat aphid. Environ. Entomol. 2006, 35, 1690–1695. [Google Scholar] [CrossRef]
  342. Li, C.; Zhang, X.; Li, F.; Nan, Z.; Schardl, C.L. Disease and pest resistance of endophyte infected and non-infected drunken horse grass. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Research and Practice Series; Volume 13, pp. 111–114.
  343. Ahmad, S.; Johnson-Cicalese, J.M.; Dickson, W.K.; Funk, C.R. 1986. Endophyte-enhanced resistance in perennial ryegrass to the bluegrass billbug, Sphenophorus parvulus. Entomol. Exp. Appl. 1986, 41, 3–10. [Google Scholar] [CrossRef]
  344. Ahmad, S.; Funk, C.R. Bluegrass billbug tolerance of ryegrass cultivars and selections. J. Econ. Entomol. 1983, 76, 414–416. [Google Scholar] [CrossRef]
  345. Clay, K.; Hardy, T.N.; Hammond, A.M. Fungal endophytes of grasses and their effects on an insect herbivore. Oecologia 1985, 66, 1–6. [Google Scholar] [CrossRef]
  346. Hardy, T.; Clay, K.; Hammond, A.M., Jr. Fall armyworm (Lepidoptera: Noctuidae): A laboratory bioassay and larval preference study for the fungal endophyte of perennial ryegrass. J. Econ. Entomol. 1985, 78, 571–574. [Google Scholar] [CrossRef]
  347. Hardy, T.; Clay, K.; Hammond, A.M., Jr. Leaf age and reated factors affecting endophyte-mediated resistance to fall armyworm (Lepidoptera: Noctuidae) in tall fescue. Environ. Entomol. 1986, 15, 1083–1089. [Google Scholar] [CrossRef]
  348. Breen, J.P. Enhanced resistance to fall armyworm (Lepidoptera: Noctuidae) in Acremonium endophyte infected turfgrasses. J. Econ. Entomol. 1993, 86, 621–629. [Google Scholar] [CrossRef]
  349. Clay, K.; Cheplick, G.P. Effect of ergot alkaloids from fungal endophyte-infected grasses on fall armyworm (Spodoptera frugiperda). J. Chem. Ecol. 1989, 15, 169–182. [Google Scholar] [CrossRef] [PubMed]
  350. Ahmad, S.; Govindarajan, S.; Johnson-Cicalese, J.M.; Funk, C.R. Association of a fungal endophyte in perennial ryegrass with antibiosis to larvae of the southern armyworm. Entomol. Exp. Appl. 1987, 43, 287–294. [Google Scholar] [CrossRef]
  351. Shiba, T.; Sugawara, K. Fungal loline alkaloids in grass endophyte associations confer resistance to the rice leaf bug, Trigonotylus caelestialium. Entomol. Exp. Appl. 2009, 130, 55–62. [Google Scholar] [CrossRef]
  352. Jensen, J.G.; Popay, A.J. Perennial ryegrass infected with AR37endophyte reduces survival of porina larvae. N. Z. Plant Prot. 2004, 57, 323–328. [Google Scholar]
  353. Popay, A.J.; Cotching, B.; Moorhead, A.; Ferguson, C.M. AR37 reduces porina populations and plant damage in the field. Proc. N. Z. Grassl. Assoc. 2012, 74, 165–169. [Google Scholar] [CrossRef]
  354. Babu, J.V. Bioactive chemicals of importance in endophyte-infected grasses. Ph.D. Thesis, University of Waikato, Hillcrest, New Zealand, 2009. [Google Scholar]
  355. Cook, R.; Lewis, G.C. Fungal endophytes and nematodes of agricultural and amenity grasses. In Biotic Interactions in Plant-Pathogen Associations; Jeger, M.J., Spence, N.J., Eds.; CABI Publishing: Wallingford, UK, 2001; pp. 35–61. [Google Scholar]
  356. Kimmons, C.A.; Gwinn, K.D.; Bernard, E.C. Nematode reproduction on endophyte-infected and endophyte-free tall fescue. Plant Dis. 1990, 74, 757–761. [Google Scholar] [CrossRef]
  357. Kirkpatrick, T.L.; Barham, J.D.; Bateman, R.J. Host status for Meloidogyne graminis of tall fescue selections and clones with and without the endophyte Acremonium coenophialum. In Proceedings of the 1st International Symposium on Acremonium/Grass Interactions, New Orleans, LA, USA, 9 October–14 November 1990; pp. 154–156. [Google Scholar]
  358. Elmi, A.A.; West, C.P.; Robbins, R.T.; Kirkpatrick, T.L. Endophyte effects on reproduction of root-knot nematode (Meloidogyne marylandi) and osmotic adjustment in tall fescue. Grass Forage Sci. 2000, 55, 166–172. [Google Scholar] [CrossRef]
  359. Stewart, T.M.; Mercer, C.F.; Grant, J.L. Development of Meloidogyne nassi on endophyte-infected and endophyte-free perennial ryegrass. Australas. Plant Pathol. 1993, 22, 40–41. [Google Scholar] [CrossRef]
  360. Pedersen, J.F.; Rodriguez-Kabana, R.; Shelby, R.A. Ryegrass cultivars and endophyte in tall fescue affect nematodes in grass and succeeding soybean. Agron. J. 1988, 80, 811–814. [Google Scholar] [CrossRef]
  361. Bacetty, A.A.; Snook, M.E.; Glenn, A.E.; Noe, J.P.; Nagabhyru, P.; Bacon, C.W. Chemotaxis disruption in Pratylenchus scribneri by tall fescue root extracts and alkaloids. J. Chem. Ecol. 2009, 35, 844–850. [Google Scholar] [CrossRef]
  362. West, C.P.; Izekor, E.; Oosterhuis, D.M.; Robbins, R.T. The effect of Acremonium coenophialum on the growth and nematode infestation of tall fescue. Plant Soil 1988, 112, 3–6. [Google Scholar] [CrossRef]
  363. Gwinn, K.D.; Bernard, E.C. Interactions of endophyte-infected grasses with the nematodes Meloidogyne marylandii and Pratylenchus scribneri. In Proceedings of the 2nd International Symposium on Acremonium/Grass Interactions, Palmerston North, New Zealand, 2–12 February 1993; pp. 156–160. [Google Scholar]
  364. Watson, R.N.; Prestidge, R.A.; Ball, O.J.-P. Suppression of white clover by ryegrass infected with Acremonium endophyte. In Proceedings of the 2nd International Symposium on Acremonium/Grass Interactions, Palmerston North, New Zealand, 2–12 February 1993; pp. 218–221. [Google Scholar]
  365. Eerens, J.P.J.; Visker, M.H.P.W.; Lucas, R.J.; Easton, H.S.; White, J.G.H. Influence of the ryegrass endophyte (Neotyphodium lolii) in a cool moist environment: IV. Plant parasitic nematodes. N. Z. J. Agric. Res. 1998, 41, 209–217. [Google Scholar] [CrossRef] [Green Version]
  366. Barker, G.M. Mollusc herbivory influenced by endophytic claviciptaceous fungal infections in grasses. Ann. Appl. Biol. 2008, 153, 381–393. [Google Scholar] [CrossRef]
  367. Watson, B. The effect of endophyte in perennial ryegrass and tall fescue on red and black headed pasture cockchafers. In Proceedings of the 6th International Symposium on Fungal Endophytes on Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 347–352.
  368. Lopez, J.E.; Faeth, S.H.; Miller, M. Effect of endophytic fungi on herbivory by red legged grasshoppers (Orthoptera: Acrididae) on Arizona fescue. Environ. Entomol. 1995, 24, 1576–1580. [Google Scholar] [CrossRef]
  369. Watson, R.N.; Hume, D.E.; Bell, N.L.; Neville, F.J. Plant-parasitic nematodes associated with perennial ryegrass and tall fescue with and without Acremonium endophyte. N. Z. Plant Protect. 1995, 48, 199–203. [Google Scholar] [CrossRef]
  370. Wiewiora, B.; Zurek, G.; Zurek, M. Endophyte-mediated disease resistance in wild populations of perennial ryegrass (Lolium perenne). Fungal Ecol. 2015, 15, 1–8. [Google Scholar] [CrossRef]
  371. Xia, C.; Li, N.; Zhang, Y.; Li, C.; Zhang, X.; Nan, Z. Role of Epichloë endophytes in defense responses of cool-season grasses to pathogens: A review. Plant Disease 2018, 102, 2061–2073. [Google Scholar] [CrossRef] [Green Version]
  372. Christensen, M.J.; Latch, G.C.M.; Tapper, B.A. Variation within isolates of Acremonium endophytes from perennial rye-grasses. Mycol. Res. 1991, 95, 918–923. [Google Scholar] [CrossRef]
  373. Christensen, M.J. Antifungal activity in grasses infected with Acremonium and Epichloë endophytes. Australas. Plant Pathol. 1996, 25, 186–191. [Google Scholar] [CrossRef]
  374. Niones, J.T.; Takemoto, D. VibA, a homologue of a transcription factor for fungal heterokaryon incompatibility, is involved in antifungal compound production in the plant-symbiotic fungus Epichloë festucae. Eukaryot Cell 2015, 14, 13–24. [Google Scholar] [CrossRef] [Green Version]
  375. Wang, X.; Zhou, Y.; Ren, A.; Gao, Y. Effect of endophyte infection on fungal disease resistance of Leymus chinensis. Acta Ecol. Sin. 2014, 23. Available online: http://en.cnki.com.cn/Article_en/CJFDTotal-STXB201423003.htm. (accessed on 17 June 2020).
  376. Purev, E.; Kondo, T.; Takemoto, D.; Niones, J.T.; Ojika, M. Identification of ε-Poly-L-lysine as an antimicrobial product from an Epichloë endophyte and isolation of fungal ε-PL synthetase gene. Molecules 2020, 25, 1032. [Google Scholar] [CrossRef] [Green Version]
  377. Nissinen, R.; Helander, M.; Kumar, M.; Saikkonen, K. Heritable Epichloë symbiosis shapes fungal but not bacterial communities of plant leaves. Sci. Rep. 2019, 9, 5253. [Google Scholar] [CrossRef]
  378. Roberts, E.L.; Ferraro, A. Rhizosphere microbiome selection by Epichloë endophytes of Festuca arundinacea. Plant Soil 2015. [Google Scholar] [CrossRef]
  379. Zhong, R.; Xia, C.; Ju, Y.; Li, N.; Zhang, X.; Nan, Z.; Christensen, M.J. Effects of Epichloë gansuensis on root-associated fungal communities of Achnatherum inebrians under different growth conditions. Fungal Ecol. 2018, 31, 29–36. [Google Scholar] [CrossRef]
  380. Ju, Y.; Zhong, R.; Christensen, M.J.; Zhang, X. Effects of Epichloë gansuensis endophyte on the root and rhizosphere soil bacteria of Achnatherum inebrians under different moisture conditions. Front. Microbiol. 2020, 11. [Google Scholar] [CrossRef]
  381. Roberts, E.; Lindow, S. Loline alkaloid production by fungal endophytes of Fescue species select for particular epiphytic bacterial microflora. ISME J. 2014, 8, 359–368. [Google Scholar] [CrossRef] [Green Version]
  382. Mormile, B.W. Influence of seed microbiome on fitness of Epichloë infected tall fescue seedlings. Master’s Thesis, Southern Connecticut State University, New Haven, CT, USA, 2016. [Google Scholar]
  383. Tian, P.; Nan, Z.; Li, C.; Spangenberg, G. Effect of the endophyte Neotyphodium lolii on susceptibility and host physiological response of perennial ryegrass to fungal pathogens. Eur. J. Plant Pathol. 2008, 122, 593–602. [Google Scholar] [CrossRef]
  384. Nan, Z.B.; Li, C.J. Neotyphodium in native grasses in China and observations on endophyte/host interactions. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium, Soest, Germany, 27–29 September 2000; Paul, V.H., Dapprich, P.D., Eds.; pp. 41–50. [Google Scholar]
  385. Wang, X.Y.; Qin, J.H.; Chen, W.; Zhou, Y.; Ren, A.Z.; Goa, Y.B. Pathogen resistance advantage of endophyte-infected over endophyte-free Leymus chinensis is strengthen by pre-drought treatment. Eur. J. Plant Pathol. 2016, 144, 477–486. [Google Scholar] [CrossRef]
  386. Panka, D.; Jeske, M.; Troczynski, M. Occurrence of Neotyphodium and Epichloë fungi in meadow fescue and red fescue in Poland and screening of endophyte isolates as potential biological control agents. Acta Sci. Pol. Hortorum Cultus 2013, 12, 67–83. [Google Scholar]
  387. Xia, C.; Zhang, X.X.; Christensen, M.J.; Nan, Z.B.; Li, C.J. Epichloë endopyte affects the ability of powdery mildew (Blumeria graminis) to colonise drunken horse grass (Achnatherum inebrians). Fungal Ecol. 2015, 16, 26–33. [Google Scholar] [CrossRef]
  388. Perez, L.I.; Gundel, P.E.; Ghersa, C.M.; Omacini, M. Family issues: Fungal endophyte protects host grass from the closely related pathogen Claviceps purpurea. Fungal Ecol. 2013, 6, 379–386. [Google Scholar] [CrossRef]
  389. Trevathan, L.E. Performance of endophyte-free and endophyte-infected tall fescue seedlings in soil infested with Cochliobolus sativus. Can. J. Plant Pathol. 1996, 18, 415–418. [Google Scholar] [CrossRef]
  390. Chen, W.; Liu, H.; Wurihan; Gao, Y.; Card, S.D.; Ren, A. The advantages of endophyte infected over uninfected tall fescue in the growth and pathogen resistance are counteracted by elevated CO2. Sci. Rep. 2017, 7, 6952. [Google Scholar] [CrossRef]
  391. Reddy, M.N.; Faeth, S.H. Damping-off of Festuca arizonica caused by Fusarium. Am. J. Plant Sci. 2010, 1, 104–105. [Google Scholar] [CrossRef] [Green Version]
  392. Hume, D.E.; Quigley, P.E.; Aldaoud, R. Influence of Neotyphodium infection on plant survival of disease tall fescue and ryegrass. In Neotyphodium/Grass Interactions; Bacon, C.W., Hill, N.S., Eds.; Springer Science + Business Media: New York, NY, USA, 1997; pp. 171–172. [Google Scholar]
  393. Bonos, S.A.; Wilson, M.M.; Meyer, W.A.; Funk, R.C. Suppression of red thread in fine fescues through endophyte-mediated resistance. Appl. Turfgrass Sci. 2005, 2, 1–7. [Google Scholar] [CrossRef]
  394. Zabalgogeazcoa, I.; Gundel, P.E.; Helander, M.; Saikkonen, K. Non-systemic fungal endophytes in Festuca rubra plants infected by Epichloë festucae in subarctic habitats. Fungal Divers. 2013, 60, 25–32. [Google Scholar] [CrossRef]
  395. Welty, R.E.; Barker, R.E.; Azevedo, M.D. Reaction of tall fescue infected and noninfected by Acremonium coenophialum to Puccinia graminis subsp. Graminicola Plant Dis. 1991, 75, 883–886. [Google Scholar] [CrossRef]
  396. Panka, D.; Jeske, M.; Troczynski, M. Effect of Neotyphodium uncinatum endophyte on meadow fescue yielding, health status and ergovaline production in host-plants. J. Plant Protect. 2011, 51, 362–370. [Google Scholar]
  397. Wheatley, W.M.; Nicol, H.I.; Hunt, E.R.; Nikandrow, A.; Cother, N. An association between perennial ryegrass endophyte, a leafspot caused by Pyrenophora semeniperda and preferential grazing by sheep. In Proceedings of the 3rd International Conference on Harmful and Beneficial Microorganisms in Grassland, Pasture and Turf, Paderbom, Germany, 26–29 September 2000; pp. 71–75. [Google Scholar]
  398. Burpee, L.L.; Bouton, J.H. Effect of eradication of the endophyte Acremonium coenophialum on epidemics of Rhizoctonia blight in tall fescue. Plant Dis. 1993, 77, 157–159. [Google Scholar] [CrossRef]
  399. Gwinn, K.D.; Gavin, A.M. Relationship between endophyte infestation level of tall fescue seed lots and Rhizoctonia zeae seedling disease. Plant Dis. 1992, 76, 911–914. [Google Scholar] [CrossRef]
  400. Clarke, B.B.; White, J.F.; Hurley, R.H.; Torres, M.S.; Suns, S.; Huff, D.R. Endophyte-Mediated Suppression of Dollar Spot Disease in Fine Fescues. Plant Dis. 2006, 90, 994–998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  401. Tian, Z.; Wang, R.; Ambrose, K.; Clarke, B.B.; Belanger, F.C. The Epichloë festucae antifungal protein has activity against the plant pathogen Sclerotinia homoeocarpa, the causal agent of dollar spot disease. Sci. Rep. 2017, 7, 5643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  402. Wäli, P.R.; Helander, M.; Nissinen, O.; Saikkonen, K. Susceptibility of endophyte-infected grasses to winter pathogens (snow molds). Can. J. Bot. 2006, 84, 1043–1051. [Google Scholar] [CrossRef]
  403. Vignale, M.V.; Astiz-Gassó, M.M.; Novas, M.V.; Iannone, L.J. Epichloid endophytes confer resistance to the smut Ustilago bullata in the wild grass Bromus auleticus (Trin.). Biol. Control. 2013, 67, 1–7. [Google Scholar] [CrossRef]
  404. Siegel, M.R.; Latch, G.C.M.; Johnson, M.C. Fungal endophytes of grasses. Annu. Rev. Phytopathol. 1987, 25, 293–315. [Google Scholar] [CrossRef]
  405. Latch, G.C.M.; Hunt, W.F.; Musgrave, D.R. Endophytic fungi affect growth of perennial ryegrass. N. Z. J. Agric. Res. 1985, 28, 165–168. [Google Scholar] [CrossRef]
  406. Belesky, D.P.; Fedders, J.M. Does endophyte influence regrowth of tall fescue? Ann. Bot. 1996, 78, 499–505. [Google Scholar] [CrossRef] [Green Version]
  407. Monnet, F.; Vaillant, N.; Hitmi, A.; Sallanon, H. Photosynthetic activity of Lolium perenne as a function of endophyte status and zinc nutrition. Funct. Plant Biol. 2005, 32, 131–139. [Google Scholar] [CrossRef]
  408. Easton, H.S.; Fletcher, L.R. The importance of endophyte in agricultural systems—Changing plant and animal productivity. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 11–18.
  409. Young, C.A.; Hume, D.E.; McCulley, R.L. Forages and Pastures Symposium: Fungal endophytes of tall fescue and perennial ryegrass: Pasture friend or foe? J. Animal Sci. 2013, 91, 2379–2394. [Google Scholar] [CrossRef] [Green Version]
  410. Caradus, J.R. The commercial impact of Neotyphodium endophyte science and technology. In Proceedings of the 8th International Grass Endophyte Symposium, Lanzhou, China, 13–16 August 2012; pp. 203–206. [Google Scholar]
  411. Fletcher, L.R. Novel Endophytes in New Zealand Grazing Systems: The Perfect Solution or a Compromise? In Epichloaëe, Endophytes of Cool Season Grasses: Implications, Utilization and Biology; Young, C.A., Aiken, G.E., McCulley, R.L., Strickland, J.R., Schardl, C.L., Eds.; Samuel Roberts Noble Foundation: Ardmore, OK, USA, 2012; pp. 5–13. [Google Scholar]
  412. Thom, E.R.; Popay, A.J.; Hume, D.E.; Fletcher, L.R. Evaluating the performance of endophytes in farm systems to improve farmer outcomes—A review. Crop Pasture Sci. 2012, 63, 927–943. [Google Scholar] [CrossRef]
  413. Hume, D.E.; Cosgrove, G.P. Endophyte-what is it and its significance in New Zealand pastoral agriculture. In Proceedings of the 20th Annual Conference of the Grassland Society of NSW, Orange, Australia; 19–21 July 2005; pp. 31–36. [Google Scholar]
  414. Fletcher, L.R. “Non-toxic” endophytes in ryegrass and their effect on livestock health and production. In Ryegrass Endophyte—An Essential New Zealand Symbiosis; Woodfield, D.R., Matthew, C., Eds.; NZ Grassland Association Grassland Research and Practice Series; NZ Grassland Association: Napier, New Zealand, 1999; Volume 7, pp. 133–139. [Google Scholar]
  415. Tapper, B.A.; Latch, G.C.M. Selection against toxin production in endophyte infected perennial ryegrass. In Ryegrass Endophyte—An Essential New Zealand Symbiosis; Woodfield, D.R., Matthew, C., Eds.; NZ Grassland Association Grassland Research and Practice Series; NZ Grassland Association: Napier, New Zealand, 1999; Volume 7, pp. 107–111. [Google Scholar]
  416. Hume, D.E.; Ryan, D.L.; Cooper, B.M.; Popay, A.J. Agronomic performance of AR37-infected ryegrass in northern New Zealand. Proc. N. Z. Grassl. Assoc. 2007, 69, 201–205. [Google Scholar] [CrossRef]
  417. Fletcher, L.R. Managing ryegrass-endophyte toxicosis. In Neotyphodium in Cool-Season Grasses; Roberts, C.A., West, C.P., Spiers, D.E., Eds.; Blackwell Publishing: Ames, IA, USA, 2005; pp. 229–241. [Google Scholar]
  418. Easton, H.S.; Tapper, B.A. Neotyphodium down under—Research developments in New Zealand. In Proceedings of the 5th International Symposium, Neotyphodium/Grass Interactions, Fayetteville, AR, USA, 23–26 May 2004. [Google Scholar]
  419. Bluett, S.J.; Thom, E.R.; Clark, D.A.; Macdonald, K.A.; Minneé, E.M.K. Effects of perennial ryegrass infected with either AR1 or standard (wild) endophyte on dairy production in the Waikato. N. Z. J. Agric. Res. 2005, 48, 197–212. [Google Scholar] [CrossRef]
  420. Bluett, S.J.; Thom, E.R.; Clark, D.A.; Waugh, C.D. Effects of a novel ryegrass endophyte on pasture production, dairy cow milk production and calf liveweight gain. Aust. J. Exp. Agric. 2005, 45, 11–19. [Google Scholar] [CrossRef]
  421. Ussher, G. Northlands Pasture Toxin Project. N. Z. Large Herds Assoc. Annu. Conf. 2003, 34, 62–64. [Google Scholar]
  422. Milne, G.D. Technology transfer of novel ryegrass endophytes in New Zealand. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 237–239.
  423. Easton, H.S.; Cooper, B.M.; Lyons, T.B.; Pennell, C.G.L.; Popay, A.J.; Tapper, B.A.; Simpson, W.R. Selected endophyte and plant variation. In Proceedings of the 4th International Neotyphodium/Grass Interactions Symposium, Soest, Germany, 27–29 September 2000; Paul, V.H., Dapprich, P.D., Eds.; pp. 351–356. [Google Scholar]
  424. Ferguson, C.M.; Barratt, B.I.P.; Bell, N.; Goldson, S.L.; Hardwick, S.; Jackson, M.; Jackson, T.A.; Phillips, C.B.; Popay, A.J.; Rennie, G.; et al. Quantifying the economic cost of invertebrate pests to New Zealand’s pastoral industry. N. Z. J. Agric. Res. 2019, 62, 255–315. [Google Scholar] [CrossRef] [Green Version]
  425. Tapper, B.A.; Lane, G.A. Janthitrems found in a Neotyphodium endophyte of perennial ryegrass. In Proceedings of the 5th International Symposium Neotyphodium/Grass Interactions, Fayetteville, AR, USA, May 23–26 2004; p. 301. [Google Scholar]
  426. Finch, S.C.; Fletcher, L.R.; Babu, J.V. The evaluation of endophyte toxin residues in sheep fat. N. Z. Vet. J. 2012, 60, 56–60. [Google Scholar] [CrossRef]
  427. Thom, E.R.; Waugh, C.D.; Minneé, E.M.K.; Waghorn, G.C. A new generation ryegrass endophyte—The first results from dairy cows fed AR37. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practrice Series; Volume 13, pp. 293–296.
  428. Popay, A.J.; Wyatt, R.T. Resistance to Argentine stem weevil in perennial ryegrass infected with endophytes producing different alkaloids. In Proceedings of the 48th NZ Plant Protection Conference, Wellington, New Zealand, 21–23 July 1995; pp. 229–236. [Google Scholar]
  429. Popay, A.J.; Thom, E.R. Endophyte effects on major insect pests in Waikato dairy pasture. Pasture persistence symposium. N. Z. Grassl. Res. Pract. Ser. 2009, 15, 121–126. [Google Scholar]
  430. Stewart, A.; Kerr, G.; Lissaman, W.; Rowarth, J. Endophyte in Ryegrass and Tall Fescue. In Pasture and Forage Plants for New Zealand; Davies, K., Casey, M., Eds.; NZ Grassland Assoaciation Grassland and Research Practice Series; NZ Grassland Assoaciation Grassland: Napier, New Zealand, 2014; Volume 8, pp. 66–77. [Google Scholar]
  431. Popay, A.J.; Rijswijk, K.; Goldson, S.L. Argentine stem weevil: Farmer awareness and the effectiveness of different ryegrass/endophyte associations. J. N. Z. Grassl. 2017, 79, 147–152. [Google Scholar]
  432. Hume, D.E.; Popay, A.J.; Cooper, B.M.; Eerens, J.P.J.; Lyons, T.B.; Pennell, G.C.L.; Tapper, B.A.; Latch, G.C.M.; Baird, D.B. Effect of a novel endophyte on the productivity of perennial ryegrass (Lolium perenne) in New Zealand. In Proceedings of the 5th International Symposium on Neotyphodium/Grass Interactions, Fayetteville, AR, USA, 23–26 May 2004; p. 313. [Google Scholar]
  433. Thom, E.R.; Waugh, C.D.; Minneé, E.M.; Waghorn, G.C. Effects of novel and wild-type endophytes in perennial ryegrass on cow health and production. N. Z. Vet. J. 2013, 61, 87–97. [Google Scholar] [CrossRef]
  434. Fletcher, L.R.; Finch, S.C.; Sutherland, B.L.; de Nicolo, G.; Mace, W.J.; van Kote, C.; Hume, D.E. The occurrence of ryegrass staggers and heat stress in sheep grazing ryegrass endophyte associations with diverse alkaloid profiles. N. Z. Vet. J. 2017, 65, 232–241. [Google Scholar] [CrossRef] [Green Version]
  435. van Zijll, E.; Dobrowolski, M.P.; Sandford, A.; Smith, K.F.; Willocks, M.J.; Spangenberg, G.C.; Forster, J.W. Detection and characterisation of novel fungal endophyte genotypic variation in cultivars of perennial ryegrass (Lolium perenne L.). Aust. J. Agric. Res. 2008, 59, 214–221. [Google Scholar] [CrossRef]
  436. Eady, C.C.; Corkran, J.R.; Bailey, K.M.; Kerr, G.A.; Nicol, A.M. Estimation of ergovaline intake of cows from grazed perennial ryegrass containing NEA2 or standard endophyte. J. N. Z. Grassl. 2007, 79, 189–196. [Google Scholar]
  437. Ruppert, K.G.; Matthew, C.; McKenzie, C.M.; Popay, A.J. Impact of Epichloë endophytes on adult Argentine stem weevil damage to perennial ryegrass seedlings. Entomol. Exp. Appl. 2017, 163, 328–337. [Google Scholar] [CrossRef]
  438. Popay, A.J.; McNeill, M.R.; Goldson, S.L.; Ferguson, C.M. The current status of Argentine stem weevil (Listronotus bonariensis) as a pest in the North Island of New Zealand. N. Z. Plant Protect. 2011, 64, 55–62. [Google Scholar] [CrossRef] [Green Version]
  439. Caradus, J.R.; Card, S.D.; Finch, S.C.; Hume, D.E.; Johnson, L.J.; Mace, W.J.; Popay, A.J. Ergot alkaloids in New Zealand pastures and their impact. N. Z. J. Agric. Res. 2020. [Google Scholar] [CrossRef]
  440. Logan, C.M.; Edwards, G.R.; Kerr, G.A.; Williams, S. Ryegrass staggers and liveweight gain of ewe lambs and hoggets grazing four combinations of perennial ryegrass and strains of endophyte. Proc. N. Z. Soc. Anim. Prod. 2015, 75, 175–178. [Google Scholar]
  441. Cameron, N.E. Grass Endophyte. U.S. Patent 9,133,434 B1, 26 May 2015. [Google Scholar]
  442. Clayton, W. Molecular and cellular analysis of the endophyte Neotyphodium uncinatum and its association with Festulolium. Master’s Thesis, Massey University, Palmerston North, New Zealand, 2013; p. 143. [Google Scholar]
  443. Nboyine, J.A.; Saville, D.; Boyer, S.; Cruickshank, R.H.; Wratten, S.D. When host-plant resistance to a pest leads to higher plant damage. J. Pest Sci. 2017, 90, 173–182. [Google Scholar] [CrossRef]
  444. Barker, G.M.; Patchett, B.J.; Gillanders, T.J.; Brown, G.S.; Montel, S.J.Y.; Cameron, N.E. Feeding and oviposition by Argentine stem weevil on Epichloë uncinata-infected, loline-containing Festulolium. Proc. N. Z. Plant Protect. 2015, 68, 212–217. [Google Scholar]
  445. Thompson, F.N.; Stuedemann, J.A. Pathophysiology of fescue toxicosis. Agric. Ecosyst. Environ. 1993, 44, 263–281. [Google Scholar] [CrossRef]
  446. Strickland, J.R.; Oliver, J.W.; Cross, D.L. Fescue toxicosis and its impact on animal agriculture. Vet. Hum. Toxicol. 1993, 35, 454–464. [Google Scholar]
  447. Bouton, J.H.; Hill, N.S.; Hoveland, C.S.; McCann, M.A.; Thompson, F.N.; Hawkins, L.L.; Latch, G.C.M. Performance of tall fescue cultivars infected with nontoxic endophytes. In Proceedings of the 4th International Neotyphodium/ Grass Interactions Symposium, Soest, Germany, 27–29 September 2000; Paul, V.H., Dapprich, P.D., Eds.; pp. 179–185. [Google Scholar]
  448. Parish, J.A.; McCann, M.A.; Watson, R.H.; Hoveland, C.S.; Hawkins, L.L.; Hill, N.S.; Bouton, J.H. Use of non-ergot alkaloid-producing endophytes for alleviating tall fescue toxicosis in sheep. J. Anim. Sci. 2003, 81, 1316–1322. [Google Scholar] [CrossRef] [PubMed]
  449. Parish, J.A.; McCann, M.A.; Watson, R.H.; Paiva, N.N.; Hoveland, C.S.; Parks, A.H.; Upchurch, B.L.; Hill, N.S.; Bouton, J.H. Use of non-ergot alkaloid-producing endophytes for alleviating tall fescue toxicosis in stocker cattle. J. Anim. Sci. 2003, 81, 2856–2868. [Google Scholar] [CrossRef] [PubMed]
  450. Watson, R.H.; McCann, M.A.; Parish, J.A.; Hoveland, C.S.; Thompson, F.N.; Bouton, J.H. Productivity of cow-calf pairs grazing tall fescue pastures infected with either the wild-type endophyte or a nonergot alkaloid-producing endophyte strain, AR542. J. Anim. Sci. 2004, 82, 3388–3393. [Google Scholar] [CrossRef] [PubMed]
  451. Ball, D.M.; Lacefield, G.D.; Agee, C.S.; Hoveland, C.S. Introduction and acceptance of novel endophyte tall fescue in the USA. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 249–251.
  452. Hill, N.; Roach, P. Endophyte survival during seed storage: Endophyte–host interactions and heritability. Crop Sci. 2009, 49, 1425–1430. [Google Scholar] [CrossRef]
  453. Popay, A.J.; Jensen, J.G.; Cooper, B.M. The effect of non-toxic endophytes in tall fescue on two major insect pests. Proc. N. Z. Grassl. Assoc. 2005, 67, 169–173. [Google Scholar] [CrossRef]
  454. Ball, O.J.-P.; Coudron, T.A.; Tapper, B.A.; Davies, E.; Trently, D.; Bush, L.P.; Gwinn, K.D.; Popay, A.J. Importance of host plant species, Neotyphodium endophyte isolate, and alkaloids on feeding by Spodoptera frugiperda (Lepidoptera: Noctuidae) larvae. J. Econ. Entomol. 2006, 99, 1462–1473. [Google Scholar] [CrossRef] [PubMed]
  455. Ball, O.J.-P.; Gwinn, K.D.; Pless, C.D.; Popay, A.J. Endophyte isolate and host grass effects on Chaetocnema pulicaria (Coleoptera: Chrysomelidae) feeding. J. Econ. Entomol. 2011, 104, 665–672. [Google Scholar] [CrossRef]
  456. Hunt, M.G.; Newman, J.A. Reduced herbivore resistance from a novel grass-endophyte association. J. Appl. Ecol. 2005, 42, 762–769. [Google Scholar] [CrossRef]
  457. Tozer, K.N.; Ates, S.; Mapp, N.R.; Smith, M.C.; Lucas, R.J.; Edwards, G.R. Effects of MaxPTM endophyte in tall fescue on pasture production and composition, and sheep grazing preference, in a dryland environment. In Proceedings of the 6th International Symposium Fungal Endophytes Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; N.Z Grassland Association Grassland Research and Practice Series; Volume 13, pp. 259–272.
  458. Hopkins, A.A.; Young, C.A.; Simpson, W.R.; Panaccione, D.G.; Mittal, S.; Bouton, J.H. Agronomic performance and lamb safety of tall fescue novel endophyte combinations in the south-central USA. Crop Sci. 2010, 50, 1552–1561. [Google Scholar] [CrossRef]
  459. Shymanovich, T.; Crowley, G.; Ingram, S.; Steen, C.; Panaccione, D.G.; Young, C.A.; Watson, W.; Poore, M. Endophytes matter: Variation of dung beetle performance across different endophyte-infected tall fescue cultivars. Appl. Soil Ecol. 2020, 152. [Google Scholar] [CrossRef]
  460. Van Hanja, N.; de Bruin, J. Tall Fescue Endophyte E34. U.S. Patent 7,642,424 B2, 5 January 2010. [Google Scholar]
  461. Dillard, S.L.; Smith, S.R.; Hancock, D.W. variability of ergovaline and total ergot alkaloid expression among endophytic tall fescue cultivars. Crop Sci. 2019, 59, 2866–2875. [Google Scholar] [CrossRef]
  462. Craig, A.M.; Blythe, L.L.; Duringer, J.M. The role of the Oregon State University Endophyte service laboratory in diagnosing clinical cases of endophyte toxicoses. J. Agric. Food Chem. 2014, 62, 7376–7381. [Google Scholar] [CrossRef] [PubMed]
  463. Beck, P.A.; Stewart, C.B.; Gunter, S.A.; Singh, D. Evaluation of tall fescues for stocker cattle in the Gulf Coastal Plain. Prof. Anim. Sci. 2009, 25, 569–579. [Google Scholar] [CrossRef]
  464. Roulund, N.; Jensen, A.M.D. Tall Fescue Endophyte Isolate 647. U.S. Patent 9,706,779 B2, 23 August 2013. [Google Scholar]
  465. Nihsen, M.E.; Piper, E.L.; West, C.P.; Crawford, R.J., Jr.; Denard, T.M.; Johnson, Z.B.; Roberts, C.A.; Spiers, D.A.; Rosenkrans, C.F., Jr. Growth rate and physiology of steers grazing tall fescue inoculated with novel endophytes. J. Anim. Sci. 2004, 82, 878–883. [Google Scholar] [CrossRef]
  466. Rolston, M.P.; Agee, C. Delivering Quality Seed to Specification—The USA and NZ Novel Endophyte Experience. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassand Research and Practice Series; Volume 13, pp. 229–231.
  467. Rolston, M.P.; Archie, W.J.; Simpson, W. Tolerance of AR1 Neotyphodium endophyte to fungicides used in perennial ryegrass seed production. Proc. N. Z. Plant Protect. 2002, 55, 322–325. [Google Scholar] [CrossRef] [Green Version]
  468. Hume, D.E.; Barker, D.J. Growth and Management of Endophytic Grasses in Pastoral Agriculture. In Neotyphodium in Cool-Season Grasses; Roberts, C.A., West, C.P., Speirs, D.E., Eds.; Wiley-Blackwell: Hoboken, NJ, USA, 2005; pp. 201–226. [Google Scholar]
  469. Easton, S.; Tapper, B. Neotyphodium research and application in New Zealand. In Neotyphodium in Cool-Season Grasses; Roberts, C.A., West, C.P., Speirs, D.E., Eds.; Wiley-Blackwell: Hoboken, NJ, USA, 2008; pp. 35–42. [Google Scholar]
  470. Andrea, J.G.; Roberts, C.A. Transferring endophytes technology to North American farmers. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 233–236.
  471. Parish, J.A.; Watson, R.H. On-farm impacts of endophyte technology in the United States. In Proceedings of the 6th International Symposium on Fungal Endophytes of Grasses; Christchurch, New Zealand, 25–28 March 2007, Popay, A.J., Thom, E.R., Eds.; NZ Grassland Association Grassland Research and Practice Series; Volume 13, pp. 243–248.
  472. Murray, F.R.; Latch, G.C.M.; Scott, D.B. Surrogate transformation of perennial ryegrass Lolium perenne, using genetically modified Acremonium endophyte. Mol. Gen. Genet. 1992, 233, 1–9. [Google Scholar] [CrossRef]
  473. Schardl, C.L. Molecular and genetic methodologies and transformation of grass endophytes. In Biotechnology of Endophytic Fungi of Grasses; Bacon, C.W., White, J., Jr., Eds.; CRC Press: Boca Raton, FL, USA, 1994; pp. 151–165. [Google Scholar]
  474. Shi, T.-Q.; Liu, G.-N.; Ji, R.-Y.; Shi, K.; Song, P.; Ren, L.-J.; Huang, H.; Ji, X.-J. CRISPR/Cas9-based genome editing of the filamentous fungi: The state of the art. Appl. Microbiol. Biotechnol. 2017, 101, 7435–7443. [Google Scholar] [CrossRef]
  475. Johnson, L.J.; Voisey, C.R.; Faville, M.J.; Moon, C.D.; Simpson, W.R.; Johnson, R.D.; Stewart, A.V.; Caradus, J.R.; Hume, D.E. Advances and perspectives in breeding for improved grass-endophyte associations. In Proceedings of the Improving Sown Grasslands through Breeding and Management, Joint Symposium EFG/Eucarpia, Zurich, Switzerland, 24–27 June 2019; Grassland Science in Europe: Eucarpia, Italy, 2019; Volume 24, pp. 351–363. [Google Scholar]
  476. Bastias, D.A.; Johnson, L.J.; Card, S.D. Symbiotic bacteria of plant-associated fungi: Friends or foes? Curr. Opinion Plant Biol. 2020, 56, 1–8. [Google Scholar] [CrossRef]
  477. Easton, H.S.; Lyons, T.B.; Cooper, B.M.; Mace, W.J. Loline alkaloids for better protection of pastures from insect pests. Proc. N. Z. Grassl. Assoc. 2009, 71, 151–154. [Google Scholar] [CrossRef]
  478. Bassett, S.A.; Johnson, R.D.; Simpson, W.R.; Laugraud, A.; Jordan, T.W.; Bryan, G.T. Identification of a gene involved in the regulation of hyphal growth of Epichloë festucae during symbiosis. FEMS Microbiol. Lett. 2016, 363. [Google Scholar] [CrossRef]
  479. Spiering, M.J.; Moon, C.D.; Wilkinson, H.H.; Schardl, C.L. Gene clusters for insecticidal loline alkaloids in the grass-endophytic fungus Neotyphodium uncinatum. Genetics 2005, 169, 1403–1414. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  480. Pennell, C.G.L. Pesticidal Plant Extract Containing Loline Derivatives. U.S. Patent 9,375,011 B2, 12 March 2008. [Google Scholar]
  481. Yue, Q.; Miller, C.J.; White, J.F., Jr.; Richardson, M.D. Isolation and characterization of fungal inhibitors from Epichloë festucae. J. Agric. Food Chem. 2000, 48, 4687–4692. [Google Scholar] [CrossRef] [PubMed]
  482. Fernando, K.; Reddy, P.; Hettiarachchige, I.K.; Spangenberg, G.C.; Rochfort, S.J.; Guthridge, K.M. Novel antifungal activity of Lolium-associated Epichloë endophytes. Microorganisms 2020, 8, 955. [Google Scholar] [CrossRef] [PubMed]
Table 4. Invertebrate organisms (insects, nematodes and molluscs) impacted by Epichloë endophytes; for other older references related to effects of Epichloë endophytes in ryegrass and tall fescue on insects, refer to Breen (1994) [252].
Table 4. Invertebrate organisms (insects, nematodes and molluscs) impacted by Epichloë endophytes; for other older references related to effects of Epichloë endophytes in ryegrass and tall fescue on insects, refer to Breen (1994) [252].
OrganismImpactAlkaloid InvolvedEpichloë Strain/TypeReference
Insects
Acheta domesticus—house cricketsToxic to nymphsns *Ryegrass types[253]
Adoryphorus coulonii—Red-headed cockchaferReduced (10–20%) root consumption at >1000 µg/g DMLolineMeadow fescue types[254]
Agallica constricta—leaf hopperResistancensFescue types[255]
Agrostis ipsilon—Black cutwormDeterrence and toxicityErgovaline and/or ergine most potent, with lolines also effectiveE. lolii x E. typhina hybrid from ryegrass[256,257]
Aploneura lentisci—root aphidReduced survival; possible neurotoxinUnknown (in case of AR5), and possibly epoxy janthitremsAR37, AR5, AR6, and standard ryegrass endophyte[258,259,260,261,262,263,264]
Reduced root aphid numbers per plantPossibly lolines—NFL and NALFescue types[265,266,267]
Minimal effectDespite having similar ergovaline levels in roots as AR5NEA2 and NEA6 endophytes[264]
Increased numbersnsAR1 endophyte[268]
Balanococcus poae—Pasture mealybugReduced survivalnsRyegrass types including AR1[258,269,270,271]
Reduced infestationnsFescue types that do not express ergovaline[272]
Blissus leucopterus hirtus—hairy chinch bugDeterrence and toxicity to larvae and adultsnsFescue and ryegrass types[273,274,275,276,277]
No effect Fescue types[278]
Costelytra zealandica or C. giveni—Grass grubReduced root feeding and larval weight gain; a deterrent effect Loline and increased levels due to grass grub attackFescue and meadow fescue types; E. uncinatum[91,92,279,280,281,282,283,284,285,286]
Cerodontha australis—wheat sheath minerToxicity or deterrence to larvae, but no effect on ovipositionnsAR47 and AR48 ryegrass strains[287]
Crambus roman—sod webwormDeterrentnsRyegrass types (turf)[288]
Ctenocephalides felis—cat flea larvaeContact toxicityNFLFescue types[289]
Cyclocephala lurida—southern masked chaferReduced numbersnsFescue types[217]
Diuraphis noxia—Russian wheat aphidToxic to nymphs and adults; deterrent to adultsnsRyegrass and fescue types[290,291]
Draeculacephala spp.—leaf hopperResistancensFescue types[25,292]
Drosophila melanogaster—fruit flyToxic to adultsnsFescue types[293]
Exitianus exitiosus—leaf hopperResistancensFescue types[255]
Exomala orientalisReduced survivalnsFescue types[294]
Graminella nigrifrons—leaf hopperResistancensFescue types[255]
Graphania mutans—cutwormNot a deterrent, but disrupted developmentPeramineRyegrass types[295]
Heteronychus arator—African black beetle Antifeeding effect on adultsErgopeptine alkaloids - ergotamine, ergovaline, ergocryptineStandard ryegrass endophyte; AR22, AR12 endophytes[260,270,280,296,297,298,299,300]
Reduced numbersnsAR37 endophyte[260]
Deterrent, antifeeding effect on larval and adult stagesLolineFescue and meadow fescue types; E. uncinata[254,301,302]
No effectPeramine, lolitrem B, paxilline, festuclavine, lysergol, and lysergic acid amideRyegrass and fescue types[280,297,298]
Lepidogryllus spp.—mottled field cricketDeterrentLolineMeadow fescue types; E. uncinatum[303]
Listronotis bonariensis—Argentine stem weevilFeeding deterrent for both adults and larvae; reduced ovipositionPeramine—higher concentration required to control larvaeRyegrass types; AR1, AR5, NEA2 endophytes[84,245,260,270,299,304,305,306,307,308,309,310,311,312,313,314,315,316,317]
Feeding deterrent and toxin of larvae, but not adultsLolitrem BRyegrass types[315,318,319,320]
Feeding deterrentPaxillineRyegrass types[84]
Reduce larval damage of tillersnsAR37 endophyte[260]
Feeding deterrent and death of larvaeLoline level above 400 µg/g DM; NANL possibly more potent than NFL at moderate concentrationsMeadow fescue types[279,321,322,323]
Feeding deterrentErgovaline; ergocryptine; ergotamineRyegrass types[295,324]
No effect Ryegrass and fescue types[325]
Oncopeltus fasciatus—large milkweed bug Feeding deterrent and toxicNFLFescue types[140,326]
Ostrinia nubilalis—European corn borer larvaeToxic effects and reduced larval weight gainNALFescue types[327]
Parapediasa teterella—bluegrass webwormDeterrent, reduced feedingnsFescue and ryegrass types[328,329,330]
Periplaneta Americana—American cockroachContact toxicityNFLFescue types[289]
Phenococcus solani—mealybugReduced numbersnsFescue types[331]
Philobota spp.—Pasture tunnel mothsReduced numbersnsAR37[262]
Popillia japonica—Japanese beetle larvaeContact toxicityNFLFescue types[289]
Reduced feedingParticularly NFL and NAL; and lesser effect of ergotamine, ergonovine, ergocryptineFescue types[294,332]
Inconsistent effects Fescue types[294,333]
No effect Fescue and ryegrass types[334,335,336,337]
Prosapia bicincta—leaf hopperResistancensFescue types[255]
Pseudococcidae—mealybugsReduced numbersnsAR37[262]
Rhopalosiphum padi—aphidFeeding deterrent and toxicLolineFescue types[325,326,338,339,340,341]
Reduced numbersnsE. gansuense[342]
No effectErgovalineRyegrass and fescue types[338]
Rhopalosiphum maidis—Corn leaf aphidSome resistance, but less than for R. padi and S. graminumnsRyegrass types; lesser impact of fescue types[326]
Schizaphis graminum—aphidToxic causing reduced numbersLolineFescue types; E festucae and E. uncinatum[326,327,340]
Feeding deterrent and toxicPeramineRyegrass and fescue types[338]
No effectErgovaline
ResistancensFescue types[140]
Sphenphorus parvulus—Bluegrass billbug Resistance/ toxicity to adultsnsRyegrass and fescue types (turf)[288,292,343,344]
Spodoptero frugiperda—fall army wormReduced worm survival and liveweight gainsnsFescue and ryegrass types[345,346,347,348]
NFL, NAL Fescue types[327]
Ergotamine, ergonovine, ergocryptineFescue types[349]
Spodoptera eridania—southern army wormToxicnsRyegrass types[350]
Teleogryllus commodus—black field cricketDeterrentLolineMeadow fescue types; E. uncinatum[303]
Trigonotylus caelestialium—rice leaf bugResistanceLolineFescue types[351]
Wiseana cervinata—PorinaReduced survivalnsAR37 ryegrass type[80,192,256,352,353]
Reduce feeding and weight gainPaxilline [354]
LolineFescue types[282]
Mites
Tetranychus cinnabarinusReduced numbersnsE. gansuense[342]
Nematodes (refer to [355] Cook and Lewis 2001)
Helicotylenchus pseudorobustus—spiral nematodesReduced numbersnsFescue types[356]
Meloidogyne marylandiFewer egg masses and eggs and reduced infectionnsFescue types[356,357,358]
Reduced infectionns, but not ergovalineRyegrass types[90]
Meloidogyne nassiReduced galls and femalesnsRyegrass types[359]
Paratrichodorus minor—stubby root nematodesReduced numbersnsFescue types[360]
Pratylenchus scribneri—Lesion nematode Repellent and deathNFL at high concentrations; and ergovalineFescue types[356,361]
Reduced numbers nsFescue types[362,363]
Attractant and causes deathErgovaline, ergotamineFescue types[361]
RepellentErgocryptine, ergonovineFescue types
Attractant at <20 µg/m and repellent at high concentrations NFLFescue types
Pratylenchus spp.Reduced numbers in soilnsRyegrass types[364,365]
Tylenchorhynchus acutus—stunt nematodesReduced numbers in soilnsFescue types[362]
Molluscs
Deroceras reticulatumReduced feedingLolitrem B and possibly lolines Used artificial diets incorporating the secondary metabolites[366]
No effectPeramine
Stimulated feedingErgotamine and ergovaline
AttractantPaxilline, lolitriol, a-paxitriol and b-paxitriol
* ns = not specified.
Table 5. Pathogens impacted by Epichloë endophytes in planta.
Table 5. Pathogens impacted by Epichloë endophytes in planta.
PathogenImpact of EndophyteAlkaloid InvolvedEpichloë Strain/TypeReference
Alternaria alternataModerate resistanceEnhanced superoxide dismutase or peroxidases activityRyegrass types[383]
Reduced incidence of infectionns *Host: Elymus cylindricus [384]
Bipolaris sorokinianaNo effect in planta E. bromicola[375]
No effect in planta E. gansuensis[342]
Reduced incidence of infectionnsHost: Leymus chinensis[385]
nsFescue types[386]
Resistance to infectionEnhanced superoxide dismutase or peroxidases activityRyegrass types[383]
Blumeria graminis—powdery mildewLower disease incidencensE. gansuensis[342,387]
Cladosporium sp.No effect in planta E. bromicola[375]
Claviceps purpureaReduced infection unless plants water stressednsAnnual ryegrass types[388]
Cochliobolus sativus—soil pathogenNo effect Fescue types[389]
Curvularia lunataNo effect in planta E. bromicola[375]
Moderate resistanceEnhanced superoxide dismutase or peroxidases activityRyegrass types[383]
Reduced incidence of infectionnsHost: Leymus chinensis[385]
Reduced disease symptomsnsFescue types[390]
Drechsler sp. Reduced incidence infectionnsFescue types[386]
Drechslera erythrospilaInhibited hyphal growthnsRyegrass and fescue types[373]
Reduced disease symptoms in plantaProtease and endoglucanase activityE. fesctucae[374]
Drechslera siccans—brown blightResistance to infectionnsRyegrass types[370]
Fusarium avenaceumResistance to infectionEnhanced superoxide dismutase or peroxidases activityRyegrass types[383]
F. avenaceumReduced incidence of infectionnsHost: Elymus cylindricus [384]
F. culmorumReduced incidence of infectionnsHost: Elymus cylindricus
F. oxysporumReduced incidence of infectionnsHost: Elymus cylindricus
Increased resistancensFescue arizonica type[391]
F. poaeReduced incidence of infectionnsFescue types[386]
Fusarium spp.No effect Ryegrass and fescue types[392]
Resistance to infectionnsRyegrass types[370]
Laetisaria fuciformis—red threadLower disease incidence and severitynsMeadow fescue types[393]
Microdochim bolleyiNo effect Ryegrass and fescue types[392]
Phaeosphaeria—leaf spotNo effect Meadow fescue types[394]
Puccinia graminis subsp. graminicolaNo effect Fescue types[395]
Puccinia spp. No effect E. uncinatum[396]
Pyrenophora semeniperda—leaf spotReduced disease symptoms in plantansRyegrass types[397]
Rhizoctonia blightNo effect Fescue types[398]
Rhizoctonia zeaeReduced disease symptoms in plantaPhenolic compoundsFescue types[399]
Reduced hyphal growthnsE. uncinatum[373]
R. solaniReduced incidence of infectionnsFescue types[386]
Sclerotinia homoeocarpa—Dollar spot diseaseLower disease incidence and severityAntifungal proteinMeadow fescue types[400,401]
Typhula ishikariensis—snow moldIncreased susceptibilitynsMeadow fescue types[402]
Ustilago bullata—head smutSuppressed infectionnsE. tembladerae[403]
* ns = not specified.
Table 7. The effect of AR37 endophyte strain in perennial ryegrass on insect pests. (Taken from [260]).
Table 7. The effect of AR37 endophyte strain in perennial ryegrass on insect pests. (Taken from [260]).
Endophyte StrainTillers Damaged by ASW (%)Number of Black Beetles per m2Tillers Damaged by Porina Larvae (%)Number of Root Aphids per Plant *
AR372.12313.62 (0.5)
Standard2.81728.7171 (1.23)
Nil endophyte25.76434.9244 (1.93)
LSD0.0514.22619.9(0.67)
* Log-transformed data in parentheses.
Table 8. Mean total ergot alkaloids and ergovaline concentrations (µg kg−1) in the leaf blade and leaf sheath BarOptima Plus E34, and Kentucky 31 varieties of tall fescue sampled during 2012 and 2014 across Georgia and Kentucky. (Taken from [461]).
Table 8. Mean total ergot alkaloids and ergovaline concentrations (µg kg−1) in the leaf blade and leaf sheath BarOptima Plus E34, and Kentucky 31 varieties of tall fescue sampled during 2012 and 2014 across Georgia and Kentucky. (Taken from [461]).
Tall Fescue VarietyTotal Ergot Alkaloid Concentration (µg kg−1)Ergovaline Concentration (µg kg−1)
Leaf BladeLeaf SheathLeaf BladeLeaf Sheath
BarOptima Plus E34133 b337 b37 b343 b
KY311667 a6312 a268 a2848 a
p-value<0.0001<0.0001<0.0001<0.0001
a,b Within a column, means without a common superscript letter differ significantly (p < 0.05).
Table 9. Mean over two years average daily gain (ADG), grazing days per ha, and blood serum prolactin levels (in February) of 11 month old calves grazed on different endophytic tall fescue pastures in the Coastal Plain region of southwestern Arkansas. (Taken from [463]).
Table 9. Mean over two years average daily gain (ADG), grazing days per ha, and blood serum prolactin levels (in February) of 11 month old calves grazed on different endophytic tall fescue pastures in the Coastal Plain region of southwestern Arkansas. (Taken from [463]).
Tall Fescue and EndophyteADG (kg/day)Grazing Days per haBlood Serum Prolactin (ng/mL)
KY310.585291.5
Endophyte free1.0838462
BarOptima E340.9355338
Jesup AR542 (MaxQ)0.8861179
SEM *0.083014
* SEM—standard error of the mean; for Jesup AR542, n = 2; for KY-31, EF, and BarOptima E-34, n = 3.
Table 10. Mean concentrations (µg/g of DM) in the herbage of measured ergot alkaloids and loline levels (N-formyl loline (NFL) and N-acetyl loline (NAL)), average daily weight gain (ADG) of 2 year old steers, and blood serum prolactin levels across two sites in the USA. (Taken from [465]).
Table 10. Mean concentrations (µg/g of DM) in the herbage of measured ergot alkaloids and loline levels (N-formyl loline (NFL) and N-acetyl loline (NAL)), average daily weight gain (ADG) of 2 year old steers, and blood serum prolactin levels across two sites in the USA. (Taken from [465]).
Tall Fescue and EndophyteEndophyte Infection Rate (% Viable in Seed)Alkaloid Levels (µg/g of DM)ADG (kg/day)Prolactin (ng/mL)
Total Ergot AlkaloidsNFLNAL
HiMag—ArkShield9401611170.6 a155 a
KY31800.703051170.34 b17 b
HiMag—Nil endophyte00000.62 a108 a
a,b Within a column, means without a common superscript letter differ significantly (p < 0.05).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Caradus, J.R.; Johnson, L.J. Epichloë Fungal Endophytes—From a Biological Curiosity in Wild Grasses to an Essential Component of Resilient High Performing Ryegrass and Fescue Pastures. J. Fungi 2020, 6, 322. https://doi.org/10.3390/jof6040322

AMA Style

Caradus JR, Johnson LJ. Epichloë Fungal Endophytes—From a Biological Curiosity in Wild Grasses to an Essential Component of Resilient High Performing Ryegrass and Fescue Pastures. Journal of Fungi. 2020; 6(4):322. https://doi.org/10.3390/jof6040322

Chicago/Turabian Style

Caradus, John R., and Linda J. Johnson. 2020. "Epichloë Fungal Endophytes—From a Biological Curiosity in Wild Grasses to an Essential Component of Resilient High Performing Ryegrass and Fescue Pastures" Journal of Fungi 6, no. 4: 322. https://doi.org/10.3390/jof6040322

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop