Next Article in Journal
Phytoplankton Size Structure and Diversity in the Transitional System of the Aquatina Lagoon (Southern Adriatic Sea, Mediterranean)
Next Article in Special Issue
Diversity of Endophytic Microbes in Taxus yunnanensis and Their Potential for Plant Growth Promotion and Taxane Accumulation
Previous Article in Journal
Emergence of a Hybrid IncI1-Iα Plasmid-Encoded blaCTX-M-101 Conferring Resistance to Cephalosporins in Salmonella enterica Serovar Enteritidis
Previous Article in Special Issue
Utility of Plant Growth-Promoting Rhizobacteria for Sustainable Production of Bermudagrass Forage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Endophytes in Agriculture: Potential to Improve Yields and Tolerances of Agricultural Crops

Department of Chemistry and Biotechnology, School of Science, Computing and Engineering Technologies, Swinburne University of Technology, Hawthorn, VIC 3122, Australia
*
Authors to whom correspondence should be addressed.
Microorganisms 2023, 11(5), 1276; https://doi.org/10.3390/microorganisms11051276
Submission received: 10 April 2023 / Revised: 8 May 2023 / Accepted: 10 May 2023 / Published: 12 May 2023
(This article belongs to the Special Issue Endophytes: Improving Plants Performance)

Abstract

:
Endophytic fungi and bacteria live asymptomatically within plant tissues. In recent decades, research on endophytes has revealed that their significant role in promoting plants as endophytes has been shown to enhance nutrient uptake, stress tolerance, and disease resistance in the host plants, resulting in improved crop yields. Evidence shows that endophytes can provide improved tolerances to salinity, moisture, and drought conditions, highlighting the capacity to farm them in marginal land with the use of endophyte-based strategies. Furthermore, endophytes offer a sustainable alternative to traditional agricultural practices, reducing the need for synthetic fertilizers and pesticides, and in turn reducing the risks associated with chemical treatments. In this review, we summarise the current knowledge on endophytes in agriculture, highlighting their potential as a sustainable solution for improving crop productivity and general plant health. This review outlines key nutrient, environmental, and biotic stressors, providing examples of endophytes mitigating the effects of stress. We also discuss the challenges associated with the use of endophytes in agriculture and the need for further research to fully realise their potential.

1. Introduction

Mutualistic symbiotic endophyte-plant relationships are surmised to be ubiquitous amongst higher plants; every plant on earth contains one or more endophytic fungal (EF) or bacterial (EB) species residing within the leaf, stem, roots, and flower or fruit of plants [1,2]. These relationships have been well documented to commonly give rise to specialised characteristics within plants as a result of the endophyte presence [3]. These properties range from an improved defence against disease and pathogens—antimicrobial, anticancer, antioxidant, anti-inflammatory, insecticidal—to the promotion of plant growth through improved nutrient acquisition and stress tolerance [4].
The relationship between plants and endophytes is one ranging from symbiosis to light pathogenicity. Typically, the plant provides carbohydrates that are vital for endophyte growth, and in return the endophytes’ secondary metabolites can be provided to the plants with tolerances to a breadth of stressors such as drought, increased salinity, nutrient imitation, etc. [1]. As described by Schulz and Boyle [5], 51% of the novel bioactive compounds discovered in-planta were of endophytic origin. This potential pool of bioactive compounds results in the ability to identify and naturally produce composites with medicinal, pharmaceutical, and agricultural applications on a large-scale [6].
Society currently stands at risk of increasing our antimicrobial resistance to bacteria across a range of infections, including bacteria involved in human diseases and plant pathogens causing food spoilage. Both risks weigh upon the fundamental aspects of our society, being healthcare and food supply, which in turn compounds the impact of an ever-increasing population [7,8]. The discovery of novel bioactive secondary metabolites of endophytic origin offers a new direction to combat pathogens. The bioactive metabolites that are produced are considered secondary due to the speculation that they are not essential for the organisms’ growth or reproduction, and instead are a result of evolution in order to protect its source of shelter and nutrition, the host [9].
Endophyte-plant relationships are complex and, only in the last few decades, has the web begun to untangle, with positive characteristics being reported in individualised endophyte-plant interactions. Often, reports are heavily based in favour of the beneficial characteristics, as expected, where confirmational evidence is found. However, negative impacts are scarcely reported or investigated, leaving space for scepticism amongst the theories for large scale applications. As the knowledge base of endophyte-plant relationships expands, it is important to conduct investigations with due diligence in both the positive and negative effects of endophytes, to gain a true understanding of the capabilities of their applications.

2. Requirements for Increased Crop Productivity: Climate Change and Population Growth

The population is expanding at a rate that poses a threat to the current capacity of agricultural practices and improvements to crop yields is required to sustain the continuous growth. The global population is predicted to top 9.7 billion in 2050 and 10.4 billion by 2080 [10], not only increasing the total produce yield required to sustain said population, but also reducing the current farmable land in consequence of an increased demand for housing land. As outlined in Figure 1, providing residential land (grey) from limited farmland(green) provokes a need to improve crop tolerances to colonise previously unsuitable/marginal land (orange). Particularly in Australia, farmable land is limited by the water availability and the soil salt concentration, therefore, increasing the ability of plants—with endophytes—to grow at lower moisture or higher salt concentrations unlocks this land for agricultural uses [11,12,13].
The impacts of this increased demand of residential land—if left unsolved—further compound problems of starvation across Asia and Africa (Figure 2). Alongside starvation rates increasing to 9.9% (globally) in 2020, up from 8.4% in 2019, a gradual rise of undernutrition, consequence of the covid-19 pandemic and its effects on global food supply, is projected [14]. It is vitally important that starvation trends take a downward trend if the projected population figures are to come true. Failing to address the current starvation and undernutrition could exponentially increase food supply chain disruptions, impacting low-socioeconomic/third-world populations significantly. Although dire situations such as that in Africa require major interventions in irrigation and fertilisation, endophyte-enhanced plants can still play a role in the biofortification of crop plants.
Nutritional crop limitations, with regards to essential dietary nutrients, are also associated to undernutrition. Therefore, it is important to understand the properties of agricultural crops in relation to the nutritional limitations to nourish troubled nations efficiently and effectively [15].
Figure 2. Food insecurity illustrations representing the severity of insecurity from 2016 (A) to 2019 (B), outlining the increasing nature of food supply issues, particularly in southern Africa where food supply and undernutrition is increasing at an alarming rate [16].
Figure 2. Food insecurity illustrations representing the severity of insecurity from 2016 (A) to 2019 (B), outlining the increasing nature of food supply issues, particularly in southern Africa where food supply and undernutrition is increasing at an alarming rate [16].
Microorganisms 11 01276 g002
Crop loss as a result of pests and diseases is the primary income loss for farmers and local economies [17]. In a situation where the world can ill afford waste, biotic stress losses compound the impact of a growing population. Crop loss consists of an initial loss through the direct interaction with pests and pathogens, as well as the secondary losses caused by the indirect effects of diseases and nutritional content within the soils, having long-term effects [18]. Therefore, it is difficult to quantify the secondary impacts on crop loss due to a range of random variables that are difficult to account for, such as the prominence of pests and disease at the given time of analysis, in comparison to other seasons, due to the repetitive cycle of growth between harvest and sowing seasons. This gives little opportunity to distinguish the nature of loss between crop generations [19].
Crop yield loss is highly dependent on the treatment of the crops, with fungicides, pesticides, and insecticides being the driving force behind increasing crop productivity. The ability to substitute the use of harmful chemicals with naturally sourced compounds produced by endophytes in-vitro, is expected to have a large impact on the long-term health of all living organisms involved in the process from planting to consumption, whether that be humans or other animals [18].

3. Agricultural Potential of Endophytes

Current industrial approaches to combatting stressors are through fertilisers, herbicides, and pesticides which have several questionable effects on consumers and long-term soil quality, where taking a natural approach would offer a less harmful and more consumer-friendly product [20]. Endophytes—both bacteria and fungi—have emerged as highly potent natural resources, which have been utilised to improve stress tolerances of plants post-inoculation and decreased susceptibility to diseases and predators [21].
The overall topic of endophytes is novel, with the main interests in their applications only coming to the forefront in recent decades, these applications being outlined in Figure 3 below. The knowledge of potential solutions that endophytes can provide remains largely incomplete, as defined by the estimated 1 million endophytes on earth and only 100,000 of which have been appropriately described and studied, as of 2015 [1]. Studies conducted in this space typically take an approach of fungal or bacterial analysis, but rarely both. This could potentially undermine or pass over the key complex trends within plants with desirable characteristics. It is important that studies begin to understand the entire “endomicrobiome”—the community of microorganisms that reside within the plant tissue—to understand the complete internal environment and the interplay between species [22]. There are still many unknown elements to endophyte-plant interaction [1], however, a holistic approach provides the potential to give context to the plants internal biome via transcriptomics, proteomics, and metabolomics to understand all aspects of endophyte intervention on plant health [23].
Endophyte presence has shown significant benefits in a number of studies, commonly improving the fresh and dry weights of endophyte-associated plants (EAP) two-fold [24,25,26,27]. As outlined by Singh et al. [28]’s study on the effects of endophytes on zinc and iron deficiencies, EAP consequently gained 2–2.3 times fresh and dry weight as a result of improved nutrient intake abilities. Whereby, the endophyte presence resulted in increased indole-3-aceatic acid (IAA) and phytosiderophore production, which increased nutrient solubilisation significantly [28].
The summary table below (Table 1) outlines the wide variety of reported benefits from the literature reviewed, some of which are known plant pathogens in some species yet provide beneficial impacts to other, unrelated species of plants. The complexities of the genetic factors involved in endophyte intervention strategies must be further researched [29]. It is highly likely that the endophytes listed in Table 1 could have further or different beneficial characteristics when tested on different model plants. This is a key problem that requires solving for large scale approaches to improving crop health.
Recruiting beneficial endophytes in a field application is not as simple as theoretically suggested because the complex relations between the biosystems of soils (rhizospheric microbes and actinomycetes) and plants make for a more difficult introduction of any microorganisms [30]. It has been suggested that the colonisation of a seedling is a convoluted event that is very competitive with other microorganisms and the presence of inhibitors within seedlings [31]. Particularly within the initial 24 h post-sowing where the colonisation by endophytes as early as possible tends to improve a sustained presence in mature plants [32,33]. Inoculation approaches apply to all plant-microbe interactions that cannot undergo vertical transmission, which is relevant for many crop plants such as wheat and grains as they do not readily transmit microbes, posing potent difficulties in the pursuit of large-scale farm-based applications of EAP [33].
There are many strategies currently being utilised to combat micronutrient deficiencies in crops, some of which being the direct supplementation of nutrients, direct genetic intervention on plants, fertilization, and food fortification [34]. Although they each have pros and cons, the endophytic treatment of plants will soon be potentially among the strategies to improve food nutrition. Many of the major diet-based nutrient deficiencies globally (iron, zinc, etc.) have direct connections to the benefits identified from EAP [28,35,36]. Theoretically, endophyte bioprospecting may give rise to new approaches that internally improve food nutrition, without the need for potentially harmful chemical treatments or GMOs. However, significant knowledge gaps must be filled, and the effects of the endophytes must be understood before anything can be deemed safe for human consumption.
Table 1. Summary table of endophytes reported to have agricultural applications.
Table 1. Summary table of endophytes reported to have agricultural applications.
EndophyteHost Plant *EffectReferenceType
Acremonium sp. Lp1,2Lolium perenneAntifungal[37]Fungi
Aspergillus terreus AM2Moringa oleiferaAntifungal, IAA production[38,39]Fungi
Trichoderma harzianumHardwood barkAntifungal, P solubilisation[40,41,42]Fungi
Epichloe festucae Fr1, 11, E365Lolium perenneAntifungal[37]Fungi
Epichloe sp.Not reportedBiotic stress tolerance[43]Fungi
Paraphaeosphaeria sporulosaActinidia deliciosaBiotic stress tolerance[44,45]Fungi
Cochliobolus sp. 23-1Panicum coloratum, Chloris gayanaCa accumulation[46]Fungi
Ophiosphaerella sp. 15-2Panicum coloratum, Chloris gayanaCa accumulation[46]Fungi
Penicillium bilaiiElaeis guineensisCa accumulation[47]Fungi
Penicillium oxalicum P4Elaeis guineensisCa accumulation[47]Fungi
Setosphaeria rostrata GR1APanicum coloratum, Chloris gayanaCa accumulation[46]Fungi
Fusarium fujikuroi IMI58289Manihot esuclenta, Oryza sativaGA production[48,49,50]Fungi
Gibberella fujikuroiManihot esculentaGA production[48,49]Fungi
A. lipoferumZea maysGA production[51]Fungi
Penicillium citrinum IR-3-3Ixeris repenesGA production[52]Fungi
Penicillium commune KNU5379Seasamum indicumGA production[49]Fungi
Penicillium funiculosumGlycine maxGA production[53]Fungi
A. brasilenseZea maysGA production[51]Fungi
Sphaceloma manihoticola Lu949Manihot esuclentaGA production[48,49]Fungi
Aspergillus fumigatus LHL06Glycine max, Zea maysGA production, pathogen resistance[52,54]Fungi
Beauveria bassianaVitis vinifera, wheatGeneral growth, pest resistance[55,56,57,58]Fungi
Colletotrichum tofieldiae CT04_08450Arabidopsis thalianaGeneral growth, hormone production[59]Fungi
Diaporthe sp.Festuca rubraGeneral growth, IAA production, N, Ca, Mg, Fe accumulation[60]Fungi
Neotyphodium lolii NEA4Lolium perenneGeneral growth, Pest deterrence, pathogen resistance excl. staggers[61]Fungi
Acremonium coenophialumFestuca arundinaceaGrowth, pest resistance, pathogen resistance[62]Fungi
Aspergillus flavusEuphorbia geniculataIAA production, pathogen resistance[63]Fungi
Piriformospora indicaMultipleIAA production, salt tolerance[64,65]Fungi
Gilmaniella sp. AL12Atractylodes lanceaJasmonic acid production[66]Fungi
Phomopsis liquidambariBischofia polycarpaN & P accumulation[67]Fungi
Heteroconium chaetospiraChinese cabbageNitrogen metabolism, general growth[68]Fungi
Aspergillus niger AP5, P85Arachis hypogaeaP solubilisation, IAA production, Ca accumulation[47,69]Fungi
Glomus mosseaeNot reportedP, K, Mg, Cu, Zn, Mn accumulation[70]Fungi
Trichoderma arundinaceumNot reportedJasmonic acid production[71]Fungi
Trichoderma harzianum TRI5Not reportedBiotic stress tolerance[72]Fungi
Trichoderma virens IB 119/12Glycine maxPathogen resistance[73]Fungi
Cladosporium cladosporioides BOU1Solanum melongenaPest resistance[74]Fungi
Epichloe coenphiala AR584Tall fescuePest resistance[75]Fungi
Fusarium oxysporum 24o, V5W2Musa spp.Pest resistance[76,77]Fungi
Metarhizium anisopliae QS155Not reportedPest resistance[41]Fungi
Epichloe gansuensisAchnatherum inebriansSalicylic acid production[78]Fungi
Bacillus subtillis 26DNot reportedBiotic stress tolerance[79]Bacteria
Streptomyces sp. 11EVigna radiataAuxin production, N fixation, salt tolerance, siderophore production[80,81]Bacteria
Azospirillum brasilense B510Oryza sativaIAA production, N fixation[82,83]Bacteria
Paenibacillus sp. ANM59, ANM76Cicer arietinumIAA production, P solubilisation, salt tolerance[84]Bacteria
Bacillus thuringiensisZea maysInsect deterrent[85]Bacteria
Ewingella americana EU-M4ARActZea maysK accumulation[86]Bacteria
Pantoea agglomerans EU-E1RT3-1Zea maysK accumulation[86]Bacteria
Pseudomonas brenneri EU-A2SK1Zea maysK accumulation[86]Bacteria
Mesorhizobium ciceri BRM5Cicer arietinumN fixation, IAA production, salt tolerance[87,88]Bacteria
Azospirillum brasilense Cd, Az39Triticum aestivumN fixation, P solubilisation, auxin production[89,90]Bacteria
Bacillus sp. 13EVigna radiataN fixation, P solubilisation, auxin production[80]Bacteria
Bacillus endophyticus 14EVigna unguiculata, soybeanN fixation, P solubilisation, auxin production, salt tolerance[80]Bacteria
Bacillus altitudinis Q7Ginkgo bilobaPathogen resistance[91]Bacteria
Bacillus polymyxa L6Not reportedPathogen resistance[92]Bacteria
Serratia plymuthica HRO-C48Not reportedPest resistance[93]Bacteria
Burkholderia sp. SSGBeta vulgarisS metabolism, N fixation, K accumulation, IAA production[94,95]Bacteria
Azotobacter chroococcum Avi2Not reportedSalt tolerance, drought stress, general growth[96]Bacteria
Serratia marcescens AL2-16Achyranthes asperaSiderophore production, IAA production, ammonia production, general growth[97,98,99]Bacteria
Pseudomonas fluorescens L228, L111, L321Not reportedSiderophore production, P solubilisation, general growth, pest resistance[100]Bacteria
Enterobacter sp. SA187Indigofera argenteaSulphur metabolism, reduced ROS accumulation, salt tolerance[101]Bacteria
Burkholderia phytofirmans PsJNNot reportedZn accumulation[102]Bacteria
Staphylococcus hominis 7EVigna radiataZn, P accumulation, hormone production, antifungal[80]Bacteria
* “Host plant” refers to the plant from which the organism was been isolated for the first time.

4. Plant-Endophyte Relationship

Unpredictable environments consistently place stressors upon plants through abiotic and biotic sources and it is the plants prerogative to efficiently mitigate these stressors to maintain homeostasis [2,3,5,103,104,105,106,107,108,109,110]. The impact that endophytes have on the internal environment and behaviours of a plant is closely tied to the gene regulation and signalling as a result of the endophyte presence [111]. Dependent upon the origin of the stressor, plant-endophyte interactions differ through the type of endophytes present, the signalling mechanism, the secondary metabolites, etc. The effects of the endophytes on plants are constantly being investigated and new depths of the relationships between the two are being uncovered [112]. In the many instances of the significant effects of endophytes on plant behaviours—nutrient uptake improvements, greater stress tolerances, etc.—there are no direct and complete descriptions of the mechanisms by which this is achieved.
Plant-endophyte symbiosis is best demonstrated as the plants observed providing EF with carbon sources, mainly in the form of sucrose as determined by tracing 13CO provided to the plant. This evidence supports the symbiotic model between the fungi and plant, where the plant is nurturing the endophytes within them, suggesting that endophyte presence is not of a parasitic nature [68]. Additionally, Usuki et al. [68] identified a negative correlation between H. chaetospira and soil nutrition (carbohydrate, nitrogen, and phosphorus), as the plant’s needs are met by the environment the endophyte impact is reduced. This evidence outlines the symbiotic nature of the plant-endophyte interactions, whereby plants rely on alternative pathways of nutrient acquisition when soil nutrition is sub-optimal. This aalso suggests that plants may have a selective permeability to specific endophytes relative to the plants’ internal and external environments [113].

5. Role of Endophytes in Plant Health

5.1. Biotic Stressors

5.1.1. Pathogens

As aforementioned, primary crop loss is as a result of pathogen and pest intervention impacting agricultural practices to the degree of billions of dollars [114]. Pathogens, in which there are more than 100 differenct types of plant pathogens, can be spread throughout crops via airborne propagation, insect vectors, or via introduced methods such as irrigation [115,116]. Delgado-Baquerizo et al. [117] connected warmer climates to an increased number of pathogens in plants, linking climate change to increased disease levels. An analysis of the entire microbiome can aid in understanding the difference in behaviour between harmful pathogens and the development of beneficial endophytes, as well as explaining any shift in endophyte behaviour, from beneficial to harmful [22].
Endophytes can improve plant biosecurity through the improved antimicrobial activity of EAP, as developed in some commercially available perennial ryegrasses [61]. The majority of literature surrounding endophytic improvement relative to biotic stressors comes from investigating Epichloe endophytes, a family of fungi with strong relationships with grasses that have both beneficial and pathogenic effects [78,118]. Beneficial endophytes belonging to Epichloe improve biotic resistance through a range of mechanisms, mainly through the upregulation of jasmonic (JA) and salicylic acid (SA) production pathways—as discussed following [78,118,119]. Xia et al. [43] studied the impact of Epichloe endophytes on plants infected by powdery mildew (Blumeria graminis), suggesting that EF are able to mitigate the chlorophyll loss of activity and damage (measure of disease severity) compared to non-EF containing plants.
Along with antimicrobial properties provided by endophytes, increases in metabolic and photosynthetic rates in EAP help combat the effects of pathogen presence. EAP have shown to have higher rates of growth under pathogen infection, compared to the basal rates of non-EAP [118]. It could be considered that larger and healthier plants have reduced pathogen impact and increased capabilities to mitigate infection [44]. This is exemplified by endophytic increases in calcium (Ca) acquisition, which is involved in the systematic acquired resistance signalling, whereby healthier levels of Ca result in an appropriate ability to respond to stressors [120,121,122].
The antimicrobial properties of endophytes are on display within kiwifruit plants, as their associated endophytes’ can resist canker disease caused by epiphytic bacteria Pseudomonas syringae [44,45,123]. Paraphaeosphaeria sporulosa is documented as an associated EF to kiwifruit plants (Actinidia deliciosa) with antimicrobial activity against P. syringae through the production of diketopiperazine [44,124]. Diketopiperazine—a class of organic compounds—are known to have antimicrobial properties, especially those that contain proline, which is a common factor in stress tolerances [125]. Similar results have been reported in wheat pathogens, causing some of the most economically destructive diseases. These diseases, caused by fungi Magnaporthe oryzae Triticum (MoT) and Fusarium graminearum, can contribute to extraordinary amounts of crop loss [72,79,126]. Chakraborty et al. [79] discovered lipoproteins produced by Bacillus subtillis 26D capable of inhibiting MoT growth and germ tube formation in vitro. Taylor et al. [72] outlines that the endophyte Trichoderma harzianum’s ability to produce trichodiene—a known biocontrol agent—is a sesquiterpene that inhibits Fusarium graminearum, the pathogen responsible for fusarium head blight.
The complexities of endophytic communities can be demonstrated through the disease ryegrass staggers, a common condition in sheep, cattle, and horses during winter as grazing approaches roots containing Neotyphodium lolii [127,128]. Alkaloid toxins produced by N. lolii are harmful to livestock but can be beneficial to the plant resistance to stem weevils through the same mode of action [127,129]. It is the complex balance between benefit and cost that must be assessed to responsibly apply these novel techniques. Increasing the biomass of crops reduces the amount of N. lolii containing roots consumed by livestock, but does not mitigate it entirely, therefore it is currently a case-by-case scenario for applying these techniques.
Examples, such as those aforementioned, provide validity to the theoretical applications of endophytes as biocontrol agents. However, there must be more research surrounding the impacts of the introduced species on the natural biome of a plant, alongside the cost to the plants and any downstream consumers in order to ensure safety of the practice [130,131]. As there are many secondary metabolites produced by the endophytes, there is always potential that some of these compounds can negatively impact the plants and consumers if they are left to bioaccumulate [131].

5.1.2. Secondary Metabolites

The metabolic profiling of endophytes in recent years has identified a wealth of secondary metabolites with antimicrobial, anticancer, and pesticidal properties. Including compounds such as flavonoids, carotenoids, melatonins, terpenoids, phenolics, alkaloids, peptides [132,133]. 42% of the newly approved drugs between 1981–2019 are of, or a derivative of, natural sources, many of which are produced by endophytes [134]. Although the large majority of the metabolite discoveries are found in the context of medicinal applications, the same bioprospecting principles can be used in the context of agricultural practices [133]. Many metabolites with agricultural applications are multifunctional in terms of plant growth promotion (PGP), along with improved abiotic or biotic stress tolerances.
Endophyte-derived bioactive compounds such as acyl-homoserine lactones, diacetylphloroglucinol (DAPG), and pyoluteorin, have been utilised in managing soil-borne pathogens, suppressing weeds, and PGP [135,136]. The application of these compounds to crops offers the potential for the development of eco-friendly, low-cost, and sustainable agricultural practices. Similarly, non-ribosomal peptides and polyketides have been identified as components of the systemic acquired resistance (SAR) in plants, which can strengthen the host plant immunity to various pathogens [137]. Additionally, amongst these secondary metabolites are plant hormones, which stimulate the germination of seeds, root growth, and nutrient uptake [101]. The origin of many of these metabolites are not well understood, as evidence outlines relationships where the compounds are produced by plants, microbes, or co-production between both organisms. Therefore, an understanding of the endophytic secondary metabolites and the pathways by which they are produced is vital to enhance plant productivity and resistance to pests in agricultural practices.

5.1.3. Pests

Complementing pathogen resistance, endophytes can impact the other half of biotic stress tolerances, which is pest resistance [138,139]. Volatile organic compounds (VOC) produced by plants can have a variety of pesticidal functions, they can be directly toxic, a repellent, or an attractant for pest pathogens (caterpillars, spiders, etc.). The term VOC is a broad umbrella term encompassing compounds such as pheromones, phenolics, and alkaloids [139,140]. It is widely reported that endophytes produce VOCs as well as stimulating VOC production within plants via the jasmonic pathway (JA). Evidence shows that EF stimulating this pathway through the production of key compounds in the signalling cascade, thus increasing resistance to pests through antiherbivore VOC production [141,142]. Endophytes that possess this ability are a viable biocontrol agent in principle. Unlike many other stressors, biotic resistance is one of the only stress tolerances that currently have field applications in the form of EAP [143,144]. Commercially available ryegrass strains AR1, AR37, NEA2, and NEA4 are examples of EAP to improve yield, increasing lateral root growth by up to 92%, and significantly increasing the number of tillers aiding in persistence. Furthermore, these strains show increased resistance to weevils and black beetles by approximately 75% and 60%, respectively [61].
The most studied and widely available biopesticide is the use of Bacillus thuringiensis, a pore-forming toxic protein producing EB with insecticidal characteristics. Cry toxin expression in crops has shown to improve pest resistance, as it impacts the larval stages of many phyla of insects [85]. The pore-forming toxins target the digestive tracts of the larvae through midgut cell membrane insertion resulting in osmotic shock. B. thuringiensis products are commercially available as sprayable products and take up 35% of the biocontrol agent markets [77]. Second to the aforementioned, Beauveria bassiana is the next most studied endophytic biopesticide, and is an EF capable of parasitising more than 200 insect species [57]. Wang et al. [145] describes in great detail the production of degrading enzymes by which B. bassiana is able to infiltrate pest hosts, followed by the production of the insecticidal compounds beauvericin, bassianin, bassianolide, beauverolides, tenellin, oosporein, and oxalic acid, all of which contribute to the interference with cell functions and antimicrobial responses within the host insect [146,147]. Similar insecticidal characteristics have also been reported from other entomopathogenic microbes, including Rhizoctonia solani, Fusarium sp. [77], Epichloe coenophiala [75], Pseudomonas fluorescens [98,99], Serratia plymuthica [93], Cladosporium cladosporioides [74], and Metarhizium anisopliae [41].
The discussed endophytes represent a proof of concept of endophyte-based approaches to be implemented into agricultural practices. Not only as biocontrol agents, but to open the door for human-consumed crops to receive treatments relative to nutrient deficiencies and abiotic stress tolerances to improve yields.

5.2. Phytohormones

5.2.1. Auxins (Indole-3-Acetic Acid)

Endophyte presence often promotes plant growth as a result of the phytohormone production or upregulation, and is the most widely known mechanism for physiological and structural changes [110]. The production of phytohormones is a direct relationship to growth signalling. Hormones produced by plants, EB, and EF—such as Indole-3-acetic acid (IAA)—can stimulate taproot/adventitious root growth [148]. IAA is a multifunctional signalling molecule that transcriptionally effects small, nuclear, short-lived proteins resulting in a range of auxin-regulated transcription throughout the cells [149]. The introduction of phytohormone-producing EB and EF to plants in stressed conditions has shown clear persistence across a range of stress parameters [150].
Piriformospora indica is a key EF identified to produce IAA as a secondary metabolite, and is one of the most commonly studied EF as a plant inoculant [65,151]. Auxin signalling has been reported to be involved in iron solubilisation by inducing Fe (III) reductase in Malus xiaojinensis. Furthermore, IAA signalling is associated to 1-aminocyclopropane-1-carboxylate (ACC) deaminase activity, an important enzyme involved in limiting the effects of salt stress [152,153]. ACC deaminase degrades ACC, a precursor to ethylene, which is a plant hormone responsible for the reactive oxygen species (ROS) accumulation under salinity stress [153,154]. Of the literature reviewed and current knowledge base, endophytes capable of producing auxins have a strong relationship with improved plant growth, one of the strongest impacts on overall plant behaviour and physiology [110,155].

5.2.2. Gibberellins

Gibberellins (GA), like auxins are a class of growth promoting phytohormones involved in a vast array of effects on plant cells, particularly stem elongation, flowering, and leaf spread. 136 GAs have been discovered, only a few of which meet the criteria of bioactivity (containing 19 carbon backbone and hydroxyl group), GA1, GA3, GA4, and GA7 have been identified as the most bioactive of the GAs [48,156,157,158], however, GAs that are considered ‘inactive’ are precursors to bioactive GAs with subtle differences, which can be assimilated enzymatically to form bioactive GAs [150,157]. GAs are regulated by DELLA proteins via a negative feedback loop. DELLA repression due to the binding of DELLA and GID1—a GA receptor—blocks signalling until the repressor is ubiquinated by ubiquitin-transferase SCF [159]. There is evidence to suggest that endophytes play a role in the transcription of DELLA and SCF proteins, in turn improving GA effects alongside any GAs produced as secondary metabolites [101]. GA-producing endophytes have been characterised down to the presence or absence of GA gene clusters. Endophytes that possess these genes typically belong to Ascomycetes phyla [150]. Association with GA producing endophytes has shown to benefit plants, yet the direct mechanisms are yet to be appropriately described [150]. Identifying a key trend in the characteristics of a taxonomical classification—contrary to the majority of endophyte knowledge—are rare in the current climate of understanding characteristic trends relative to the beneficial properties that endophytes can bestow on plants.

5.2.3. Stress Response Hormones

Three phytohormones commonly associated with a range of stress responses can be collectively termed as the stress response hormones, and include abscisic acid (ABA), salicylic acid (SA), and JA [160]. Current knowledge of these three systems suggests a complex interplay between the three signals and that the presence of endophytes only entangles the relationships more [161,162].

Abscisic Acid

Phytohormone ABA is commonly known as the plant stress hormone, involved in major signalling pathways involved in biotic stress response, as well as being responsible for general plant growth outside of the stressed conditions [163]. Under drought stress, ABA is produced rapidly by plants to aid in the plant’s survival in sub-optimal conditions. Increased levels of ABA stimulate stomatal closure through H2O2 production in order to preserve moisture, as well as decreasing lateral root growth and increased vertical (main) root elongation to vitally improve water uptake, and accelerate sentience [163]. However, contrary to early theories, endophyte association negatively correlates with ABA levels, as shown by Waqas et al. [164] who reported a negative correlation of ABA and Penicillium sp. and Pterolepis glomerata. Khan et al. [150] suggest that this indicates that ABA is an independent mechanism of stress mitigation. Conversely, Wang et al. [161] reported positive correlations between EB and ABA levels, suggesting that conjecture remains surrounding the effects of endophyte presence on ABA levels, posing the potential for ABA-dependent and ABA-independent pathways. Despite these findings, it must be considered that the negative relationships have been more heavily reported. This is potentially due to the salinity and drought stress that the plant finds itself under; ABA produces H2O2—a ROS—which is undesirable under saline conditions, and therefore reducing levels of ABA is a likely plant resistance mechanism to minimize ROS accumulation [163].

Salicylic Acid

Although endophyte presence typically reduces the ABA levels in plants, theoretically reducing the drought tolerance levels of plants, the SA system can counter for the reduced defence mechanisms that ABA provides [6]. SA, a phytohormone capable of initiating the systematic acquired resistance (SAR), provides a whole plant response to the localised infection by pathogens [165]. The SAR is a wide affecting system much like the innate immune system in animals. Plant SA production occurs through isochorismate synthase and phenylalanine ammonia-lyase pathways, however, not all steps of these pathways have been discovered [78,166]. Incomplete knowledge of these mechanisms creates a gap when assessing the impacts of endophytes on SA production. Understanding the plant mechanism in its entirety can potentially uncover key compounds and enzymes involved in the pathogen resistance in EAP [167].
Kou et al. [78] assessed the effects of Epichloe gansuensis on Achnatherum inebrians, its ability to increase the production of SA, and its pathogen resistance relative to endophyte presence. EAP showed increased levels of SA and the transcription of genes relevant to the SA pathway. Increased SA production induces a pathenogenic-related gene transcription, involving bacterial and fungal lysing enzymes [168]. The mode and degree of action of the SAR is relative to the specific signalling and context of each cell, making it difficult to accurately quantify the impact of the variables, such as endophytes and other phytohormones [169]. Acibenzolar-S-methyl is the most studied analog of SA and is a key inducer of the SAR, however, its function remains incompletely understood like most of the SAR. The literature is limited to a broad description of the SAR and remains vague when explaining the mechanisms by which plants act against pathogens. Evidence suggests that the SAR induces pathogenesis-related proteins [168,169]. This is due to the large number of proteins and genes involved in the pathogenesis and the SAR, which include antimicrobial peptides, other phytohormones, protein & MAP kinases and transcription factors [170].

Jasmonic Acid

Jasmonic Acid (JA) is a key phytohormone involved in a wide variety of cellular regulation from senescence, flowering, and abscission. JA is also the key signalling molecule involved in stress signalling as a result of plant wounding from pests. More specifically, JA is released when the plant is consumed, stimulating CsKPI genes responsible for Kunitz protease inhibitor (KPI) expression—a known anti-herbivore compound [141,142,171,172]. Protease inhibitors, when consumed, prevent the digestive enzymatic activity of insects which reduces the metabolic potential of the insect, thus, making the plant undesirable as a source of food and increasing pest (insect) resistance as a result of the increased levels of JA [173]. CsKPI genes have been the subject of genetic engineering, however, the conjecture and limitations surrounding GMOs leaves an opening for alternative methods to promoting genetic transcription, including endophytes [171].
Endophytes have shown clear implications on the JA levels of plants, often antagonistically to SA pathways [173]. The direct treatment of plants with JA are sub-optimal due to this relationship with SA, as it leaves plants susceptible to viral and bacterial infections [174]. Endophyte Gilmaniella sp. provided evidence that endophytes produce JA and VOC in Atractylodes lancea [66]. Mengistu [71] explains the pathway by which EF Trichoderma arundinaceum stimulates the JA related defence genes, achieved through the production of trichodiene [72,175]. Trichodiene, the forementioned compound, with activity against F. graminearum, which is the pathogen responsible for wheat blight [72].
The complexities of the interrelationships of the ‘stress signalling hormones’ (JA, SA, ABA) have only recently begun to be appropriately understood. Yet the complete pathways of each hormone—both irrespective and respective of each other—must continue to be described [161,162]. This is in order to most appropriately apply the theories of endophytic inoculation on agricultural practices.

Cytokinin’s

Cytokinins (CK) are a class of phytohormones involved in apical dominance and lateral root growth signalling which directly act as an antagonist to auxins. Recently, CKs have been the subject of studies regarding pathogen resistance, with evidence suggesting that CK are involved in regulating the SA pathway, as well as SA independent resistance [167,176,177,178]. Großkinsky et al. [178] identified that the endophyte G20-18 was capable of controlling the forementioned P. sryingae infection, when CK levels was adequate. Reduction in these levels diminished the capabilities of G20-18 to control the bacterial pathogen [178]. It has been reported that parthenogenesis occurs by regulating the host CK production, and external sources of CK such as endophytes offer alternative routes to pathogen suppression via CK [179,180]. The interference of pathogen development as a result of endophyte presence has been identified to occur amongst the cytoskeleton and cell trafficking of pathogens. Gupta et al. [181] tested CK and derivatives directly against B. cinerea and found that zeatin, kinetin, thidiazuron, and 6-benzylaminopurine were able to inhibit the growth of the fungi. The suggested mode of action is the interference of proliferation through spores and tube elongation.
The well-known aspect to CK is also relevant to benefiting from endophytic association, as higher levels of CK results in the stimulation of lateral root growth and an overall increase in biomass [151,167]. This is the case with P. indica association, whereby plant biomass is significantly increased due to its presence as a result of higher CK levels [167]. As aforementioned, endophytes have the capacity to produce phytohormones as well as stimulate the transcription of them. This is the case with CK where some endophytes have the capacity to produce their own CK to improve plant growth [164,182].

Reactive Oxygen Species

ROS are derivations of oxygen that are more reactive than molecular oxygen (O2), and are signalling molecules that are involved in plant senescence, degrading cellular components which are no longer required, and maintaining redox homeostasis. Under stress conditions, ROS production increases significantly and can surpass the metabolism capacity of osmolytes which results in accumulation [183,184,185]. Accumulation of ROS caused by stress can seriously damage the cellular components such as DNA, RNA, and proteins, potentially resulting in cell death [186]. H2O2, super-oxide (O2), hydroxyl radicals (HO), and nitric oxide (NO) are all ROS, where plants combat the accumulation through antioxidants such as proline, peroxidases (POD), superoxide dismutase (SOD), catalase (CAT), gamma-aminobutyric acid (GABA) [183].
Endophyte presence has shown to positively impact the osmolyte and antioxidant production within plant tissues, providing an improved ability to metabolise ROS [119,187]. Studies investigating the inoculation of B. subtillus and its effect on stress tolerance improvements found that the EB stimulates CAT, SOD, and POD production in chickpeas. As a result, plants experience greater tolerance under salt stressed conditions, however, these results were observed in both saline and optimal conditions and is therefore not specific to stressed conditions in these circumstances [188,189]. Contrary to this, a study of glutathione peroxidases (GPX) conducted by Bela et al. [185] identified genes relevant to GPX production activated by oxidative stress, that resulted in reduced cell death caused by heat and salinity, and could potentially act as cellular redox sensors.
As Khan et al. [189] identified, endophyte presence can increase antioxidant production in all conditions, increasing the basal level of antioxidants. Compounded with oxidative stress-induced reactions to salinity which further increases the overall defence to ROS. Improved nitrogen availability has shown to improve salt tolerance due to increased proline biosynthesis, as proline is a secondary amine [190]. The ability of the endophytes to enhance bioavailable nitrogen to plants provides another direct link from endophyte-based improvements and increased stress tolerance.
Hormones produced by plants and microbes are involved in a complex model of regulation and identifying the hormone levels in plants is one of the best insights into the context of a plant’s health. Table 2 below summarises the key effects of each plant hormone affected by endophyte presence.

5.3. Nutrient Limitations

5.3.1. Iron

One of the most common deficiencies in plants is Fe deficiency. Fe is involved in the production of chlorophyll, is a common enzymatic co-factor, and plays a major role in the electron transport chain (cellular respiration) [197]. Alkaline soils make up approximately 30% of the farmable land and cause Fe deficiencies more than any other condition, therefore overcoming this limitation greatly improves the ability to farm on marginal land [198]. Graminaceous plants—most grass and cereal species—respond to conditions of Fe deficiency by producing phytosiderophores (PS), a class of organic substances that can solubilise rhizospheric Fe and Zinc (Zn), known as strategy I [6,35,199]. Some endophytes produce and excrete PS around the roots of non-graminaceous plants, solubilizing Fe (III) through the binding of mugineic acids (MA), conjugating to form Fe (III)-MA which resolves the requirement of reducing Fe (III) to Fe (II) for plants to absorb rhizospheric metals as outlined in Figure 4 (strategy II) [36]. Studies have suggested that this mechanism of PS-involved metal transport can solubilise other metals such as Zn, copper (Cu), cadmium (Cd), and Ni [200]. Thus, identifying a direct link between endophyte presence and an improved response to metal deficiencies via PS production. Fe plays an important role in a number of cellular processes downstream and Fe deficiencies extend beyond direct Fe cellular processes, impacting plant growth across the board [28,201].
The direct mechanisms of impact of endophytes on Fe accumulation are poorly understood. Tt has been speculated that Fe deficiency stimulates ROS accumulation, resulting in a suppression of Fe-related and regulated genes FIT, IRT1, FRO2 [28,202,203]. Additionally, endophytes have been observed to stimulate IAA induced Fe(III) reductase activity, aiding in Fe accumulation via strategy I [192]. Thirdly, endophytes have been observed to increase PS levels in plants, giving strategy II capabilities to [109,201]. Regardless of the mode of action, endophytes have shown to play key roles in improving Fe availability and represent a potential solution to the most common consumer nutrient deficiency [28,46]. However, reported results originate from the investigations of relationships between individual endophytes and plants. To appropriately analyse and describe the effects of endophytes on Fe acquisition, a larger knowledge base is required to assess trends across phylogenetic groups of both plants and EF and EB, for there is potential for multiple points of injunction on plant processes.

5.3.2. Zinc

Zn behaves similarly to that of Fe relative to plant nutrition; typically, deficiencies occur in calcareous and clay containing soils due to adsorption of reactive Zn2+ ions. Zn is the second most organically abundant transition metal after Fe [204]. Deficiencies result in chlorosis, stunted plant growth, and reduced cellular functions as Zn is an important enzymatic co-factor [204,205]. It is important to ensure the mitigation of dietary inadequacies in consumers down-stream, and nutritionally fortifying the crops that are most susceptible to Zn deficiency, such as corn, provides a solution to undernutrition.
Zn2+ can be acquired by graminaceous plants through two strategies, the first of which is the chelation pathway, which also falls under strategy II nutrient acquisition [6]. Chelation is the process by which plant or endophyte exudates (typically polydentate, or multi-dentate, ligands) bind with the charged ions (such as Fe3+, Zn2+, Ca+, and magnesium (Mg2+)) to form chelate complexes. Chelating agents are amino acids (glutamic acid & glycine) and organic acids (Ethylenediaminetetraacetic acid, Peracetic acid, Diethylene Triamine Pentaacetic Acid, Nitrilotriacetic acid, Hydroxyethylethylenediaminetriacetic acid, & Ethylenediamine) that scavenge metal ions from the soil [206,207]. Although not the primary function, the PS—MA pathway is capable of taking up other cations, although its affinity remains firmly targeted towards Fe [36,200]. Secondly, plants can absorb Zn in ionic form (Zn2+—stable due to a complete outer shell) with the use of ion-transporter channels at the expense of ATP (active transport) [205]. As aforementioned, chelating compounds can be produced and excreted by some endophytes, and these exudates can aid in accumulation via the lower energy pathway of complex formation [36]. There is, however, the potential for hyperaccumulation as a result of endophyte presence as a possibility and excess zinc is toxic to plants. Where Zn accumulation occurs outside of cellular regulation (EAP), plants can be unable to ‘turn off’ accumulation, and endophyte excreted PS could continue to accumulate Zn resulting in the toxification of the plant. It is unknown whether appropriate regulators are involved in the PS accumulation pathway [204,208].
P. indica has proven to show improvements in Zn accumulation in wheat grown under Zn deficient conditions, increasing overall vegetative weight, and branching of the roots. An increase in POD levels was observed in EAP and under oxidative stress it can be theorised that EAP would tolerate ROS better than non-EAP as a result of adequate Zn [183,209]. Highlighting the overall complexity of plant nutrition and the interwoven nature of plant systems, tying many metabolic functions together due to common factors such as Zn [209]. The aforementioned literature suggests that endophytes can significantly improve Zn uptake leading to improved plant growth.

5.3.3. Nitrogen

Nitrogen (N) is a key nutrient to plants as it is a major component of chlorophyll and amino acids, and in turn impacts the energy production and protein functions of cells [196]. The presence of endophytes has shown to significantly alter the chemical processes of basic plant survival mechanisms, outlined by a 2007 study that analyses the propensity of Heteroconium chaetospira to provide Chinese cabbage plants with nitrogen that plants are typically unable to access [68]. Pinus contorta has been observed to thrive in extremely low N conditions, alerting Padda et al. [210] to a potential alternate nitrogen source, with the likely cause being symbionts. 14 endophytes in this study were assessed as inoculants of P. contorta and were found to significantly increase seedling size and biomass under N-limited conditions. Most importantly, 11 endophytes increased free nitrogen in the plant by >200% and it was concluded that EB are able to fix N from the atmosphere to help P. contorta thrive in N-limited soil. High levels of nitrogenase were reported in endophytes isolated from P. contorta, an enzyme that classifies EB as nitrogen-fixing and explains the mode of action of these results [210,211]. There are many studies that have identified endophyte association improving N bioavailability through N13 radioisotope tracing from atmospheric N [68].
The significance of N to overall plant health cannot be understated as it plays a role in almost all cell functions through proteins [196]. Improving N availability through endophyte presence or fertilisation has a large flow-on affect to overall plant health. One example demonstrated in Figure 5 is how N availability can apply to drought resistance by metabolising ROS with proline, and in turn reducing accumulation and allowing IAA synthesis to continue uninhibited. [163,183,190,196]. Table 3 (page 19) summaries the impacts of plant nutrition on overall plant health as these tie closely to the theory of the vast downstream effects outlined in Figure 5.
N fertilizers, such as urea, are common practice to improve crop yields, as they are the most effective crop health improvers. However, urea can be harmful to the environment and livestock if inappropriately treated [212]. Using alternate sources to improve N bioavailability, such as EAP, greatly reduces the threat to the environment and the effort required to sustain healthy crops [213,214].

5.3.4. Phosphorus

Crop production is commonly limited by phosphorus (P) bioavailability due to phosphate formation, and the complex fixations with Fe, Ca, and aluminium (Al) [67]. A study conducted by Varga et al. [215] concluded that endophyte association (WP5 and WP42) can increase P solubilisation from Ca3(PO4)2 and AlPO4 in Poplus trees 80 and three-fold, respectively. The proteomics of plants with versus without sufficient P showed a general increase in metabolic related proteins, particularly those involved in sugar metabolism, of which P is an important co-factor. The increased metabolism rates are likely responsible for the increased dry weight of plants observed in the study. Multitudes of studies similar to Varga et al. [215] are required to understand the overall trends throughout plant endophyte relations, taking ‘omics’ approaches aid in providing context to the physiological impacts intracellularly.
Unlike many other essential plant nutrients, the role of P in plants—particularly agricultural plants—has been heavily studied as it is one of the greatest limiting factors in crop production, as fertilizers like the aforementioned urea are currently used to improve P bioavailability [67,213]. P fertilisers are of low efficiency due to P fixation and pose a great risk to the environment as they can contain toxic elements such as lead (Pb), mercury (Hg), and arsenic (As) [216,217]. Endophytes possess the potential to remove harmful contaminants from bioaccumulating in crops or livestock through reducing current fertilization practices. The large amount of agricultural land is saturated with P from previous fertilization, but remains fixed to Fe, Ca, and Al, as EAP inoculated with the right strain can access this soil-bound P and reduce the need for fertilization [67].
Endophyte association increases P solubilisation by increased gluconic, malic, citric, salicylic, and benzene acetic acid production and phosphatase activity [218,219]. Kumar et al. [220] investigated the impacts of P. indica presence under P limited conditions which resulted in a 150% and 20% biomass increase relative to P-limited and non-P-limited conditions, respectively. Similarly, studies indicate that endophytes can mitigate P-limiting conditions, with both control and P-limited plants resulting in a similar biomass and growth parameters [221,222]. Suggested effects of EAP are an increased production of organic acids in root associated microorganisms, as acidification of root environments aid the solubilisation of fixed P in soil [218,223]. Endophytes provide a strong alternative to crop acquisition of P which is typically delivered in a particularly problematic method.

5.3.5. Potassium

Potassium (K) is one of the most common elements in the earth’s crust, however, bioavailable K in soil is relatively low. The large majority (90–98%) of soil K is unavailable due to fixation, as K plays a key role in soil structure [224]. K+ is a vital cation involved in a number of cellular processes, including but not limited to metabolic regulation, photosynthesis, stress signalling, and maintaining redox homeostasis [224]. Extracellular excretion of MA involves symport membrane proteins involving K+ in the exchange [225,226], which ties K deficiency to an inability to obtain Fe via strategy II (the lower energy required pathway). K is also a key co-factor in carbohydrate metabolism, as it is an enzyme activator of carbohydrate metabolising enzymes impacting the general mass, growth, and reproduction of the plant [227,228].
Due to limited amounts of bioavailable K, attaining the maximum potential from the soil is crucial in maximising yields. Endophyte Burkholderia sp. significantly increased K availability from a range of fixed sources. To the degree that additional K supplementation did not have a statistically significant impact on the bioavailable K, as endophytes were able to provide sufficient levels of K [219]. However, it is possible that EAP cannot make up for soil K shortcomings and may need to be used in conjunction with current fertilisation practices to improve crop management procedures. Even if endophytes reduce the net requirements for fertilisation, it can be seen as a stabilisation of agricultural land, removing a proportion of the threat from the environment. Isolated endophyte studies have also identified the ability of endophytes to solubilise K in media, clearly indicating the quantitative measure to which endophytes can improve nutrient solubilisation. Rana et al. [86] demonstrates that endophytes identified as K solubilising—Pseudomonas brenneri, Ewingella americana, and Pantoea agglomerans—showed to improve root length, Fe concentration, protein, and chlorophyll content in-planta. This provides solutions to a number of deficiencies and plant short comings, proving the importance of adequate K in plants.

5.3.6. Sulphur

Sulphur availability is also an important factor in plant growth, as it is involved in the production of proteins that can impact the physiology and structure of plants via important amino acids cysteine and methionine (Figure 6) [25,94,229]. Sufficient levels of cysteine and methionine provide sufficient building blocks for polyamines and POD which are the antioxidants involved in the detoxification of ROS [101]. Identifying a link between endophyte impact on the transcriptome and tolerance to salinity stress is a key factor in agronomical crop failures [186].
A 2021 study of EB in sulphur limited conditions observed a significant abundance shift in bacteria of the Burkholderia genus under sulphur limited conditions [94]. A more in-depth study of Enterobacter sp. SA187 in Arabidopsis identified the mechanisms by which the particular EB interacts with the host plant to improve sulphate tolerance [101]. The study concluded that the treatment of sulphur-deprived plants with SA187 showed improvements in sulphur metabolism by the same degree as treatment directly with sulphur supplementation, by impacting the ethylene signalling pathway.
Improved sulphur metabolism as a result of endophyte presence is due to transcriptional regulation impacts of EB, which expand potential environmental sulphur sources to include sulphate and alkane sulfonates [94]. Transcriptome analysis of Arabidopsis identified 873 up-regulated and 610 down-regulated genes as a result of SA187 presence, outlining the significant and intricate impacts that endophytes have on plants, due to the volume of genes impacted, it is likely that SA187 benefits plants in other micronutrient deficiencies [101]. Although there is significant evidence to suggest that these transcriptome changes are directly tied to the benefits, radiolabelled tracing of sulphur would help quantify the direct physiological impact on the plant.

5.3.7. Calcium

P. indica shows great potential to improve multiple nutrient acquisitions in plants, one of which is Ca [122]. The determinant factor of cell wall integrity and rigidity, and key growth regulator, Ca is important for the fundamental structures of healthy plants. Ca2+ is also involved in secondary messaging, including a broad spectrum of functions, such as biotic and abiotic stress responses. Being one of the first signals in the pathogen defence cascade of the SAR [120,121,122]. P. indica [122], Penicillium bilaii, Penicillium oxalicum, Aspergillus niger [230], Ophiosphaerella sp., Cochliobolus sp., and Setosphaeria rostrata [46] all show the capacity to improve Ca solubilisation in-planta. Ca acquisition correlates heavily with P and Mg solubilisation due to the common source of Ca originating from calcium phosphate and dolomite [46,231]. Ca is taken into roots through permeable ion channels via active transport, and endophytes improve intake by stimulating this process. Endophytes also improve nutrient uptake by increasing the capacity of accumulation through root growth stimulation [232]. The improved solubilisation of Ca by endophytes is not completely understood and requires further research collectively with P solubilising EB and EF to understand the full picture [230].

5.3.8. Magnesium

Mg is involved in growth regulation, cell wall structure, chlorophyl concentration, and photosynthesis rates as an enzymatic co-factor [26,233]. Furthering this point, evidence suggests sufficient Mg levels are correlated to a number of active sites of enzymes responsible for the acquisition and transport of other nutrients [234]. Mg helps biotic tolerances by improving the structure of cell walls, moreover, Mg2+ has been described to act competitively for binding sites of pathogens [235,236]. Similar to other cations such as Fe, Zn, and Ca, Mg is taken into plants by one of two pathways, active transport, or chelation (strategy II) [36,206].
Endophytes P. indica and Lolium perenne show improved Mg acquisition flowing on to impacting general growth and activity of plants [103,233,237]. Prasad et al. [237] details that P. indica has a positive relationship with Arabidopsis thaliana chlorophyll concentration both under Mg deficient and sufficient conditions. This uggests that endophytes are capable of overcoming nutrient deficiencies, whcih is due to the aforementioned increases in chelating compounds produced by endophytes [36,206]. Moreover, utilisation of Mg is higher in EAP resulting in higher levels in the plant tissues [103,206].
Theories and evidence suggest that endophytic improvement of Mg, and all nutrients aforementioned, presents a flexible application of EAP. Whereby endophytes can provide stress tolerances in conditions of low nutrition, as well as providing improved nutrient acquisition, transportation, and utilisation under optimal conditions, improving gross yield and biomass [233,237].
Table 3. Summary table of effects of endophytes on key plant nutrients.
Table 3. Summary table of effects of endophytes on key plant nutrients.
NutrientMechanism of Endophyte InterventionEffect on PlantReference
IronPhytosiderophore production—reducing the energy requirements to transport Fe into roots.Reduces ROS accumulation.
Key roles in photosynthesis and the electron transport chain.
[28,36,197,202,203,218]
ZincPhytosiderophore production—as above.Enzymatic co-factor in root growth.
Component of chlorophyll.
[28,36,204,205,209]
NitrogenAtmospheric fixation when soil N is low (mainly bacteria due to fungi lacking nitrogenase).Key factor in amino compounds.
Increases proline concentration.
[67,71,82,95,186,210,238]
PhosphorusOrganic acid production & soil acidification to increase solubilisation, and increased proteins involved in sugar metabolism.Root growth stimulation.
Key component in DNA and RNA.
[67,215,216,218,219,222]
PotassiumImproving flow of free K+ into roots. However, remains limited by available soil K.Metabolic regulation, photosynthesis, stress signalling, and maintaining redox homeostasis.[25,219,224,225]
SulphurProtein production to utilise sulphate and alkane sulfonates.Detoxification of ROS.
Key structural component of proteins.
[94,101,229]
CalciumTied to P and Mg solubilisation due to complex formationMaintains redox homeostasis.
Key structural component of cell walls.
[53,120,148,230,231]
MagnesiumNot determinedKey co-factor to enzymes involved in a diverse range of metabolic functions.[234,235,237]

6. Large-Scale Agricultural Application

The continuation of endophytic studies is required to analyse the stability of inoculated plants in a large scale/commercial setting beyond current applications. Endophytes have been successfully incorporated into commercialised perennial rye grasses on several occasions, improving the performances of grass growth on an annual scale [61]. The knowledge of currently available endophyte-treated products is protected behind patents and intellectual property rights, making it difficult to critically analyse the effects of endophyte presence [111]. Viable results stemming from EAP in grasses is highly relevant to providing validity to proposing endophyte-based applications to agroeconomical crops, maximising yield whilst maintaining surveillance of adverse effects on the livestock consuming such products [239]. At the time of writing, there is no reported negative side effects from EAP consumption within animal trials, yet the costs of endophyte inoculation to plants and humans are yet to be studied and described appropriately. It is vitally important that all aspects of EAP are understood to consider them as safe treatments for consumption by the general public.
Applying the theory of commercial rye grasses to hypothesise the successful incorporation of endophytes into human consumed crops, many aspects of food supply can be improved. Endophytes that are capable of increasing biomass through a range of mechanisms aid in the total yield of produce from crops, improving supply security. Furthered by abiotic stress tolerance increases allowing for greater yields in sub-optimal conditions and opening new land for agricultural use. Additionally, biosecurity is increased through increased biotic stress tolerances in EAP, reducing crop loss to pests and pathogens. EAP can also be an important strategy to combat malnourishment globally, improving nutrient levels within widely consumed crops (rice, wheat, cereals) as a form of dietary supplementation. Globally, 742 million tonnes of rice are produced per year, a marginal increase in yield efficiency can result in an extraordinary change in our abilities to provide for malnourished populations. Endophytes possess the ability to increase gross yield through increasing efficiency per hectare and increasing total farmable land.

7. Conclusions

In conclusion, this review paper has highlighted the significant role that endophytes can play in promoting plant growth and improving nutrient acquisition in agriculture. The interaction between endophytes and their host plants has been shown to result in enhanced nutrient uptake, stress tolerance, and disease resistance, ultimately leading to improved crop yields. Furthermore, endophytes offer a promising and eco-friendly alternative to traditional agricultural practices, reducing the need for synthetic fertilizers and pesticides. Research in this area is ongoing, with numerous endophytes being identified and characterised every year in an attempt to develop practical applications for their use in agriculture. As such, the potential of endophytes as a sustainable solution for improving crop productivity and reducing environmental impacts cannot be understated.
Despite the significant progress made in understanding the role of endophytes in agriculture, there is still much to be discovered. Continued research is needed to gain a deeper understanding of the interactions between endophytes and plants, as well as the factors that affect these interactions. This will involve identifying and characterizing new endophytes, studying their modes of action, and understanding their effects on plant growth, and nutrient uptake under different environmental conditions. Additional research is needed to develop practical applications for the use of endophytes in agriculture. This will involve testing different endophyte inoculation methods, optimizing their efficacy, analysing human impacts, and developing strategies to scale-up their use in farming systems. It is also important to assess the potential risks associated with the use of endophytes, such as the spread of harmful microbes or impacts on soil microbiota.
Due to its novelty future, endophyte studies can take on a range of scopes, whether it be the development of stress tolerant plants, bio inoculums, natural antimicrobials/pesticides, or understanding the fundamental impacts of stress on, plants, the endomicrobiome, and their relationship. There are many contributions still required to realise the potential of endophytes in agriculture. Any and all work conducted on endophytes helps fill in the gaps in the knowledge of endophytes and their properties to one day have enough data and evidence to make a broader and clearer analysis of endophyte-based approaches to agriculture.
Overall, the results presented in this review paper demonstrate the immense potential of endophytes in agriculture and highlight the need for further research and development in this field. By harnessing the benefits of endophytes, we can move towards a more sustainable and efficient agricultural system, promoting both environmental and economic sustainability, all the while, fortifying crop plants against stressors and reducing the need for chemical treatments of the land.

Author Contributions

Conceptualization, D.W., E.A.P., A.J.C. and B.Z.; investigation, D.W.; writing—original draft preparation, D.W.; writing—review and editing, D.W.; visualization, D.W.; supervision, E.A.P., A.J.C. and B.Z.; project administration, D.W. and B.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Deshmukh, S.; Verekar, S.; Bhave, S. Endophytic fungi: A reservoir of antibacterials. Front. Microbiol. 2015, 5, 715. [Google Scholar] [CrossRef] [PubMed]
  2. Strobel, G.; Daisy, B. Bioprospecting for microbial endophytes and their natural products. MMBR 2003, 67, 491–502. [Google Scholar] [CrossRef] [PubMed]
  3. Yadav, G.; Meena, M. Bioprospecting of endophytes in medicinal plants of Thar Desert: An attractive resource for biopharmaceuticals. Biotechnol. Rep. 2021, 30, e00629. [Google Scholar] [CrossRef]
  4. Manganyi, M.C.; Ateba, C.N. Untapped potentials of endophytic fungi: A review of novel bioactive compounds with biological applications. Microorganisms 2020, 8, 1934. [Google Scholar] [CrossRef]
  5. Schulz, B.; Boyle, C. The endophytic continuum. Mycol. Res. 2005, 109, 661–686. [Google Scholar] [CrossRef]
  6. Baron, N.C.; Rigobelo, E.C. Endophytic fungi: A tool for plant growth promotion and sustainable agriculture. Mycology 2022, 13, 39–55. [Google Scholar] [CrossRef] [PubMed]
  7. Singh, O.V. Food Borne Pathogens and Antibiotic Resistance; John Wiley & Sons: Hoboken, NJ, USA, 2017. [Google Scholar]
  8. AURA. Fourth Australian Report on Antimicrobial Use and Resistance in Human Health; AURA: Sydney, NSW, Australia, 2021; ISBN 978-1-922563-21-7. [Google Scholar]
  9. Quinn, G.A.; Banat, A.M.; Abdelhameed, A.M.; Banat, I.M. Streptomyces from traditional medicine: Sources of new innovations in antibiotic discovery. J. Med. Microbiol. 2020, 69, 1040–1048. [Google Scholar] [CrossRef]
  10. United Nations, Department of Economic and Social Affairs, Population Division. World Population Prospects 2022: Summary of Results; United Nations: New York, NY, USA, 2022. [Google Scholar]
  11. Putri, R.; Naufal, M.; Nandini, M.; Dwiputra, D.; Wibirama, S.; Sumantyo, J. The Impact of Population Pressure on Agricultural Land towards Food Sufficiency (Case in West Kalimantan Province, Indonesia). In Proceedings of the IOP Conference Series: Earth and Environmental Science, Yogyakarta, Indonesia, 22–23 October 2019; p. 012050. [Google Scholar]
  12. Anggraini, D.P.; Sutrisno, J.; Barokah, U. The impact analysis of rice field conversion on food security in Sragen Regency. Proc. E3S Web Conf. 2021, 316, 02035. [Google Scholar] [CrossRef]
  13. Patel, S.K.; Sharma, A.; Singh, G.S. Traditional agricultural practices in India: An approach for environmental sustainability and food security. Energy Ecol. Environ. 2020, 5, 253–271. [Google Scholar] [CrossRef]
  14. FAO; IFAD; UNICEF; WFP; WHO. The State of Food Security and Nutrition in the World 2021; FAO: Rome, Italy, 2021. [Google Scholar]
  15. Kraemer, K.; Gilbert, C. Do vitamin A deficiency and undernutrition still matter? Community Eye Health 2013, 26, 61–63. [Google Scholar]
  16. Roser, M.; Ritchie, H. Hunger and Undernourishment. Available online: https://ourworldindata.org/hunger-and-undernourishment (accessed on 24 October 2022).
  17. Savary, S.; Willocquet, L. Simulation modeling in botanical epidemiology and crop loss analysis. Plant Health Instr. 2014, 147. [Google Scholar]
  18. Cerda, R.; Avelino, J.; Gary, C.; Tixier, P.; Lechevallier, E.; Allinne, C. Primary and Secondary Yield Losses Caused by Pests and Diseases: Assessment and Modeling in Coffee. PLoS ONE 2017, 12, e0169133. [Google Scholar] [CrossRef]
  19. Smith, H.M.; Samach, A. Constraints to obtaining consistent annual yields in perennial tree crops. I: Heavy fruit load dominates over vegetative growth. Plant Sci. 2013, 207, 158–167. [Google Scholar] [CrossRef] [PubMed]
  20. Blair, A.; Ritz, B.; Wesseling, C.; Beane Freeman, L. Pesticides and human health. Occup. Environ. Med. 2015, 72, 81. [Google Scholar] [CrossRef]
  21. Gill, S.S.; Gill, R.; Trivedi, D.K.; Anjum, N.A.; Sharma, K.K.; Ansari, M.W.; Ansari, A.A.; Johri, A.K.; Prasad, R.; Pereira, E.; et al. Piriformospora indica: Potential and Significance in Plant Stress Tolerance. Front. Microbiol. 2016, 7, 332. [Google Scholar] [CrossRef] [PubMed]
  22. 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]
  23. Pammi, M.; Aghaeepour, N.; Neu, J. Multiomics, artificial intelligence, and precision medicine in perinatology. Pediatr. Res. 2022, 93, 308–315. [Google Scholar] [CrossRef] [PubMed]
  24. Govindaraj, M.; Masilamani, P.; Albert, V.A.; Bhaskaran, M. Effect of physical seed treatment on yield and quality of crops: A review. Agric. Rev. 2017, 38, 1–14. [Google Scholar] [CrossRef]
  25. Hoffmann, C.; Stockfisch, N.; Koch, H.-J. Influence of sulphur supply on yield and quality of sugar beet (Beta vulgaris L.)—Determination of a threshold value. Eur. J. Agron. 2004, 21, 69–80. [Google Scholar] [CrossRef]
  26. Sheng, M.; Tang, M.; Chen, H.; Yang, B.; Zhang, F.; Huang, Y. Influence of arbuscular mycorrhizae on photosynthesis and water status of maize plants under salt stress. Mycorrhiza 2008, 18, 287–296. [Google Scholar] [CrossRef]
  27. Su, Z.-z.; Wang, T.; Shrivastava, N.; Chen, Y.-y.; Liu, X.; Sun, C.; Yin, Y.; Gao, Q.-k.; Lou, B.-g. Piriformospora indica promotes growth, seed yield and quality of Brassica napus L. Microbiol. Res. 2017, 199, 29–39. [Google Scholar] [CrossRef] [PubMed]
  28. Singh, D.; Geat, N.; Rajawat, M.V.S.; Prasanna, R.; Kar, A.; Singh, A.M.; Saxena, A.K. Prospecting endophytes from different Fe or Zn accumulating wheat genotypes for their influence as inoculants on plant growth, yield, and micronutrient content. Ann. Microbiol. 2018, 68, 815–833. [Google Scholar] [CrossRef]
  29. Mei, C.; Flinn, B.S. The use of beneficial microbial endophytes for plant biomass and stress tolerance improvement. Recent Pat. Biotechnol. 2010, 4, 81–95. [Google Scholar] [CrossRef] [PubMed]
  30. Khaksar, G.; Treesubsuntorn, C.; Thiravetyan, P. Effect of endophytic Bacillus cereus ERBP inoculation into non-native host: Potentials and challenges for airborne formaldehyde removal. Plant Physiol. Biochem. 2016, 107, 326–336. [Google Scholar] [CrossRef]
  31. Gaiero, J.R.; McCall, C.A.; Thompson, K.A.; Day, N.J.; Best, A.S.; Dunfield, K.E. Inside the root microbiome: Bacterial root endophytes and plant growth promotion. Am. J. Bot. 2013, 100, 1738–1750. [Google Scholar] [CrossRef]
  32. Nelson, E.B. Microbial dynamics and interactions in the spermosphere. Annu. Rev. Phytopathol. 2004, 42, 271–309. [Google Scholar] [CrossRef]
  33. Robinson, R.J.; Fraaije, B.A.; Clark, I.M.; Jackson, R.W.; Hirsch, P.R.; Mauchline, T.H. Wheat seed embryo excision enables the creation of axenic seedlings and Koch’s postulates testing of putative bacterial endophytes. Sci. Rep. 2016, 6, 25581. [Google Scholar] [CrossRef]
  34. Wakeel, A.; Farooq, M.; Bashir, K.; Ozturk, L. Micronutrient Malnutrition and Biofortification: Recent Advances and Future Perspectives. In Plant Micronutrient Use Efficiency; Academic Press: Cambridge, MA, USA, 2018; pp. 225–243. [Google Scholar]
  35. Römheld, V. The role of phytosiderophores in acquisition of iron and other micronutrients in graminaceous species: An ecological approach. In Iron Nutrition and Interactions in Plants; Springer: Berlin/Heidelberg, Germany, 1991; pp. 159–166. [Google Scholar]
  36. Prasad, R.; Shivay, Y.S.; Mandi, S. Chapter 21—Phytosiderophores and absorption of iron and other cations by plants. In Cation Transporters in Plants; Upadhyay, S.K., Ed.; Academic Press: Cambridge, MA, USA, 2022; pp. 385–399. [Google Scholar]
  37. Christensen, M.J. Antifungal activity in grasses infected with Acremonium and Epichloe endophytes. Australas. Plant Pathol. 1996, 25, 186–191. [Google Scholar] [CrossRef]
  38. Hashem, A.H.; Shehabeldine, A.M.; Abdelaziz, A.M.; Amin, B.H.; Sharaf, M.H. Antifungal Activity of Endophytic Aspergillus terreus Extract Against Some Fungi Causing Mucormycosis: Ultrastructural Study. Appl. Biochem. Biotechnol. 2022, 194, 3468–3482. [Google Scholar] [CrossRef]
  39. Vassileva, M.; Malusá, E.; Eichler-Löbermann, B.; Vassilev, N. Aspegillus terreus: From Soil to Industry and Back. Microorganisms 2020, 8, 1655. [Google Scholar] [CrossRef]
  40. Dabiré, T.G.; Bonzi, S.; Somda, I.; Legrève, A. Evaluation of the potential of Trichoderma harzianum as a plant growth promoter and biocontrol agent against Fusarium damping-off in onion in Burkina Faso. Asian J. Plant Pathol. 2016, 10, 49–60. [Google Scholar] [CrossRef]
  41. Shah, S.; Ash, G.J.; Wilson, B.A. Resporulation of Metarhizium anisopliae granules on soil and mortality of Tenebrio molitor: Implications for wireworm management in sweetpotato. Ann. Appl. Biol. 2023, 182, 65–76. [Google Scholar] [CrossRef]
  42. Vinale, F.; Flematti, G.; Sivasithamparam, K.; Lorito, M.; Marra, R.; Skelton, B.W.; Ghisalberti, E.L. Harzianic acid, an antifungal and plant growth promoting metabolite from Trichoderma harzianum. J. Nat. Prod. 2009, 72, 2032–2035. [Google Scholar] [CrossRef] [PubMed]
  43. Xia, C.; Li, N.; Zhang, X.; Feng, Y.; Christensen, M.J.; Nan, Z. An Epichloe endophyte improves photosynthetic ability and dry matter production of its host Achnatherum inebrians infected by Blumeria graminis under various soil water conditions. Fungal Ecol. 2016, 22, 26–34. [Google Scholar] [CrossRef]
  44. Chen, Q.; Yu, J.-J.; He, J.; Feng, T.; Liu, J.-K. Isobenzofuranones and isocoumarins from kiwi endophytic fungus Paraphaeosphaeria sporulosa and their antibacterial activity against Pseudomonas syringae pv. actinidiae. Phytochemistry 2022, 195, 113050. [Google Scholar] [CrossRef]
  45. Xin, X.-F.; Kvitko, B.; He, S.Y. Pseudomonas syringae: What it takes to be a pathogen. Nat. Rev. Microbiol. 2018, 16, 316–328. [Google Scholar] [CrossRef]
  46. Spagnoletti, F.N.; Tobar, N.E.; Fernández Di Pardo, A.; Chiocchio, V.M.; Lavado, R.S. Dark septate endophytes present different potential to solubilize calcium, iron and aluminum phosphates. Appl. Soil Ecol. 2017, 111, 25–32. [Google Scholar] [CrossRef]
  47. Yin, Z.; Shi, F.; Jiang, H.; Roberts, D.P.; Chen, S.; Fan, B. Phosphate solubilization and promotion of maize growth by Penicillium oxalicum P4 and Aspergillus niger P85 in a calcareous soil. Can. J. Microbiol 2015, 61, 913–923. [Google Scholar] [CrossRef]
  48. Bömke, C.; Rojas, M.C.; Gong, F.; Hedden, P.; Tudzynski, B. Isolation and characterization of the gibberellin biosynthetic gene cluster in Sphaceloma manihoticola. Appl. Environ. Microbiol. 2008, 74, 5325–5339. [Google Scholar] [CrossRef]
  49. Choi, W.-Y.; Rim, S.-O.; Lee, J.-H.; Lee, J.-M.; Lee, I.-J.; Cho, K.-J.; Rhee, I.-K.; Kwon, J.-B.; Kim, J.-G. Isolation of gibberellins-producing fungi from the root of several Sesamum indicum plants. J. Microbiol. Biotechnol. 2005, 15, 22–28. [Google Scholar]
  50. Rademacher, W.; Graebe, J.E. Gibberellin A4 produced by Sphacelomamanihoticola, the cause of the superelongation disease of cassava (Manihotesculenta). Biochem. Biophys. Res. Commun. 1979, 91, 35–40. [Google Scholar] [CrossRef] [PubMed]
  51. Lucangeli, C.; Bottini, R. Effects of Azospirillum spp. on endogenous gibberellin content and growth of maize (Zea mays L.) treated with uniconazole. Symbiosis 1997, 23, 63–72. [Google Scholar]
  52. Khan, S.A.; Hamayun, M.; Yoon, H.; Kim, H.-Y.; Suh, S.-J.; Hwang, S.-K.; Kim, J.-M.; Lee, I.-J.; Choo, Y.-S.; Yoon, U.-H. Plant growth promotion and Penicillium citrinum. BMC Microbiol. 2008, 8, 231. [Google Scholar] [CrossRef] [PubMed]
  53. Khan, A.L.; Hamayun, M.; Kim, Y.-H.; Kang, S.-M.; Lee, I.-J. Ameliorative symbiosis of endophyte (Penicillium funiculosum LHL06) under salt stress elevated plant growth of Glycine max L. Plant Physiol. Biochem. 2011, 49, 852–861. [Google Scholar] [CrossRef]
  54. Abdelaziz, A.M.; El-Wakil, D.A.; Attia, M.S.; Ali, O.M.; AbdElgawad, H.; Hashem, A.H. Inhibition of Aspergillus flavus Growth and Aflatoxin Production in Zea mays L. Using Endophytic Aspergillus fumigatus. J. Fungi 2022, 8, 482. [Google Scholar] [CrossRef]
  55. Mantzoukas, S.; Lagogiannis, I.; Mpousia, D.; Ntoukas, A.; Karmakolia, K.; Eliopoulos, P.A.; Poulas, K. Beauveria bassiana Endophytic Strain as Plant Growth Promoter: The Case of the Grape Vine Vitis vinifera. J. Fungi 2021, 7, 142. [Google Scholar] [CrossRef]
  56. Saragih, M.; Trizelia; Nurbailis; Yusniwati. Endophytic Colonization and Plant Growth Promoting Effect by Entomopathogenic fungus, Beauveria bassiana to Red Chili (Capsicum annuum L.) with Different Inoculation Methods. IOP Conf. Ser. Earth Environ. Sci. 2019, 305, 012070. [Google Scholar] [CrossRef]
  57. Nakahara, Y.; Shimura, S.; Ueno, C.; Kanamori, Y.; Mita, K.; Kiuchi, M.; Kamimura, M. Purification and characterization of silkworm hemocytes by flow cytometry. Dev. Comp. Immunol. 2009, 33, 439–448. [Google Scholar] [CrossRef]
  58. Wang, X.; Radwan, M.M.; Taráwneh, A.H.; Gao, J.; Wedge, D.E.; Rosa, L.H.; Cutler, H.G.; Cutler, S.J. Antifungal activity against plant pathogens of metabolites from the endophytic fungus Cladosporium cladosporioides. J. Agric. Food Chem. 2013, 61, 4551–4555. [Google Scholar] [CrossRef]
  59. Hiruma, K.; Gerlach, N.; Sacristán, S.; Nakano, R.T.; Hacquard, S.; Kracher, B.; Neumann, U.; Ramírez, D.; Bucher, M.; O’Connell, R.J.; et al. Root Endophyte Colletotrichum tofieldiae Confers Plant Fitness Benefits that Are Phosphate Status Dependent. Cell 2016, 165, 464–474. [Google Scholar] [CrossRef]
  60. Huang, W.; Cai, Y.; Hyde, K.; Corke, H.; Sun, M. Biodiversity of endophytic fungi associated with 29 traditional Chinese medicinal plants. Fungal Divers. 2008, 33, 61–75. [Google Scholar]
  61. Barenbrug, A. Barenbrug Autumn 2022 Product Guide; Barenbrug Agriculture: Melbourne, Australia, 2022; pp. 1–12. [Google Scholar]
  62. Christensen, M.J.; Latch, G.C.M. Variation among isolates of Acremonium endophytes (A. coenophialum and possibly A. typhinum) from tall fescue (Festuca arundinacea). Mycol. Res. 1991, 95, 1123–1126. [Google Scholar] [CrossRef]
  63. Abdel-Motaal, F.; Kamel, N.; El-Zayat, S.; Abou-Ellail, M. Early blight suppression and plant growth promotion potential of the endophyte Aspergillus flavus in tomato plant. Ann. Agric. Sci. 2020, 65, 117–123. [Google Scholar] [CrossRef]
  64. Oelmüller, R.; Sherameti, I.; Tripathi, S.; Varma, A. Piriformospora indica, a cultivable root endophyte with multiple biotechnological applications. Symbiosis 2009, 49, 1–17. [Google Scholar] [CrossRef]
  65. Sirrenberg, A.; Göbel, C.; Grond, S.; Czempinski, N.; Ratzinger, A.; Karlovsky, P.; Santos, P.; Feussner, I.; Pawlowski, K. Piriformospora indica affects plant growth by auxin production. Physiol. Plant. 2007, 131, 581–589. [Google Scholar] [CrossRef]
  66. Ren, C.-G.; Dai, C.-C. Jasmonic acid is involved in the signaling pathway for fungal endophyte-induced volatile oil accumulation of Atractylodes lancea plantlets. BMC Plant Biol. 2012, 12, 128. [Google Scholar] [CrossRef]
  67. Tang, M.-J.; Zhu, Q.; Zhang, F.-M.; Zhang, W.; Yuan, J.; Sun, K.; Xu, F.-J.; Dai, C.-C. Enhanced nitrogen and phosphorus activation with an optimized bacterial community by endophytic fungus Phomopsis liquidambari in paddy soil. Microbiol. Res. 2019, 221, 50–59. [Google Scholar] [CrossRef]
  68. Usuki, F.; Narisawa, K. A mutualistic symbiosis between a dark septate endophytic fungus, Heteroconium chaetospira, and a nonmycorrhizal plant, Chinese cabbage. Mycologia 2007, 99, 175–184. [Google Scholar] [CrossRef]
  69. Abdelgawad, M.A.; Hamed, A.A.; Nayl, A.A.; Badawy, M.; Ghoneim, M.M.; Sayed, A.M.; Hassan, H.M.; Gamaleldin, N.M. The Chemical Profiling, Docking Study, and Antimicrobial and Antibiofilm Activities of the Endophytic fungi Aspergillus sp. AP5. Molecules 2022, 27, 1704. [Google Scholar] [CrossRef]
  70. Chen, M.; Yang, G.; Sheng, Y.; Li, P.; Qiu, H.; Zhou, X.; Huang, L.; Chao, Z. Glomus mosseae Inoculation Improves the Root System Architecture, Photosynthetic Efficiency and Flavonoids Accumulation of Liquorice under Nutrient Stress. Front. Plant Sci. 2017, 8, 931. [Google Scholar] [CrossRef]
  71. Mengistu, A.A. Endophytes: Colonization, Behaviour, and Their Role in Defense Mechanism. Int. J. Microbiol. 2020, 2020, 6927219. [Google Scholar] [CrossRef] [PubMed]
  72. Taylor, L.; Gutierrez, S.; McCormick, S.P.; Bakker, M.G.; Proctor, R.H.; Teresi, J.; Kurtzman, B.; Hao, G.; Vaughan, M.M. Use of the volatile trichodiene to reduce Fusarium head blight and trichothecene contamination in wheat. Microb. Biotechnol. 2022, 15, 513–527. [Google Scholar] [CrossRef] [PubMed]
  73. Haddad, P.E.; Leite, L.G.; Lucon, C.M.M.; Harakava, R. Selection of Trichoderma spp. strains for the control of Sclerotinia sclerotiorum in soybean. Pesqui. Agropecu. Bras. 2017, 52, 1140–1148. [Google Scholar] [CrossRef]
  74. Islam, T.; Gupta, D.R.; Surovy, M.Z.; Mahmud, N.U.; Mazlan, N.; Islam, T. Identification and application of a fungal biocontrol agent Cladosporium cladosporioides against Bemisia tabaci. Biotechnol. Bioteq. 2019, 33, 1698–1705. [Google Scholar] [CrossRef]
  75. Popay, A.J.; Jensen, J.G.; Mace, W.J. Root Herbivory: Grass Species, Epichloe Endophytes and Moisture Status Make a Difference. Microorganisms 2020, 8, 997. [Google Scholar] [CrossRef]
  76. Waweru, B.; Turoop, L.; Kahangi, E.; Coyne, D.; Dubois, T. Non-pathogenic Fusarium oxysporum endophytes provide field control of nematodes, improving yield of banana (Musa sp.). Biol. Control 2014, 74, 82–88. [Google Scholar] [CrossRef]
  77. Lopes, R.; Tsui, S.; Gonçalves, P.J.R.O.; De Queiroz, M.V. A look into a multifunctional toolbox: Endophytic Bacillus species provide broad and underexploited benefits for plants. World J. Microbiol. Biotechnol. 2018, 34, 94. [Google Scholar] [CrossRef]
  78. Kou, M.-Z.; Bastías, D.A.; Christensen, M.J.; Zhong, R.; Nan, Z.-B.; Zhang, X.-X. The Plant Salicylic Acid Signalling Pathway Regulates the Infection of a Biotrophic Pathogen in Grasses Associated with an Epichloe Endophyte. J. Fungi 2021, 7, 633. [Google Scholar] [CrossRef]
  79. Chakraborty, M.; Mahmud, N.U.; Gupta, D.R.; Tareq, F.S.; Shin, H.J.; Islam, T. Inhibitory Effects of Linear Lipopeptides From a Marine Bacillus subtilis on the Wheat Blast Fungus Magnaporthe oryzae Triticum. Front. Microbiol. 2020, 11, 665. [Google Scholar] [CrossRef]
  80. Bakhtiyarifar, M.; Enayatizamir, N.; Mehdi Khanlou, K. Biochemical and molecular investigation of non-rhizobial endophytic bacteria as potential biofertilisers. Arch. Microbiol. 2021, 203, 513–521. [Google Scholar] [CrossRef]
  81. Nonthakaew, N.; Panbangred, W.; Songnuan, W.; Intra, B. Plant growth-promoting properties of Streptomyces spp. isolates and their impact on mung bean plantlets’ rhizosphere microbiome. Front. Microbiol. 2022, 13, 967415. [Google Scholar] [CrossRef] [PubMed]
  82. Naher, K.; Miwa, H.; Okazaki, S.; Yasuda, M. Effects of Different Sources of Nitrogen on Endophytic Colonization of Rice Plants by Azospirillum sp. B510. Microbes Environ. 2018, 33, 301–308. [Google Scholar] [CrossRef]
  83. Molina, R.; Rivera, D.; Mora, V.; López, G.; Rosas, S.; Spaepen, S.; Vanderleyden, J.; Cassán, F. Regulation of IAA Biosynthesis in Azospirillum brasilense Under Environmental Stress Conditions. Curr. Microbiol. 2018, 75, 1408–1418. [Google Scholar] [CrossRef] [PubMed]
  84. Ahmad, M.; Naseer, I.; Hussain, A.; Zahid Mumtaz, M.; Mustafa, A.; Hilger, T.H.; Ahmad Zahir, Z.; Minggang, X. Appraising Endophyte–Plant Symbiosis for Improved Growth, Nodulation, Nitrogen Fixation and Abiotic Stress Tolerance: An Experimental Investigation with Chickpea (Cicer arietinum L.). Agronomy 2019, 9, 621. [Google Scholar] [CrossRef]
  85. Xiao, Y.; Wu, K. Recent progress on the interaction between insects and Bacillus thuringiensis crops. Philos. Trans. R. Soc. B 2019, 374, 20180316. [Google Scholar] [CrossRef] [PubMed]
  86. Rana, K.L.; Kour, D.; Kaur, T.; Devi, R.; Yadav, A.; Yadav, A.N. Bioprospecting of endophytic bacteria from the Indian Himalayas and their role in plant growth promotion of maize (Zea mays L.). J. Appl. Biol. Biotechnol. 2021, 9, 41–50. [Google Scholar]
  87. Shahid, M.; Khan, M.S.; Syed, A.; Marraiki, N.; Elgorban, A.M. Mesorhizobium ciceri as biological tool for improving physiological, biochemical and antioxidant state of Cicer aritienum (L.) under fungicide stress. Sci. Rep. 2021, 11, 9655. [Google Scholar] [CrossRef]
  88. Vijayabharathi, R.; Gopalakrishnan, S.; Sathya, A.; Srinivas, V.; Sharma, M. Deciphering the tri-dimensional effect of endophytic Streptomyces sp. on chickpea for plant growth promotion, helper effect with Mesorhizobium ciceri and host-plant resistance induction against Botrytis cinerea. Microb. Pathog. 2018, 122, 98–107. [Google Scholar] [CrossRef]
  89. Cassán, F.; Diaz-Zorita, M. Azospirillum sp. in current agriculture: From the laboratory to the field. Soil Biol. Biochem. 2016, 103, 117–130. [Google Scholar] [CrossRef]
  90. Perrig, D.; Boiero, M.L.; Masciarelli, O.A.; Penna, C.; Ruiz, O.A.; Cassán, F.D.; Luna, M.V. Plant-growth-promoting compounds produced by two agronomically important strains of Azospirillum brasilense, and implications for inoculant formulation. Appl. Microbiol. Biot. 2007, 75, 1143–1150. [Google Scholar] [CrossRef]
  91. Sun, M.; Ye, S.; Xu, Z.; Wan, L.; Zhao, Y. Endophytic Bacillus altitudinis Q7 from Ginkgo biloba inhibits the growth of Alternaria alternata in vitro and its inhibition mode of action. Biotechnol. Biotechnol. Equip. 2021, 35, 880–894. [Google Scholar] [CrossRef]
  92. Bent, E.; Chanway, C.P. The growth-promoting effects of a bacterial endophyte on lodgepole pine are partially inhibited by the presence of other rhizobacteria. Can. J. Microbiol 1998, 44, 980–988. [Google Scholar] [CrossRef]
  93. Kshetri, L.; Naseem, F.; Pandey, P. Role of Serratia sp. as biocontrol agent and plant growth stimulator, with prospects of biotic stress management in plant. In Plant Growth Promoting Rhizobacteria for Sustainable Stress Management: Volume 2: Rhizobacteria in Biotic Stress Management; Springer: Singapore, 2019; pp. 169–200. [Google Scholar]
  94. Bertoldo, G.; Della Lucia, M.C.; Squartini, A.; Concheri, G.; Broccanello, C.; Romano, A.; Ravi, S.; Cagnin, M.; Baglieri, A.; Stevanato, P. Endophytic Microbiome Responses to Sulfur Availability in Beta vulgaris (L.). Int. J. Mol. Sci. 2021, 22, 7184. [Google Scholar] [CrossRef]
  95. Kong, P.; Hong, C. Endophytic Burkholderia sp. SSG as a potential biofertilizer promoting boxwood growth. PeerJ 2020, 8, e9547. [Google Scholar] [CrossRef]
  96. Kumar, U.; Kaviraj, M.; Rout, S.; Chakraborty, K.; Swain, P.; Nayak, P.K.; Nayak, A.K. Combined application of ascorbic acid and endophytic N-fixing Azotobacter chroococcum Avi2 modulates photosynthetic efficacy, antioxidants and growth-promotion in rice under moisture deficit stress. Microbiol. Res. 2021, 250, 126808. [Google Scholar] [CrossRef] [PubMed]
  97. Devi, K.A.; Pandey, P.; Sharma, G.D. Plant Growth-Promoting Endophyte Serratia marcescens AL2-16 Enhances the Growth of Achyranthes aspera L., a Medicinal Plant. HAYATI J. Biosci. 2016, 23, 173–180. [Google Scholar] [CrossRef]
  98. David, B.V.; Chandrasehar, G.; Selvam, P.N. Pseudomonas fluorescens: A plant-growth-promoting rhizobacterium (PGPR) with potential role in biocontrol of pests of crops. In Crop Improvement through Microbial Biotechnology; Elsevier: Amsterdam, The Netherlands, 2018; pp. 221–243. [Google Scholar]
  99. Oni, F.E.; Kieu Phuong, N.; Höfte, M. Recent advances in Pseudomonas biocontrol. In Bacteria-Plant Interactions: Advanced Research and Future Trends; Caister Academic Press: Poole, UK, 2015; pp. 167–198. [Google Scholar]
  100. Lally, R.D.; Galbally, P.; Moreira, A.S.; Spink, J.; Ryan, D.; Germaine, K.J.; Dowling, D.N. Application of Endophytic Pseudomonas fluorescens and a Bacterial Consortium to Brassica napus Can Increase Plant Height and Biomass under Greenhouse and Field Conditions. Front. Plant Sci. 2017, 8, 2193. [Google Scholar] [CrossRef]
  101. Andrés-Barrao, C.; Alzubaidy, H.; Jalal, R.; Mariappan, K.G.; de Zélicourt, A.; Bokhari, A.; Artyukh, O.; Alwutayd, K.; Rawat, A.; Shekhawat, K. Coordinated bacterial and plant sulfur metabolism in Enterobacter sp. SA187–induced plant salt stress tolerance. Proc. Natl. Acad. Sci. USA 2021, 118, e2107417118. [Google Scholar] [CrossRef]
  102. Zhao, S.; Wei, H.; Lin, C.-Y.; Zeng, Y.; Tucker, M.P.; Himmel, M.E.; Ding, S.-Y. Burkholderia phytofirmans Inoculation-Induced Changes on the Shoot Cell Anatomy and Iron Accumulation Reveal Novel Components of Arabidopsis-Endophyte Interaction that Can Benefit Downstream Biomass Deconstruction. Front. Plant Sci. 2016, 7, 24. [Google Scholar] [CrossRef]
  103. Chen, Z.; Jin, Y.; Yao, X.; Chen, T.; Wei, X.; Li, C.; White, J.F.; Nan, Z. Fungal Endophyte Improves Survival of Lolium perenne in Low Fertility Soils by Increasing Root Growth, Metabolic Activity and Absorption of Nutrients. Plant Soil 2020, 452, 185–206. [Google Scholar] [CrossRef]
  104. Khan, Z.; Rho, H.; Firrincieli, A.; Hung, S.H.; Luna, V.; Masciarelli, O.; Kim, S.-H.; Doty, S.L. Growth enhancement and drought tolerance of hybrid poplar upon inoculation with endophyte consortia. Curr. Plant Biol. 2016, 6, 38–47. [Google Scholar] [CrossRef]
  105. Rana, K.L.; Kour, D.; Sheikh, I.; Yadav, N.; Yadav, A.N.; Kumar, V.; Singh, B.P.; Dhaliwal, H.S.; Saxena, A.K. Biodiversity of Endophytic Fungi from Diverse Niches and Their Biotechnological Applications. In Advances in Endophytic Fungal Research: Present Status and Future Challenges; Singh, B.P., Ed.; Springer International Publishing: Cham, Switzerland, 2019; pp. 105–144. [Google Scholar]
  106. Rodriguez, R.J.; White, J.F., Jr.; Arnold, A.E.; Redman, R.S. Fungal endophytes: Diversity and functional roles. New Phytol. 2009, 182, 314–330. [Google Scholar] [CrossRef] [PubMed]
  107. Singh, M.; Kumar, A.; Singh, R.; Pandey, K.D. Endophytic bacteria: A new source of bioactive compounds. Biotech 2017, 7, 315. [Google Scholar] [CrossRef] [PubMed]
  108. Sridhar, K.R. Diversity, Ecology, and Significance of Fungal Endophytes. In Endophytes and Secondary Metabolites; Jha, S., Ed.; Springer International Publishing: Cham, Switzerland, 2019; pp. 61–100. [Google Scholar]
  109. Tonial, F.; Falcão De Macedo Nava, F.; Luisa Gayger, A.; Bernardon Mar, T. Endophytes Potential Use in Crop Production; IntechOpen: London, UK, 2020. [Google Scholar]
  110. Yadav, A.; Yadav, K. Exploring the potential of endophytes in agriculture: A minireview. Adv. Plants Agric. Res. 2017, 6, 102–106. [Google Scholar] [CrossRef]
  111. Singh, S.; Kumar, V.; Dhanjal, D.S.; Sidhu, G.K.; Datta, S.; Kumar, S.; Singh, J. Endophytic microbes in abiotic stress management. In Microbial Endophytes; Elsevier: Amsterdam, The Netherlands, 2020; pp. 91–123. [Google Scholar]
  112. Liu, D.; Zhang, P.; Chen, D.; Howell, K. From the Vineyard to the Winery: How Microbial Ecology Drives Regional Distinctiveness of Wine. Front. Microbiol. 2019, 10, 2679. [Google Scholar] [CrossRef]
  113. Saikkonen, K.; Wäli, P.; Helander, M.; Faeth, S.H. Evolution of endophyte–plant symbioses. Trends Plant Sci. 2004, 9, 275–280. [Google Scholar] [CrossRef]
  114. Rizzo, D.M.; Lichtveld, M.; Mazet, J.A.; Togami, E.; Miller, S.A. Plant health and its effects on food safety and security in a One Health framework: Four case studies. One Health Outlook 2021, 3, 6. [Google Scholar] [CrossRef]
  115. Zappia, R.E.; Hüberli, D.; Hardy, G.E.S.J.; Bayliss, K.L. Fungi and oomycetes in open irrigation systems: Knowledge gaps and biosecurity implications. Plant Pathol. 2014, 63, 961–972. [Google Scholar] [CrossRef]
  116. Agrios, G.N. Plant Pathogens and Disease: General Introduction. In Encyclopedia of Microbiology, 3rd ed.; Schaechter, M., Ed.; Academic Press: Oxford, UK, 2009; pp. 613–646. [Google Scholar]
  117. Delgado-Baquerizo, M.; Guerra, C.A.; Cano-Díaz, C.; Egidi, E.; Wang, J.-T.; Eisenhauer, N.; Singh, B.K.; Maestre, F.T. The proportion of soil-borne pathogens increases with warming at the global scale. Nat. Clim. Chang. 2020, 10, 550–554. [Google Scholar] [CrossRef]
  118. Zhu, Y.; Zhu, S.; Zhang, F.; Zhao, Z.; Christensen, M.J.; Nan, Z.; Zhang, X. Transcriptomic Analyses Reveals Molecular Regulation of Photosynthesis by Epichloe endophyte in Achnatherum inebrians under Blumeria graminis Infection. J. Fungi 2022, 8, 1201. [Google Scholar] [CrossRef]
  119. Li, X.Z.; Simpson, W.R.; Song, M.L.; Bao, G.S.; Niu, X.L.; Zhang, Z.H.; Xu, H.F.; Liu, X.; Li, Y.L.; Li, C.J. Effects of seed moisture content and Epichloe endophyte on germination and physiology of Achnatherum inebrians. S. Afr. J. Bot. 2020, 134, 407–414. [Google Scholar] [CrossRef]
  120. Thor, K. Calcium—Nutrient and Messenger. Front. Plant Sci. 2019, 10, 440. [Google Scholar] [CrossRef] [PubMed]
  121. Dodds, P.N.; Rathjen, J.P. Plant immunity: Towards an integrated view of plant–pathogen interactions. Nat. Rev. Genet. 2010, 11, 539–548. [Google Scholar] [CrossRef]
  122. Vadassery, J.; Ranf, S.; Drzewiecki, C.; Mithöfer, A.; Mazars, C.; Scheel, D.; Lee, J.; Oelmüller, R. A cell wall extract from the endophytic fungus Piriformospora indica promotes growth of Arabidopsis seedlings and induces intracellular calcium elevation in roots. Plant J. 2009, 59, 193–206. [Google Scholar] [CrossRef] [PubMed]
  123. Wicaksono, W.A.; Jones, E.E.; Casonato, S.; Monk, J.; Ridgway, H.J. Biological control of Pseudomonas syringae pv. actinidiae (Psa), the causal agent of bacterial canker of kiwifruit, using endophytic bacteria recovered from a medicinal plant. Biol. Control 2018, 116, 103–112. [Google Scholar]
  124. Carrieri, R.; Borriello, G.; Piccirillo, G.; Lahoz, E.; Sorrentino, R.; Cermola, M.; Censi, S.B.; Grauso, L.; Mangoni, A.; Vinale, F. Antibiotic Activity of a Paraphaeosphaeria sporulosa-Produced Diketopiperazine against Salmonella enterica. J. Fungi 2020, 6, 83. [Google Scholar] [CrossRef] [PubMed]
  125. de Carvalho, M.; Abraham, W.R. Antimicrobial and Biofilm Inhibiting Diketopiperazines. Curr. Med. Chem. 2012, 19, 3564–3577. [Google Scholar] [CrossRef]
  126. Tembo, B.; Mulenga, R.M.; Sichilima, S.; M’siska, K.K.; Mwale, M.; Chikoti, P.C.; Singh, P.K.; He, X.; Pedley, K.F.; Peterson, G.L. Detection and characterization of fungus (Magnaporthe oryzae pathotype Triticum) causing wheat blast disease on rain-fed grown wheat (Triticum aestivum L.) in Zambia. PLoS ONE 2020, 15, e0238724. [Google Scholar] [CrossRef]
  127. 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, 60, 315–328. [Google Scholar] [CrossRef]
  128. Şahin, M.; Bayram, G.K.; Can, M.; Ayan, İ.; Acar, Z. Perennıal Ryegrass (Lolium perenne L.). Ph.D. Thesis, Oregon State University, Corvallis, OR, USA, 2022. [Google Scholar]
  129. Zabaleta, G.; Lee, S.T.; Cook, D.; Aguilar, M.; Iannone, L.J.; Robles, C.; Martinez, A. Indole-diterpenes alkaloid profiles of native grasses involved in tremorgenic syndromes in the Argentine Patagonia. Toxicon 2022, 217, 107–111. [Google Scholar] [CrossRef]
  130. Currie, A.F.; Gange, A.C.; Ab Razak, N.; Ellison, C.A.; Maczey, N.; Wood, S.V. Endophytic fungi in the invasive weed Impatiens glandulifera: A barrier to classical biological control? Weed Res. 2020, 60, 50–59. [Google Scholar] [CrossRef]
  131. Sharma, A.; Kaushik, N.; Sharma, A.; Marzouk, T.; Djébali, N. Exploring the potential of endophytes and their metabolites for bio-control activity. 3 Biotech 2022, 12, 277. [Google Scholar] [CrossRef] [PubMed]
  132. Caruso, D.J.; Palombo, E.A.; Moulton, S.E.; Zaferanloo, B. Exploring the Promise of Endophytic Fungi: A Review of Novel Antimicrobial Compounds. Microorganisms 2022, 10, 1990. [Google Scholar] [CrossRef] [PubMed]
  133. Dubey, A.; Malla, M.A.; Kumar, A.; Dayanandan, S.; Khan, M.L. Plants endophytes: Unveiling hidden agenda for bioprospecting toward sustainable agriculture. Crit. Rev. Biotechnol. 2020, 40, 1210–1231. [Google Scholar] [CrossRef] [PubMed]
  134. Newman, D.J.; Cragg, G.M. Natural Products as Sources of New Drugs over the Nearly Four Decades from 01/1981 to 09/2019. J. Nat. Prod. 2020, 83, 770–803. [Google Scholar] [CrossRef]
  135. Kenawy, A.; Dailin, D.J.; Abo-Zaid, G.A.; Malek, R.A.; Ambehabati, K.K.; Zakaria, K.H.N.; Sayyed, R.; El Enshasy, H.A. Biosynthesis of antibiotics by PGPR and their roles in biocontrol of plant diseases. In Plant Growth Promoting Rhizobacteria for Sustainable Stress Management: Volume 2: Rhizobacteria in Biotic Stress Management; Springer: Singapore, 2019; pp. 1–35. [Google Scholar]
  136. Julian, W.T.; Vasilchenko, A.V.; Shpindyuk, D.D.; Poshvina, D.V.; Vasilchenko, A.S. Bacterial-derived plant protection metabolite 2, 4-diacetylphloroglucinol: Effects on bacterial cells at inhibitory and subinhibitory concentrations. Biomolecules 2020, 11, 13. [Google Scholar] [CrossRef] [PubMed]
  137. Aleti, G. Secondary Metabolite Biosynthetic Gene Diversity–Bacillales Non-Ribosomal Peptides and Polyketides in Plant-Microbe Interactions. Ph.D. Thesis, Austrian Institute of Technology, Vienna, Austria, 2016. [Google Scholar]
  138. Eid, A.M.; Salim, S.S.; Hassan, S.E.-D.; Ismail, M.A.; Fouda, A. Role of endophytes in plant health and abiotic stress management. In Microbiome in Plant Health and Disease; Springer: Singapore, 2019; pp. 119–144. [Google Scholar]
  139. Pandey, P.K.; Samanta, R.; Yadav, R.N.S. Inside the plant: Addressing bacterial endophytes in biotic stress alleviation. Arch. Microbiol. 2019, 201, 415–429. [Google Scholar] [CrossRef]
  140. Brosset, A.; Blande, J.D. Volatile-mediated plant–plant interactions: Volatile organic compounds as modulators of receiver plant defence, growth, and reproduction. J. Exp. Bot. 2022, 73, 511–528. [Google Scholar] [CrossRef]
  141. Bastías, D.A.; Martínez-Ghersa, M.A.; Newman, J.A.; Card, S.D.; Mace, W.J.; Gundel, P.E. Jasmonic acid regulation of the anti-herbivory mechanism conferred by fungal endophytes in grasses. J. Ecol. 2018, 106, 2365–2379. [Google Scholar] [CrossRef]
  142. Zhu, J.; He, Y.; Yan, X.; Liu, L.; Guo, R.; Xia, X.; Cheng, D.; Mi, X.; Samarina, L.; Liu, S.; et al. Duplication and transcriptional divergence of three Kunitz protease inhibitor genes that modulate insect and pathogen defenses in tea plant (Camellia sinensis). Hortic. Res. 2019, 6, 126. [Google Scholar] [CrossRef]
  143. Datta, D.; Behera, L.; Chaudhary, V.T.; Kumar, S.; Bisen, K. Endophytes: Rendering Systemic Resistance to Plants. In Rhizosphere Microbes: Biotic Stress Management; Singh, U.B., Sahu, P.K., Singh, H.V., Sharma, P.K., Sharma, S.K., Eds.; Springer Nature Singapore: Singapore, 2022; pp. 175–195. [Google Scholar]
  144. Mastan, A.; Rane, D.; Dastager, S.G.; Babu, C.V. Development of low-cost plant probiotic formulations of functional endophytes for sustainable cultivation of Coleus forskohlii. Microbiol. Res. 2019, 227, 126310. [Google Scholar] [CrossRef] [PubMed]
  145. Wang, H.; Peng, H.; Li, W.; Cheng, P.; Gong, M. The Toxins of Beauveria bassiana and the Strategies to Improve Their Virulence to Insects. Front. Microbiol. 2021, 12, 705343. [Google Scholar] [CrossRef] [PubMed]
  146. Feng, P.; Shang, Y.; Cen, K.; Wang, C. Fungal biosynthesis of the bibenzoquinone oosporein to evade insect immunity. Proc. Natl. Acad. Sci. USA 2015, 112, 11365–11370. [Google Scholar] [CrossRef] [PubMed]
  147. Blumberg, B.J.; Short, S.M.; Dimopoulos, G. Chapter 15—Employing the Mosquito Microflora for Disease Control. In Genetic Control of Malaria and Dengue; Adelman, Z.N., Ed.; Academic Press: Boston, MA, USA, 2016; pp. 335–362. [Google Scholar]
  148. Vitorino, L.C.; Silva, F.G.; Soares, M.A.; Souchie, E.L.; Costa, A.C.; Lima, W.C. Solubilization of calcium and iron phosphate and in vitro production of Indoleacetic acid by Endophytic isolates of Hyptis marrubioides Epling (Lamiaceae). Int. Res. J. Biotechnol. 2012, 3, 47–54. [Google Scholar]
  149. Nemhauser, J.L. Back to basics: What is the function of an Aux/IAA in auxin response? New Phytol. 2018, 218, 1295–1297. [Google Scholar] [CrossRef]
  150. Khan, A.L.; Hussain, J.; Al-Harrasi, A.; Al-Rawahi, A.; Lee, I.-J. Endophytic fungi: Resource for gibberellins and crop abiotic stress resistance. Crit. Rev. Biotechnol. 2015, 35, 62–74. [Google Scholar]
  151. Liu, R.; Yang, L.; Zou, Y.; Wu, Q. Root-associated endophytic fungi modulate endogenous auxin and cytokinin levels to improve plant biomass and root morphology of trifoliate orange. Hortic. Plant J. 2022. [Google Scholar] [CrossRef]
  152. Vanderstraeten, L.; Van Der Straeten, D. Accumulation and Transport of 1-Aminocyclopropane-1-Carboxylic Acid (ACC) in Plants: Current Status, Considerations for Future Research and Agronomic Applications. Front Plant Sci 2017, 8, 38. [Google Scholar] [CrossRef]
  153. Gupta, S.; Pandey, S. ACC deaminase producing bacteria with multifarious plant growth promoting traits alleviates salinity stress in French bean (Phaseolus vulgaris) plants. Front. Microbiol 2019, 10, 1506. [Google Scholar] [CrossRef]
  154. Roy Choudhury, A.; Choi, J.; Walitang, D.I.; Trivedi, P.; Lee, Y.; Sa, T. ACC deaminase and indole acetic acid producing endophytic bacterial co-inoculation improves physiological traits of red pepper (Capsicum annum L.) under salt stress. J. Plant Physiol. 2021, 267, 153544. [Google Scholar] [CrossRef]
  155. Canellas, L.P.; Olivares, F.L. Physiological responses to humic substances as plant growth promoter. Chem. Biol. Technol. Agric. 2014, 1, 3. [Google Scholar] [CrossRef]
  156. Sponsel, V.M. Gibberellins. In Encyclopedia of Hormones; Henry, H.L., Norman, A.W., Eds.; Academic Press: New York, NY, USA, 2003; pp. 29–40. [Google Scholar]
  157. Gao, X.; Zhang, Y.; He, Z.; Fu, X. 4—Gibberellins. In Hormone Metabolism and Signaling in Plants; Li, J., Li, C., Smith, S.M., Eds.; Academic Press: Cambridge, MA, USA, 2017; pp. 107–160. [Google Scholar]
  158. Mukherjee, A.; Gaurav, A.K.; Singh, S.; Yadav, S.; Bhowmick, S.; Abeysinghe, S.; Verma, J.P. The Bioactive potential of phytohormones: A review. Biotechnol. Rep. 2022, 35, e00748. [Google Scholar] [CrossRef]
  159. Leitão, A.L.; Enguita, F.J. Gibberellins in Penicillium strains: Challenges for endophyte-plant host interactions under salinity stress. Microbiol. Res. 2016, 183, 8–18. [Google Scholar] [CrossRef]
  160. Muñoz-Espinoza, V.A.; López-Climent, M.F.; Casaretto, J.A.; Gómez-Cadenas, A. Water stress responses of tomato mutants impaired in hormone biosynthesis reveal abscisic acid, jasmonic acid and salicylic acid interactions. Front. Plant Sci. 2015, 6, 997. [Google Scholar] [CrossRef]
  161. Wang, X.-M.; Yang, B.; Ren, C.-G.; Wang, H.-W.; Wang, J.-Y.; Dai, C.-C. Involvement of abscisic acid and salicylic acid in signal cascade regulating bacterial endophyte-induced volatile oil biosynthesis in plantlets of Atractylodes lancea. Physiol. Plant. 2015, 153, 30–42. [Google Scholar] [CrossRef] [PubMed]
  162. Hou, S.; Tsuda, K. Salicylic acid and jasmonic acid crosstalk in plant immunity. Essays Biochem. 2022, 66, 647–656. [Google Scholar]
  163. Cui, X.; He, W.; Christensen, M.J.; Yue, J.; Zeng, F.; Zhang, X.; Nan, Z.; Xia, C. Abscisic Acid May Play a Critical Role in the Moderating Effect of Epichloe Endophyte on Achnatherum inebrians under Drought Stress. J. Fungi 2022, 8, 1140. [Google Scholar] [CrossRef] [PubMed]
  164. Waqas, M.; Khan, A.L.; Kamran, M.; Hamayun, M.; Kang, S.-M.; Kim, Y.-H.; Lee, I.-J. Endophytic Fungi Produce Gibberellins and Indoleacetic Acid and Promotes Host-Plant Growth during Stress. Molecules 2012, 17, 10754–10773. [Google Scholar] [CrossRef] [PubMed]
  165. Singh, S.; Awasthi, L.P.; Verma, H.N. Chapter 42—Systemic resistance inducers from plants—An ecofriendly approach for the management of viral diseases of crops. In Applied Plant Virology; Awasthi, L.P., Ed.; Academic Press: Cambridge, MA, USA, 2020; pp. 603–617. [Google Scholar]
  166. Lefevere, H.; Bauters, L.; Gheysen, G. Salicylic Acid Biosynthesis in Plants. Front. Plant Sci. 2020, 11, 338. [Google Scholar] [CrossRef]
  167. Jisha, S.; Sabu, K.; Manjula, S. Multifunctional aspects of Piriformospora indica in plant endosymbiosis. Mycology 2019, 10, 182–190. [Google Scholar] [CrossRef]
  168. Mollah, M.M.I.; Choi, H.W.; Yeam, I.; Lee, J.M.; Kim, Y. Salicylic Acid, a Plant Hormone, Suppresses Phytophagous Insect Immune Response by Interrupting HMG-Like DSP1. Front. Physiol. 2021, 12, 744272. [Google Scholar] [CrossRef] [PubMed]
  169. Tripathi, D.; Raikhy, G.; Kumar, D. Chemical elicitors of systemic acquired resistance—Salicylic acid and its functional analogs. Curr. Plant Biol. 2019, 17, 48–59. [Google Scholar] [CrossRef]
  170. Zhou, Q.; Meng, Q.; Tan, X.; Ding, W.; Ma, K.; Xu, Z.; Huang, X.; Gao, H. Protein Phosphorylation Changes during Systemic Acquired Resistance in Arabidopsis thaliana. Front Plant Sci. 2021, 12, 748287. [Google Scholar] [CrossRef] [PubMed]
  171. Haq, S.K.; Atif, S.M.; Khan, R.H. Protein proteinase inhibitor genes in combat against insects, pests, and pathogens: Natural and engineered phytoprotection. Arch. Biochem. Biophys. 2004, 431, 145–159. [Google Scholar] [CrossRef]
  172. De Leo, F.; Volpicella, M.; Licciulli, F.; Liuni, S.; Gallerani, R.; Ceci, L.R. PLANT-PIs: A database for plant protease inhibitors and their genes. Nucleic Acids Res. 2002, 30, 347–348. [Google Scholar] [CrossRef]
  173. He, Y.; Chen, T.; Zhang, H.; White, J.F.; Li, C. Fungal Endophytes Help Grasses to Tolerate Sap-Sucking Herbivores through a Hormone-Signaling System. J. Plant Growth Regul. 2022, 41, 2122–2137. [Google Scholar] [CrossRef]
  174. Lyons, R.; Manners, J.M.; Kazan, K. Jasmonate biosynthesis and signaling in monocots: A comparative overview. Plant Cell Rep. 2013, 32, 815–827. [Google Scholar] [CrossRef] [PubMed]
  175. Rynkiewicz, M.J.; Cane, D.E.; Christianson, D.W. Structure of trichodiene synthase from Fusarium sporotrichioides provides mechanistic inferences on the terpene cyclization cascade. Proc. Natl. Acad. Sci. USA 2001, 98, 13543–13548. [Google Scholar] [CrossRef]
  176. Grosskinsky, D.K.; Naseem, M.; Abdelmohsen, U.R.; Plickert, N.; Engelke, T.; Griebel, T.; Zeier, J.; Novák, O.; Strnad, M.; Pfeifhofer, H.; et al. Cytokinins mediate resistance against Pseudomonas syringae in tobacco through increased antimicrobial phytoalexin synthesis independent of salicylic acid signaling. Plant Physiol. 2011, 157, 815–830. [Google Scholar] [CrossRef]
  177. Choi, J.; Huh, S.U.; Kojima, M.; Sakakibara, H.; Paek, K.-H.; Hwang, I. The Cytokinin-Activated Transcription Factor ARR2 Promotes Plant Immunity via TGA3/NPR1-Dependent Salicylic Acid Signaling in Arabidopsis. Dev. Cell 2010, 19, 284–295. [Google Scholar] [CrossRef]
  178. Großkinsky, D.K.; Tafner, R.; Moreno, M.V.; Stenglein, S.A.; García De Salamone, I.E.; Nelson, L.M.; Novák, O.; Strnad, M.; Van Der Graaff, E.; Roitsch, T. Cytokinin production by Pseudomonas fluorescens G20-18 determines biocontrol activity against Pseudomonas syringae in Arabidopsis. Sci. Rep. 2016, 6, 23310. [Google Scholar] [CrossRef] [PubMed]
  179. Jameson, P. Cytokinins and auxins in plant-pathogen interactions–An overview. Plant Growth Regul. 2000, 32, 369–380. [Google Scholar] [CrossRef]
  180. Babosha, A.V. Regulation of resistance and susceptibility in wheat–powdery mildew pathosystem with exogenous cytokinins. J. Plant Physiol. 2009, 166, 1892–1903. [Google Scholar] [CrossRef] [PubMed]
  181. Gupta, R.; Anand, G.; Pizarro, L.; Bar-Yosef, D.L.; Kovetz, N.; Sela, N.; Yehuda, T.; Gazit, E.; Bar, M. Cytokinin Inhibits Fungal Development and Virulence by Targeting the Cytoskeleton and Cellular Trafficking. mBio 2021, 12, e03068-20. [Google Scholar] [CrossRef] [PubMed]
  182. Li, X.; Zhou, J.; Xu, R.-S.; Meng, M.-Y.; Yu, X.; Dai, C.-C. Auxin, Cytokinin, and Ethylene Involved in Rice N Availability Improvement Caused by Endophyte Phomopsis liquidambari. J. Plant Growth Regul. 2018, 37, 128–143. [Google Scholar] [CrossRef]
  183. Sachdev, S.; Ansari, S.A.; Ansari, M.I.; Fujita, M.; Hasanuzzaman, M. Abiotic Stress and Reactive Oxygen Species: Generation, Signaling, and Defense Mechanisms. Antioxidants 2021, 10, 277. [Google Scholar] [CrossRef] [PubMed]
  184. Mhamdi, A.; Van Breusegem, F. Reactive oxygen species in plant development. Development 2018, 145, dev164376. [Google Scholar] [CrossRef] [PubMed]
  185. Bela, K.; Horváth, E.; Gallé, Á.; Szabados, L.; Tari, I.; Csiszár, J. Plant glutathione peroxidases: Emerging role of the antioxidant enzymes in plant development and stress responses. J. Plant Physiol. 2015, 176, 192–201. [Google Scholar] [CrossRef]
  186. Tomar, R.S.; Kataria, S.; Jajoo, A. Behind the scene: Critical role of reactive oxygen species and reactive nitrogen species in salt stress tolerance. J. Agron. Crop Sci. 2021, 207, 577–588. [Google Scholar] [CrossRef]
  187. Hereme, R.; Morales-Navarro, S.; Ballesteros, G.; Barrera, A.; Ramos, P.; Gundel, P.E.; Molina-Montenegro, M.A. Fungal Endophytes Exert Positive Effects on Colobanthus quitensis Under Water Stress but Neutral Under a Projected Climate Change Scenario in Antarctica. Front. Microbiol. 2020, 11, 264. [Google Scholar] [CrossRef]
  188. Mahgoub, H.A.M.; Fouda, A.; Eid, A.M.; Ewais, E.E.-D.; Hassan, S.E.-D. Biotechnological application of plant growth-promoting endophytic bacteria isolated from halophytic plants to ameliorate salinity tolerance of Vicia faba L. Plant Biotechnol. Rep. 2021, 15, 819–843. [Google Scholar] [CrossRef]
  189. Khan, M.A.; Asaf, S.; Khan, A.L.; Ullah, I.; Ali, S.; Kang, S.-M.; Lee, I.-J. Alleviation of salt stress response in soybean plants with the endophytic bacterial isolate Curtobacterium sp. SAK1. Ann. Microbiol. 2019, 69, 797–808. [Google Scholar] [CrossRef]
  190. Iqbal, N.; Umar, S.; Khan, N. Nitrogen availability regulates proline and ethylene production and alleviates salinity stress in mustard. J. Plant Physiol. 2015, 178, 84–91. [Google Scholar] [CrossRef] [PubMed]
  191. Kim, J.; Harter, K.; Theologis, A. Protein-protein interactions among the Aux/IAA proteins. Proc. Natl. Acad. Sci. USA 1997, 94, 11786–11791. [Google Scholar] [CrossRef] [PubMed]
  192. Wu, T.; Zhang, H.-T.; Wang, Y.; Jia, W.-S.; Xu, X.-F.; Zhang, X.-Z.; Han, Z.H. Induction of root Fe(lll) reductase activity and proton extrusion by iron deficiency is mediated by auxin-based systemic signalling in Malus xiaojinensis. J. Exp. Bot. 2011, 63, 859–870. [Google Scholar] [CrossRef]
  193. Higgins, T.J.; Jacobsen, J.V.; Zwar, J.A. Gibberellic acid and abscisic acid modulate protein synthesis and mRNA levels in barley aleurone layers. Plant Mol. Biol. 1982, 1, 191–215. [Google Scholar] [CrossRef]
  194. Raghavendra, A.S.; Gonugunta, V.K.; Christmann, A.; Grill, E. ABA perception and signalling. Trends Plant Sci. 2010, 15, 395–401. [Google Scholar] [CrossRef]
  195. Fadiji, A.E.; Babalola, O.O. Elucidating Mechanisms of Endophytes Used in Plant Protection and Other Bioactivities With Multifunctional Prospects. Front. Bioeng. Biotechnol. 2020, 8, 467. [Google Scholar] [CrossRef]
  196. Shahid, M.A.; Sarkhosh, A.; Khan, N.; Balal, R.M.; Ali, S.; Rossi, L.; Gómez, C.; Mattson, N.; Nasim, W.; Garcia-Sanchez, F. Insights into the physiological and biochemical impacts of salt stress on plant growth and development. Agronomy 2020, 10, 938. [Google Scholar] [CrossRef]
  197. Schmidt, W.; Thomine, S.; Buckhout, T.J. Iron nutrition and interactions in plants. Front. Plant Sci. 2020, 10, 1670. [Google Scholar] [CrossRef]
  198. Murata, Y.; Itoh, Y.; Iwashita, T.; Namba, K. Transgenic Petunia with the Iron(III)-Phytosiderophore Transporter Gene Acquires Tolerance to Iron Deficiency in Alkaline Environments. PLoS ONE 2015, 10, e0120227. [Google Scholar] [CrossRef] [PubMed]
  199. Cesco, S.; Rombolà, A.D.; Tagliavini, M.; Varanini, Z.; Pinton, R. Phytosiderophores released by graminaceous species promote 59Fe-uptake in citrus. Plant Soil 2006, 287, 223–233. [Google Scholar] [CrossRef]
  200. Banakar, R.; Alvarez Fernandez, A.; Díaz-Benito, P.; Abadia, J.; Capell, T.; Christou, P. Phytosiderophores determine thresholds for iron and zinc accumulation in biofortified rice endosperm while inhibiting the accumulation of cadmium. J. Exp. Bot. 2017, 68, 4983–4995. [Google Scholar] [CrossRef] [PubMed]
  201. Ma, Y.; Rajkumar, M.; Zhang, C.; Freitas, H. Beneficial role of bacterial endophytes in heavy metal phytoremediation. J. Environ. Manag. 2016, 174, 14–25. [Google Scholar] [CrossRef]
  202. Du, Y.-C.; Kong, L.-J.; Cao, L.-S.; Zhang, W.; Zhu, Q.; Ma, C.-Y.; Sun, K.; Dai, C.-C. Endophytic Fungus Phomopsis liquidambaris Enhances Fe Absorption in Peanuts by Reducing Hydrogen Peroxide. Front. Plant Sci. 2022, 13, 872242. [Google Scholar] [CrossRef] [PubMed]
  203. Long, L.; Persson, D.P.; Duan, F.; Jørgensen, K.; Yuan, L.; Schjoerring, J.K.; Pedas, P.R. The iron-regulated transporter 1 plays an essential role in uptake, translocation and grain-loading of manganese, but not iron, in barley. New Phytol. 2018, 217, 1640–1653. [Google Scholar] [CrossRef]
  204. Nielsen, F.H. History of zinc in agriculture. Adv. Nutr. 2012, 3, 783–789. [Google Scholar] [CrossRef]
  205. Umair Hassan, M.; Aamer, M.; Umer Chattha, M.; Haiying, T.; Shahzad, B.; Barbanti, L.; Nawaz, M.; Rasheed, A.; Afzal, A.; Liu, Y. The critical role of zinc in plants facing the drought stress. Agriculture 2020, 10, 396. [Google Scholar] [CrossRef]
  206. Kumar, A.; Singh, B.; Raigond, P.; Sahu, C.; Mishra, U.N.; Sharma, S.; Lal, M.K. Phytic acid: Blessing in disguise, a prime compound required for both plant and human nutrition. Food Res. Int. 2021, 142, 110193. [Google Scholar] [CrossRef]
  207. Wang, S.; Li, L.; Ying, Y.; Wang, J.; Shao, J.F.; Yamaji, N.; Whelan, J.; Ma, J.F.; Shou, H. A transcription factor OsbHLH156 regulates Strategy II iron acquisition through localising IRO2 to the nucleus in rice. New Phytol. 2020, 225, 1247–1260. [Google Scholar] [CrossRef]
  208. Broadley, M.R.; White, P.J.; Hammond, J.P.; Zelko, I.; Lux, A. Zinc in plants. New Phytol. 2007, 173, 677–702. [Google Scholar] [CrossRef] [PubMed]
  209. Jahandideh Mahjen Abadi, V.A.; Sepehri, M.; Khatabi, B.; Rezaei, M. Alleviation of zinc deficiency in wheat inoculated with root endophytic fungus Piriformospora indica and rhizobacterium Pseudomonas putida. Rhizosphere 2021, 17, 100311. [Google Scholar] [CrossRef]
  210. Padda, K.P.; Puri, A.; Chanway, C. Endophytic nitrogen fixation—A possible ‘hidden’ source of nitrogen for lodgepole pine trees growing at unreclaimed gravel mining sites. FEMS Microbiol. Ecol. 2019, 95, fiz172. [Google Scholar] [CrossRef] [PubMed]
  211. Puri, A.; Padda, K.P.; Chanway, C.P. Evidence of endophytic diazotrophic bacteria in lodgepole pine and hybrid white spruce trees growing in soils with different nutrient statuses in the West Chilcotin region of British Columbia, Canada. For. Ecol. Manag. 2018, 430, 558–565. [Google Scholar] [CrossRef]
  212. Reis, I.A.; Souza, M.G.D.; Granja-Salcedo, Y.T.; Carvalho, I.P.C.D.; Porcionato, M.A.D.F.; Prados, L.F.; Siqueira, G.R.; De Resende, F.D. Effect of Post-Ruminal Urea Supply on Growth Performance of Grazing Nellore Young Bulls at Dry Season. Animals 2023, 13, 207. [Google Scholar] [CrossRef]
  213. Scherger, L.E.; Zanello, V.; Lexow, C. Impact of urea and ammoniacal nitrogen wastewaters on soil: Field study in a fertilizer industry (Bahía Blanca, Argentina). Bull. Environ. Contam. Toxicol. 2021, 107, 565–573. [Google Scholar] [CrossRef]
  214. Yang, Z.; Wang, Y.; Wang, K.; Zhang, Y.; Yu, N.; Liu, Z. Effects of urea application on the reproduction of Pardosa pseudoannulata: Field and laboratory studies. Chemosphere 2022, 301, 134697. [Google Scholar] [CrossRef]
  215. Varga, T.; Hixson, K.K.; Ahkami, A.H.; Sher, A.W.; Barnes, M.E.; Chu, R.K.; Battu, A.K.; Nicora, C.D.; Winkler, T.E.; Reno, L.R.; et al. Endophyte-Promoted Phosphorus Solubilization in Populus. Front. Plant Sci. 2020, 11, 567918. [Google Scholar] [CrossRef]
  216. Emami, S.; Alikhani, H.A.; Pourbabaee, A.A.; Etesami, H.; Motasharezadeh, B.; Sarmadian, F. Consortium of endophyte and rhizosphere phosphate solubilizing bacteria improves phosphorous use efficiency in wheat cultivars in phosphorus deficient soils. Rhizosphere 2020, 14, 100196. [Google Scholar] [CrossRef]
  217. Hegedűs, M.; Tóth-Bodrogi, E.; Németh, S.; Somlai, J.; Kovács, T. Radiological investigation of phosphate fertilizers: Leaching studies. J. Environ. Radioact. 2017, 173, 34–43. [Google Scholar] [CrossRef]
  218. Chhabra, S.; Dowling, D.N. Endophyte-Promoted Nutrient Acquisition: Phosphorus and Iron; Springer International Publishing: Cham, Switzerland, 2017; pp. 21–42. [Google Scholar]
  219. Baghel, V.; Thakur, J.K.; Yadav, S.S.; Manna, M.C.; Mandal, A.; Shirale, A.O.; Sharma, P.; Sinha, N.K.; Mohanty, M.; Singh, A.B.; et al. Phosphorus and Potassium Solubilization From Rock Minerals by Endophytic Burkholderia sp. Strain FDN2-1 in Soil and Shift in Diversity of Bacterial Endophytes of Corn Root Tissue with Crop Growth Stage. Geomicrobiol. J. 2020, 37, 550–563. [Google Scholar] [CrossRef]
  220. Kumar, M.; Yadav, V.; Kumar, H.; Sharma, R.; Singh, A.; Tuteja, N.; Johri, A.K. Piriformospora indica enhances plant growth by transferring phosphate. Plant Signal. Behav. 2011, 6, 723–725. [Google Scholar] [CrossRef] [PubMed]
  221. Etesami, H.; Jeong, B.R.; Glick, B.R. Contribution of arbuscular mycorrhizal fungi, phosphate–solubilizing bacteria, and silicon to P uptake by plant. Front. Plant Sci. 2021, 12, 699618. [Google Scholar] [CrossRef] [PubMed]
  222. Bargaz, A.; Elhaissoufi, W.; Khourchi, S.; Benmrid, B.; Borden, K.A.; Rchiad, Z. Benefits of phosphate solubilizing bacteria on belowground crop performance for improved crop acquisition of phosphorus. Microbiol. Res. 2021, 252, 126842. [Google Scholar] [CrossRef] [PubMed]
  223. Li, X.; Ren, A.; Han, R.; Yin, L.; Wei, M.; Gao, Y. Endophyte-Mediated Effects on the Growth and Physiology of Achnatherum sibiricum Are Conditional on Both N and P Availability. PLoS ONE 2012, 7, e48010. [Google Scholar] [CrossRef]
  224. Hafsi, C.; Debez, A.; Abdelly, C. Potassium deficiency in plants: Effects and signaling cascades. Acta Physiol. Plant. 2014, 36, 1055–1070. [Google Scholar] [CrossRef]
  225. Sakaguchi, T.; Nishizawa, N.K.; Nakanishi, H.; Yoshimura, E.; Mori, S. The role of potassium in the secretion of mugineic acids familyphytosiderophores from iron-deficient barley roots. Plant Soil. 1999, 215, 221–227. [Google Scholar] [CrossRef]
  226. Takagi, S.-i. Naturally occurring iron-chelating compounds in oat-and rice-root washings: I. Activity measurement and preliminary characterization. Soil Sci. Plant Nutr. 1976, 22, 423–433. [Google Scholar] [CrossRef]
  227. Loka, D.A.; Oosterhuis, D.M.; Baxevanos, D.; Vlachostergios, D.; Hu, W. How potassium deficiency alters flower bud retention on cotton (Gossypium hirsutum L.). Arch. Agron. Soil Sci. 2019, 65, 521–536. [Google Scholar] [CrossRef]
  228. Hu, W.; Liu, Y.; Loka, D.A.; Zahoor, R.; Wang, S.; Zhou, Z. Drought limits pollen tube growth rate by altering carbohydrate metabolism in cotton (Gossypium hirsutum) pistils. Plant Sci. 2019, 286, 108–117. [Google Scholar] [CrossRef]
  229. Shaffer, J.P.; Carter, M.E.; Spraker, J.E.; Clark, M.; Smith, B.A.; Hockett, K.L.; Baltrus, D.A.; Arnold, A.E. Transcriptional Profiles of a Foliar Fungal Endophyte (Pestalotiopsis, Ascomycota) and Its Bacterial Symbiont (Luteibacter, Gammaproteobacteria) Reveal Sulfur Exchange and Growth Regulation during Early Phases of Symbiotic Interaction. Msystems 2022, 7, e00091-22. [Google Scholar] [CrossRef] [PubMed]
  230. Acevedo, E.; Galindo-Castañeda, T.; Prada, F.; Navia, M.; Romero, H.M. Phosphate-solubilizing microorganisms associated with the rhizosphere of oil palm (Elaeis guineensis Jacq.) in Colombia. Appl. Soil Ecol. 2014, 80, 26–33. [Google Scholar] [CrossRef]
  231. Kumar, A.; Singh, U.M.; Manohar, M.; Gaur, V.S. Calcium transport from source to sink: Understanding the mechanism(s) of acquisition, translocation, and accumulation for crop biofortification. Acta Physiol. Plant. 2015, 37, 1722. [Google Scholar] [CrossRef]
  232. White, J.F.; Kingsley, K.L.; Zhang, Q.; Verma, R.; Obi, N.; Dvinskikh, S.; Elmore, M.T.; Verma, S.K.; Gond, S.K.; Kowalski, K.P. Review: Endophytic microbes and their potential applications in crop management. Pest Manag. Sci. 2019, 75, 2558–2565. [Google Scholar] [CrossRef] [PubMed]
  233. Shahabivand, S.; Parvaneh, A.; Aliloo, A.A. Root endophytic fungus Piriformospora indica affected growth, cadmium partitioning and chlorophyll fluorescence of sunflower under cadmium toxicity. Ecotoxicol. Environ. Saf. 2017, 145, 496–502. [Google Scholar] [CrossRef]
  234. Guo, W.; Nazim, H.; Liang, Z.; Yang, D. Magnesium deficiency in plants: An urgent problem. Crop J. 2016, 4, 83–91. [Google Scholar] [CrossRef]
  235. Tian, X.-Y.; He, D.-D.; Bai, S.; Zeng, W.-Z.; Wang, Z.; Wang, M.; Wu, L.-Q.; Chen, Z.-C. Physiological and molecular advances in magnesium nutrition of plants. Plant Soil. 2021, 468, 1–17. [Google Scholar] [CrossRef]
  236. Adams, N.B.; Bisson, C.; Brindley, A.A.; Farmer, D.A.; Davison, P.A.; Reid, J.D.; Hunter, C.N. The active site of magnesium chelatase. Nat. Plants 2020, 6, 1491–1502. [Google Scholar] [CrossRef]
  237. Prasad, D.; Verma, N.; Bakshi, M.; Narayan, O.P.; Singh, A.K.; Dua, M.; Johri, A.K. Functional characterization of a magnesium transporter of root endophytic fungus Piriformospora indica. Front. Microbiol. 2019, 9, 3231. [Google Scholar] [CrossRef]
  238. Song, J.; Wang, Y.; Pan, Y.; Pang, J.; Zhang, X.; Fan, J.; Zhang, Y. The influence of nitrogen availability on anatomical and physiological responses of Populus alba × P. glandulosa to drought stress. BMC Plant Biol. 2019, 19, 63. [Google Scholar] [CrossRef]
  239. Ku, Y.S.; Rehman, H.M.; Lam, H.M. Possible Roles of Rhizospheric and Endophytic Microbes to Provide a Safe and Affordable Means of Crop Biofortification. Agronomy 2019, 9, 764. [Google Scholar] [CrossRef]
Figure 1. Visual representation of the effects of an increasing population requiring more housing. Increasing the demand for housing space and township expansion encroaches on farmland, some of which is unable to move relative to this growth. An increased efficiency is required from reduced farmland or crops will have to adapt to survive in these marginal conditions. Endophytes can offer potential solutions for the aforementioned issues.
Figure 1. Visual representation of the effects of an increasing population requiring more housing. Increasing the demand for housing space and township expansion encroaches on farmland, some of which is unable to move relative to this growth. An increased efficiency is required from reduced farmland or crops will have to adapt to survive in these marginal conditions. Endophytes can offer potential solutions for the aforementioned issues.
Microorganisms 11 01276 g001
Figure 3. Overview of range of applications and desirable endophytic impacts upon host plants.
Figure 3. Overview of range of applications and desirable endophytic impacts upon host plants.
Microorganisms 11 01276 g003
Figure 4. Iron accumulation strategies. (left) Strategy I is conducted by most non-grass plants, whereby iron is taken into the plant’s epidermis through a reduction of ferric iron to ferrous iron (FRO2). This pathway involves H+ soil acidification via ATPase AHA2 and phenolic excretion to improve iron availability. (Right) strategy II plants are mainly grasses and can directly uptake ferric iron when complexed with phytosiderophore (DMA) through transporters TOM1 and YS1. This pathway is much more energy efficient and tends to have a higher rate of iron accumulation.
Figure 4. Iron accumulation strategies. (left) Strategy I is conducted by most non-grass plants, whereby iron is taken into the plant’s epidermis through a reduction of ferric iron to ferrous iron (FRO2). This pathway involves H+ soil acidification via ATPase AHA2 and phenolic excretion to improve iron availability. (Right) strategy II plants are mainly grasses and can directly uptake ferric iron when complexed with phytosiderophore (DMA) through transporters TOM1 and YS1. This pathway is much more energy efficient and tends to have a higher rate of iron accumulation.
Microorganisms 11 01276 g004
Figure 5. Flowchart outlining the flow on effect that basic nutritional improvements, such as nitrogen, have on downstream processes. Endophyte association can improve a plants ability to metabolise ROS, reducing accumulation and resulting in a greater tolerance to saline conditions.
Figure 5. Flowchart outlining the flow on effect that basic nutritional improvements, such as nitrogen, have on downstream processes. Endophyte association can improve a plants ability to metabolise ROS, reducing accumulation and resulting in a greater tolerance to saline conditions.
Microorganisms 11 01276 g005
Figure 6. Sulphur containing amino acids found in plant proteins.
Figure 6. Sulphur containing amino acids found in plant proteins.
Microorganisms 11 01276 g006
Table 2. Summary table of key plant signalling molecules produced, stimulated, or co-produced by endophytes and their effects on plant health.
Table 2. Summary table of key plant signalling molecules produced, stimulated, or co-produced by endophytes and their effects on plant health.
Signalling MoleculeEffectReference
Auxins (IAA)Promote cell elongation, root development, and apical dominance.[83,90,149,191,192]
Gibberellins (GA)Stem elongation, flowering, and leaf spread.[48,49,150,159,193]
Abscisic acid (ABA)Promote cellular conservation of water under drought and salinity stress.[83,90,193,194]
Salicylic acid (SA)Induces SAR combatting pathogen infection.[78,160,161,166,168,169,177]
Jasmonic acid (JA)Defence against pests through deterrence and elimination of pests.[66,141,142,160,162]
Cytokinin’s (CK)Promotes cell division, apical dominance, lateral root growth.[151,176,177,178,179,180,181,182]
Reactive oxygen species (ROS)Involved in stress signalling and programmed cell death. Major issues with accumulation resulting in toxification.[104,133,140,184,186,187,195,196]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Watts, D.; Palombo, E.A.; Jaimes Castillo, A.; Zaferanloo, B. Endophytes in Agriculture: Potential to Improve Yields and Tolerances of Agricultural Crops. Microorganisms 2023, 11, 1276. https://doi.org/10.3390/microorganisms11051276

AMA Style

Watts D, Palombo EA, Jaimes Castillo A, Zaferanloo B. Endophytes in Agriculture: Potential to Improve Yields and Tolerances of Agricultural Crops. Microorganisms. 2023; 11(5):1276. https://doi.org/10.3390/microorganisms11051276

Chicago/Turabian Style

Watts, Declan, Enzo A. Palombo, Alex Jaimes Castillo, and Bita Zaferanloo. 2023. "Endophytes in Agriculture: Potential to Improve Yields and Tolerances of Agricultural Crops" Microorganisms 11, no. 5: 1276. https://doi.org/10.3390/microorganisms11051276

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