Next Article in Journal
Transcriptome Profiling Reveals Potential Genes Involved in Salicylic Acid-Induced Arbutin Synthesis in Pear
Previous Article in Journal
Comparison of the Climate Change Tolerance of Native and Non-Native Species Used or Potentially Used as Ornamentals in Mediterranean Areas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Harnessing Green Helpers: Nitrogen-Fixing Bacteria and Other Beneficial Microorganisms in Plant–Microbe Interactions for Sustainable Agriculture

by
Luisa Liu-Xu
1,
Ana Isabel González-Hernández
2,
Gemma Camañes
1,
Begonya Vicedo
1,
Loredana Scalschi
1,* and
Eugenio Llorens
1,*
1
Biochemistry and Biotechnology Group, Department of Biology, Biochemistry and Natural Sciences, Universitat Jaume I, 12071 Castellón de la Plana, Spain
2
Superior Polytechnic School of Zamora, University of Salamanca, Avda. de Requejo, 33, 49029 Zamora, Spain
*
Authors to whom correspondence should be addressed.
Horticulturae 2024, 10(6), 621; https://doi.org/10.3390/horticulturae10060621
Submission received: 30 April 2024 / Revised: 2 June 2024 / Accepted: 5 June 2024 / Published: 11 June 2024

Abstract

:
The health of soil is paramount for sustaining life, as it hosts diverse communities of microorganisms that interact with plants, influencing their growth, health, and resilience. Beneficial microorganisms, including fungi and bacteria, form symbiotic relationships with plants, providing essential nutrients, promoting growth, and enhancing stress tolerance. These microorganisms, such as mycorrhizal fungi and plant growth-promoting bacteria, play crucial roles in nutrient cycling, soil health, and plant productivity. Additionally, they help lessen reliance on chemical fertilizers, thereby mitigating the environmental risks associated with their use. Advances in agricultural practices harness the potential of these beneficial microorganisms to improve crop yields while minimizing the environmental impact. However, challenges such as competition with indigenous microbial strains and environmental factors limit the universal utilization of microbial inoculants. Despite these challenges, understanding and leveraging the interactions between plants and beneficial microorganisms hold promise for sustainable agriculture and enhanced food security.

1. Introduction

Someone once said, “Soil is not just dirt, it’s alive, and the health of soil is vital for the health of plants, animals, and people”. And they were right, since soil harbors many communities of microorganisms. These microbial communities interact with plants, both externally and internally, and shape their behavior and health [1]. These interactions can range from pathogenic to mutualistic. Beneficial microorganisms are those that provide the host plant with different benefits, such as enhanced growth and stress resistance, while obtaining sources of nutrition, shelter, and ways of transmission besides air and soil. They are also used for entomological, medicinal, biotechnological, and other industrial purposes. Thus, beneficial microorganisms can represent an important tool for future sustainable agriculture as they can take part in plant nutrition, health, and product quality.
Under natural or agronomic conditions, plants are continuously exposed to ongoing disease pressures caused by pests and pathogens, leading to considerable crop losses. Furthermore, repeated cultivation practices exacerbate nutrient depletion within soil, consequently diminishing both crop yield and quality. Historically, chemical fertilizers and pesticides have been utilized as countermeasures against these challenges. However, prolonged reliance on these chemicals has led to severe health and environmental concerns, including ecological harm and pest resistance. In an effort to mitigate chemical usage, biofertilizers and biopesticides offer a more environmentally sustainable approach to enhancing crop yields [2,3].
Beneficial microorganisms are part of the microbiome, a set of microorganisms that inhabit the interior or surroundings of a living organism and play crucial roles in the health and growth of plants across ecosystems (Figure 1). The importance of the interaction between the microbiome and the host at survival and evolutionary levels has been demonstrated in numerous scientific studies. Such importance suggests considering plant and microbiota as a whole, referred to as a holobiont [4]. Plant-associated microorganisms belong to groups such as archaea, bacteria, and fungi, forming symbiotic relationships aiding in nutrient uptake, disease resistance, and overall resilience [5].
The communities forming the microbial ecosystem will vary depending on the host species, as well as the host’s age, health conditions, and environment. Similarly, we can differentiate as many ecosystems as there are parts of an individual plant. The most important ones are the rhizosphere, the phyllosphere, and the endosphere [8]. The rhizosphere is composed of communities of microorganisms that inhabit both the inner tissues of the root and its surroundings, contributing to soil nutrient acquisition, the conversion of molecules into more readily accessible forms for the plant, the modulation of root pH, and soil texture [5]. The phyllosphere comprises microorganisms capable of adapting to the changing conditions on the plant’s aerial surface, and although they do not usually form communities as rich as those in the rhizosphere, some stable communities can contribute to the plant’s development and defense [9]. Finally, the endosphere includes internal zones such as the plant’s intercellular space or apoplast [10] and comprises microorganisms that reside within the tissues of plants without causing any apparent harm to the host plant [11]. These microorganisms may play an important role in synthesizing essential compounds like amino acids, enhancing leaf growth rates, or boosting photosynthetic efficiency and stomatal conductance [12,13]. They also actively participate in nitrogen fixation and enhance plant defenses [14]. Furthermore, their influence extends to agricultural fields, impacting nutritional and sensory qualities, including flavors and odors [6].
Regarding plant health under abiotic stress, beneficial microorganisms could play an important role against drought, salinity, extreme temperatures, or low-nutrient regimes [15,16]. These functions are performed by several mechanisms, such as N2 fixation and mineral solubilization (phosphorus, potassium, zinc), reducing reliance on chemically synthesized fertilizers while directly supplying plants with macro- and micronutrients, as well as plant growth-promoting hormones [17].
On the other hand, beneficial microorganisms have the ability to suppress diverse phytopathogens through several mechanisms. These include the production of antibiotics, lytic enzymes, and siderophores and competing with pathogens for nutrients and space [18]. Beneficial microorganisms can disrupt pathogens’ quorum sensing (QS) by inhibiting the production of signal molecules through chitinases, pectinases, and lactonases [19]. Some beneficial microbes serve as natural enemies to diverse phytopathogens, exhibit antagonistic behaviors, and compete with pathogenic microorganisms, either inhibiting their growth or outright eliminating them.
Understanding and harnessing the potential of plant-beneficial microorganisms aligns directly with the European Green Deal since they can improve soil health (nutrient cycling and structure) and crop yields and reduce contamination caused by the abuse of chemical inputs. This will promote sustainable farming practices that support the Farm to Fork Strategy and several sustainable development goals (SDGs) by leading to more resilient crop production systems that can help ensure a stable food supply. As such, research into the interactions between plants and beneficial microorganisms is crucial for advancing agricultural resilience and productivity.

2. Beneficial Microorganisms

2.1. Bacteria

Bacteria proliferate in soil, where they form the most abundant and diverse communities. Plant-associated bacteria are primarily found in the rhizosphere, though some can penetrate and colonize plant roots and aerial tissues [20]. These endophytes were initially thought to be plant pathogens, but they were later found to confer several beneficial effects. Plant-associated bacteria exhibit diverse lifestyles and benefits, and classic reviews include [21]. In this review, we focus on the main plant growth-promoting bacteria (PGPB) from the perspective of plant nutrition and development.

2.1.1. Nitrogen-Fixing Bacteria

Nitrogen (N) is one of the main macronutrients needed for the correct growth and development of plants, and therefore is one of the most limiting nutritional factors. The uptake of N by the plant is entirely dependent on the amount present in the soil since plants are unable to assimilate atmospheric N2. Atmospheric N2 must be first reduced to ammonia (NH3) to be assimilable by plants. Fortunately, certain prokaryotic microorganisms, known as diazotrophs, possess the ability to reduce the atmospheric N2 to NH3 in a process known as biological nitrogen fixation (BNF) via the enzymic complex nitrogenase. Considering their life strategies, diazotrophic bacteria can be classified as nodule-forming bacteria or as non-nodular bacteria [22].
The main examples of nodule-forming bacteria are those collectively called rhizobia, which associate with plants of the Leguminosae (=Fabaceae) family. In addition to rhizobia, actinobacteria of the Frankia genus can form nodules and establish symbioses with a diverse group of plants from 23 genera from eight different families belonging to the orders Fagales, Rosales, and Cucurbitales.
Non-nodular bacteria can be found as free-living in the rhizosphere, associated with roots (associative) or inside plant tissues (endophytic). The soil free-living group includes the genera Azotobacter, Bacillus, Beijierinckia, Burkholderia, Clostridium, Desulfovibrio, Derxia, Enterobacter, Klebsiella, Paenibacillus, and Serratia [23,24]. The free-living group also includes cyanobacteria and phototrophic sulfur bacteria. In addition to those in the free-living group, cyanobacteria may live in symbiosis with fungi (forming lichens) or with plants (for example, Nostoc with bryophytes, a few gymnosperms and angiosperms, and Anabaena with the aquatic fern Azolla) [25]. Rhizospheric associative bacteria proliferate on the root surface and nourish on root exudates. The most studied genera of associative diazotrophic bacteria are Acetobacter, Azoarcus, Azospirillum, Burkholderia, and Herbaspirillum [23], but, without question, the genus Azospirillum (with an emphasis on A. brasilense and A. lipoferum) is the most important agronomically [23,26,27] and is marketed as inoculants in many countries. Endophytic bacteria invade and rapidly proliferate in inner plant parts during at least a part of their life cycle, without causing symptoms of plant disease [23]. It is considered that nitrogen-fixing endophytic bacteria have some advantages over rhizospheric associative bacteria since, being protected inside plant tissues, they have less competition and can make fixed nitrogen directly available to plants. Moreover, efficient biological nitrogen fixation requires low levels of oxygen, which are easily found inside plant tissue [28]. Acetobacter diazotrophicus (=Gluconobacter diazotrophicus) was one of the first endophytic diazotrophic bacteria to be studied with emphasis on sugarcane [29]. Moreover, Herbaspirillum is usually found as endophytic diazotrophic bacteria, particularly in grasses, such as sugarcane, rice and wheat, and pastures. Herbaspirillum seropedicae is considered to be a diazotrophic bacteria model for endophytic associations [30]. Until now, N2-fixing endophytes have been mainly isolated from monocots, from a varied number of species such as kallar grass [31], sugarcane [32], rice [33], maize [34], wheat [33], Sorghum halepense [35], miscanthus [36], and elephant grass [37]. There are also some reports of endophytic diazotrophic bacteria in mosses [38] and in conifers [39,40]. However, little is known about their presence in dicots when compared with monocots. In dicots, N2-fixing endophytes have been isolated from sweet potato [41] and coffee plants [42]. Moreover, Kandel et al. [43] isolated diatrotrophic endophytes from poplar and willow and tested them in rice, while Wozniak et al. [44] isolated diatrotrophic endophytes from Arctium lappa belonging to the Asteraceae family. A list of the endophytic diazotrophic bacteria isolated and the agricultural crops in which they were colonized into is provided in the work by Puri et al. [45] and Rana et al. [46].
All these bacteria hold the potential to improve plant performance by enhancing the availability of nutrients and to improve soil fertility, mainly through BNF and also through other features such as phosphate solubilization or iron sequestration by siderophores [27,47,48]. In addition, they produce plant growth regulators and are involved in the modulation of phytohormone and defense responses, the production of antioxidants, osmotic adjustment, and plant tolerance against biotic and abiotic stresses, as reviewed in detail by Thiebaut et al. [49].

2.1.2. Phosphate- and Potassium-Solubilizing Bacteria

Another essential macronutrient required for plant growth and metabolism is phosphorus (P), which is often unavailable for plant uptake [50]. Bacteria that convert insoluble organic and inorganic substances into soluble forms of phosphate that are easily absorbed by plants through various mechanisms are called phosphate-solubilizing bacteria (PSB) [51]. The involved mechanisms include the secretion of enzymes (phosphatase and phytolase), acids (both organic and inorganic), and chelation (siderophores and extracellular polysaccharides) (reviewed by Pan and Cai [52]). The most abundant PSB genera are Bacillus, Burkholderia, Pseudomonas, and Pseudocystis, and to a lesser extent, Enterobacter, Salmonella, Flavobacterium, Micrococcus, Thiobacillus, Azotobacter, Pantoea, Serratia, and Baeyerlingia, among others [52]. Several studies have investigated the effectiveness of PSB on P solubilization and crop performance. For example, plant-unavailable P can be dissolved in either inorganic (calcium phosphate) or organic (phytin) forms by different bacteria, such as Pseudomonas pseudoalcaligenes and Bacillus thuringiensis [53]. Moreover, Habibi et al. [54] isolated Paenibacillus, Pseudomonas, and Enterobacter from rice plants, which were able to dissolve tricalcium phosphate, one of the unavailable forms of P. PSB have the capacity to modify phosphorus speciation on their own, but they can also work in concert with phosphate fertilizers to increase phosphorus use [55].
Potassium (K) is another of the essential nutrients, a deficiency of which in soil can impair plant growth, cause disruptions to plant activities, and lower the quality of fruit [56]. The rhizospheric bacteria responsible for converting insoluble potassium (K) into soluble forms that promote plant development and yield are known as potassium solubilizing bacteria (KSB) [57]. Furthermore, several authors highlighted the role of KSB in releasing K from mineral sources such as feldspar and aluminosilicate minerals through various mechanisms, including acidolysis, chelation, exchange reactions, and organic matter and crop residue decomposition [58]. Among others, Bacillus mucilaginosus, Bacillus circulans, Bacillus cereus, Bacillus edaphicus, and Pseudomonas sp. are considered KSB.

2.2. Fungi

Beneficial or symbiotic fungi exist in almost all plants [25] mainly involving mycorrhizal, epiphytic, or endophytic lifestyles. These fungi interact with plants and other organisms, establishing complex and dynamic systems where multipartite interactions often occur [59,60]. Mycorrhiza might be able to associate with bacterial communities to colonize the plant roots, converting into an endophytic lifestyle [61]. Other studies discuss if all fungi might have past endophytic behavior [62]. Although the complexity of these interactions may seem daunting, in this review, we will focus on the main categories studied for agronomic progress.

2.2.1. Mycorrhizal Fungi

The term mycorrhiza was introduced as early as 1885 by Frank [63], as a fungus–root symbiosis that occurs in the rhizosphere, a zone rich in microbial activity. This relationship enhances nutrient availability and influences plant health. The fungus provides water and nutrients like phosphate and nitrogen to the plant, while the plant supplies carbohydrates and other organic metabolites to the fungus [64,65,66].
Mycorrhizal fungi play important roles in soil biology and chemistry, and most terrestrial plants have associations with them, including many crop species [64,67]. On the other hand, some exceptions include several species of Brassicaceae [68]. In environments in which plants do not require assistance in obtaining water and nutrients, they do not form these associations, and in some conditions like excessive plant stress, mycorrhizal fungi can turn parasitic with plants [69]. Mycorrhizae can be classified based on how they colonize plant roots. These fungi have developed diverse strategies of colonization, as well as different degrees of plant dependence. These include facultative biotrophic ectomycorrhizal fungi (EMF) and the most common obligate biotrophic arbuscular mycorrhizal fungi (AMF), also known as endomycorrhizal fungi [64]. Arbuscular mycorrhiza is one of the oldest interactions on Earth and was considered crucial for plant evolution on land [70,71,72]. The main AMF species belong to the phylum Glomeromycota, order Glomerales [65], and have been reported to improve plant health and nutrition as well as resistance to stress [66,73]. However, AMF are complex organisms, and studying their biology and use in agriculture was challenging until recently thanks to the advances in computational tools and omics data [74].
Many AMF are known to boost plant growth and yield by forming extensive hyphal networks and producing substances like glomalin. These enhance water and nutrient uptake and improve soil structure [75]. AMF is also known to increase the activity of antioxidant enzymes to counteract the negative effects of reactive oxygen species generated by salinity stress [76]. Some examples include the improvement of drought tolerance [77,78,79], salinity tolerance by species of the Glomus genus [80], zinc-induced stress [81], weed management, or pathogen and nematode resistance [82]. In comparison, EMF are not as well characterized [83].
The interaction between a host plant and mycorrhizal fungi is regulated by signaling mechanisms. Some insightful reviews which delve into the details of mycorrhizal fungi colonization and symbiosis accommodation can be found in [84,85,86,87]. In brief, plants secrete root exudates into the rhizosphere to alter the soil properties and shape the microbial community [88,89,90]. For instance, AMF spores detect the host plant through biochemical signals that include a class of plant hormones known as strigolactones. Then, AMF release signaling molecules that trigger root responses [65] and extend hyphae into the host’s root cortical cells, forming highly branched arbuscules for nutrient exchange and vesicles for storage. Interestingly, the plant’s responses to mycorrhizal colonization are similar to the molecular events in other interactions, including ones with pathogenic microorganisms [71].

2.2.2. Endophytic Fungi

Fungal endophytes have recently gained considerable interest [91]. Modern molecular and genomic techniques have facilitated significant progress in their discovery and characterization, and it is now believed that fungal endophytes may be as abundant as mycorrhizal fungi [92]. These fungi can establish symbiosis with their host plants, yet the relationship needs to be properly characterized to distinguish them from saprobes or latent plant pathogens [93].
Endophytic fungi have remarkable diversity in their lifestyles. They can be found in a wide range of plant species and can be found in any tissue (roots, stems, leaves, and seeds). They may colonize plant tissues locally or systemically, intra- or intercellularly [11]. Some valuable reviews which introduce fungal endophytes can be found in [93,94,95]. The modern study of fungal endophytes dates to the discovery of the Epichloë genus (family Clavicipitaceae), which extensively colonizes Poaceae grasses such as ryegrass (Lolium spp.) [96,97]. Similar to endophytic bacteria, they were initially thought to be detrimental since they produced mycotoxins (alkaloids) that affected grazing cattle [98]. However, these endophytes provided the plants higher resistance to pests and other stresses [99], and many studies were performed to find and use these endophytes for their benefits while reducing the potentially detrimental effects.
Most endophytes are ascomycetous fungi as they are the most ubiquitous in nature. Some well-known examples include species from genera such as Fusarium [100], Cladosporium [101], Penicillium [102], Trichoderma [103], or Aspergillus [104]. On the other hand, Basidiomycota and basal fungi are less studied due to their difficult identification and cultivation [66]. Some endophytes from these categories include the most outstanding Serendipita indica [105] or Streptomyces spp. [106]. Interestingly, some studies use the term Dark Septate Endophytes (DSEs) to describe a diverse group of ascomycetous endophytes with melanin hyphae that colonize root tissues [68], including some Alternaria spp. or Penicillium spp. Another commonly used term to categorize some beneficial fungal endophytes is plant growth-promoting fungi (PGPF) [107,108].
Endophytic fungi can establish contact with plants either through the soil or via airborne mechanisms (horizontal transmission). In addition, they can be transmitted through plant seeds to the next plant generations (vertical transmission) [109,110]. To colonize the plant tissues, the beneficial endophytic fungi may be able to evade the plant’s immune system [110] and invade tissues using penetration structures or by producing enzymes for the degradation of the cell walls [11,111]. The specific molecular events that enable a beneficial interaction instead of a pathogenic one are still being elucidated [112,113,114]. Interestingly, some endophytic fungi form relationships that are similar to mycorrhizal fungi by exchanging nutrients and enhancing plant vigor. In this way, Serendipita indica is a species that has been reported to have different behaviors that range from endophytic to mycorrhizal [115] and is a model system due to its wide range of host plants and benefits [116].
Endophytic fungi interact with their host plants in several ways. They can promote plant growth by enhancing nutrient uptake, producing growth-promoting substances, and modulating plant hormone levels. Some fungal endophytes, such as Metarhizium and Beauveria, are also able to transfer nitrogen to their plant hosts [117]. Endophytic fungi may also modulate the plant’s immune system, thereby improving its ability to withstand abiotic and biotic stresses. For instance, they can protect from pathogens and herbivores [118] by serving as a physical barrier against pathogens or producing bioactive compounds.
Additionally, fungal endophytes are a rich source of secondary metabolites that might be of human interest. Fungal endophytes, especially those from medicinal plants, have been extensively explored as potential bioactive compounds for pharmaceutical, industrial, and agricultural applications [11,119,120,121]. Interestingly, many bioactive compounds that were previously thought to be plant metabolites have been recently discovered as endophyte ones [74,83].

3. Benefits of Microorganisms in Agriculture

3.1. Advances in Agriculture

Beneficial microorganisms are essential tools for sustainable agriculture due to their multifaceted roles in enhancing soil health, promoting plant growth, and ensuring crop productivity (Figure 2). In addition, they can help mitigate climate change through their roles in carbon sequestration, reductions in greenhouse gas emissions, and the bioremediation of contaminated soils. Nevertheless, in this review, we focus on agronomic ecosystems and how the synergistic interactions between soil microbes and plants contribute to improving soil fertility and crop productivity in several ways.
Many beneficial microorganisms have been shown to improve plant health in a wide range of plant species. Conversely, other microorganisms are restricted to a very narrow range of hosts, such as the genus Epichloë, whose different species can only colonize a small range of hosts [122]. Their significance extends beyond mere nutrient cycling to encompass a range of actions that contribute to the resilience and productivity of agricultural ecosystems. For example, experiments in Mexico using the enrichment of 15N over five years revealed that atmospheric nitrogen fixation contributed significantly to maize’s nitrogen nutrition, ranging from 29% to 82% [123]. Similarly, Rose et al. showed that a commercial biofertilizer could replace 23% to 52% of nitrogen chemical fertilizers without reducing rice yield in Southeast Asia. In sugarcane plants, Serna-Cock et al. [124] demonstrated that the application of Azospirillum brasilense, Azotobacter chroccocum, and Trichoderma lignorum can effectively substitute 60% of the required nitrogen for this crop. Supporting these findings, Antunes et al. [125] demonstrated that the inoculation with Herbaspirillum seropedicae, Pseudomonas sp., and Bacillus megaterium led to a notable increase in sugarcane yield, ranging from 18% to 57.31%.
Regarding the solubilization of nutrients, the utilization of PSB as bioinoculants enhances P availability through various strategies [126,127,128]. Specifically, PSB has the ability to convert insoluble forms of P into soluble orthophosphate, effectively dissociating fixed P pools accumulated from indiscriminate P fertilizer use. Combining PSB with P fertilizers has shown synergistic effects, particularly enhancing the agronomic efficiency of P rock in soils with high P retention [129,130].
Recent studies have highlighted the benefits of PSB inoculation in various crop systems. Adnan et al. [110] demonstrated significant improvement in maize growth with PSB inoculation, especially when combined with different types of inorganic and organic P fertilizers. Similarly, Javeed et al. [131] observed enhanced maize growth parameters with PSB inoculation in conjunction with NPK fertilizer. Suleman et al. [132] reported increased P uptake and use efficiency in wheat plants inoculated with Pseudomonas sp. or Enterobacter sp. alongside diammonium phosphate fertilizer. Shirmohammadi et al. [133] found substantial increases in grain yield and plant growth parameters in rainfed wheat with the single or co-inoculation of Pseudomonas baetica and Pseudomonas helmanticensis, particularly when fertilized with triple superphosphate.
Furthermore, studies have shown that PSB consortia have a more pronounced effect on P uptake by plants. Duarah et al. [134] demonstrated higher rice plant growth with a mixture of PSB strains and NPK fertilizer, while Barra et al. [135] found enhanced ryegrass growth due to better P acquisition from Triple Super Phosphate fertilizer (TSP) with a PSB consortium. The co-inoculation of Vicia faba plants with specific bacterial strains significantly increased biomass and P content under different P supplies [48]. These findings underscore the agronomic potential of efficient PSB strains, especially when used in consortia, to enhance P use efficiency and plant growth across various cropping systems.
On the other hand, the ability of mycorrhizal fungi to improve P uptake is based on increasing absorption surface area and solubilizing soil P [136]. AMF enhanced phosphorus uptake during fast-growing stages in maize, contributing up to 19.4% of the total available soil P and significantly increasing yield [137]. Interestingly, Zhang et al. [138] have reported that AMF recruit bacteria that are able to solubilize P instead of directly affecting the P in the soil. Some soil fungi interact synergistically with rhizobacteria to enhance plant growth and nutrient acquisition. For instance, Bouhraoua et al. [139] demonstrated that inoculation with certain PSB strains, such as Pseudomonas sp., was correlated with AMF colonization, and this combination improved NPK uptake in peanut plants by up to 200%. AMF can also facilitate colonization by symbiotic bacteria, as seen in the work by Barreto de Novais et al. [140], who revealed how AMF Glomus formosanum facilitates the transfer of N-fixing Bradyrhizobium diazoefficiens in Glycine max roots. Furthermore, bacteria of the genus Frankia were studied as co-inoculants with several EMF in Alnus viridis under poor-nutrient soil by Chen et al. [141] and were proved to benefit both fungal and plant growth.
In addition to plant development, beneficial microorganisms can play a crucial role in protecting host plants from infections. The biocontrol bacteria Bacillus and Pseudomonas spp. along with the fungal genera Trichoderma, Aspergillus, and Penicillium are among the most popular biocontrol agents against both bacterial and fungal plant diseases in major crops [142]. For instance, inoculation with Bacillus strains such as B. subtilis or B. amyloliquefaciens have been proven to confer resistance against pathogens such as Botrytis cinerea in strawberry [143], Ralstonia solani in cowpea [144], and Sclerotium rolfsii in peanut [145].
Antibiotic compounds are known to be produced by bacteria such as Actinomycetes and fungi such as Penicillium [18,19]. These can include volatiles such as HCN, phenazines, pyrrolnitrin, siderophores, and more [146]. As an example, Kang et al. [147] found that HCN production in Pseudomonas chlororaphis was key in the nematicidal activity against Meloidogyne hapla in tomato plants.
Alternatively, the arsenal of antibacterial compounds produced by fungal endophytes can range from broad-spectrum to more specific agents [148]. It has been reported that endophytes generate a plethora of secondary metabolites with broad antimicrobial activity, including terpenoids, alkaloids, phenylpropanoids, aliphatic compounds, polyketides, acetol, hexanoic acid, acetic acid, and peptides [149,150]. For instance, javanicin, a compound with potent antimicrobial properties, exhibits effectiveness against a wide range of microbes, with particular efficacy against Bacillus spp. and Escherichia coli [148]. Notably, terpenoids like Phomadecalin E and 8α-acetoxyphomadecalin C, synthesized by some Microdiplodia endophytes, demonstrate potent antibacterial effects against strains of Pseudomonas aeruginosa [149]. Extracts from Trichoderma harzianum, an endophyte isolated from Salvia rosmarinus, exhibit significant antimicrobial activity against P. aeruginosa, Staphylococcus aureus, Klebsiella pneumoniae, Bacillus subtilis, and Escherichia coli, suggesting its potential as a biocontrol agent against phytopathogenic bacteria [151]. Additionally, Diaporthe phaseolorum, Aspergillus fumigatus, and A. versicolor, isolated as endophytes from healthy tomato (Solanum lycopersicum) plants, produce antibacterial metabolites like acetol, hexanoic acid, and acetic acid, which effectively combat bacterial spot disease caused by Xanthomonas vesicatoria [152]. These antimicrobial compounds may originate directly from the endophytic fungus or be induced by the host plant following endophyte inoculation [149]. However, our understanding of the secretion of these compounds and associated gene expression is still limited.
Endophytes contribute to host plant resistance against fungal pathogens through systemic responses, including strengthening cell walls and inducing defensive gene expression [153,154,155,156]. They also enhance photosynthesis rates, chlorophyll content, and the production of defensive compounds like phytoalexins [153,155,157]. Competitive exclusion, where endophytes outcompete pathogens for niche space, is another protective mechanism [154].
Certain fungal endophytes, such as those from the genus Daldinia and Fusarium, have demonstrated inhibitory effects on plant pathogens [136,137,138]. Acremonium sclerotigenum, isolated from an ancestral wheat species, is capable of inducing resistance in both wheat and tomato [158]. Moreover, species from genera like Trichoderma, Aspergillus, Colletotrichum, and Diaporthe show promising antifungal properties [159,160]. Overall, the exploration of fungal endophytes and their activities against fungal pathogens offers valuable insights into developing sustainable agricultural practices while minimizing environmental impacts. Further understanding of these interactions will facilitate the development of effective biocontrol strategies and the identification of novel antifungal compounds for agricultural applications.

3.2. Limiting Factors

Despite their potential to reduce fertilizer dependency and contribute to agricultural sustainability, microbial inoculants face several challenges that limit their universal utilization.
The efficacy of biofertilizers depends on various factors, including the targeted crop, soil conditions (such as pH, salinity, and type), competition with indigenous microbial strains, and climatic factors. Additionally, microbial strains used in commercial inoculants often do not fully represent the diversity of strains associated with plants under field conditions, leading to functional limitations; for example, while cereals derive significant benefits from diazotrophs, most microbes are unlikely to fix nitrogen in the presence of high rates of chemical fertilizers [161].
Phosphate-solubilizing microorganisms, for instance, play a particularly crucial role in soils facing extreme environmental conditions such as saline–alkaline or nutrient-deficient environments. However, reports vary regarding the optimal temperature for phosphate solubilization, ranging from 20–25 °C to 30 °C [162,163,164]. Some studies even demonstrate solubilization at extreme temperatures, from 45 °C in desert soil to as low as 10 °C [165]. Various factors influence microbial phosphate solubilization, including soil microbial interactions, plant growth stage, environmental conditions, soil types, and agronomic practices. Both aerobic and anaerobic bacteria contribute to phosphate solubilization, with different genera exhibiting varying preferences. Additionally, the presence of exogenous soluble phosphate affects the physiological activity of PSB. While low levels of soluble phosphate stimulate microbial growth and phosphorus nutrition of plants, high concentrations can suppress PSB activity, posing a barrier to their widespread use [166].
The application of microbial inoculants must also be carefully optimized. For seeds, proper seed coating is essential to ensure better contact and inoculation rates. For soil applications, combining the inoculants with soil amendments can enhance their efficacy [167]. Lastly, another current limitation is the lack of well-developed inoculant formulations, since they require extensive exploration before reaching the market. Moreover, compliance with constantly evolving regulations poses a challenge, hindering the utilization of existing beneficial microorganisms. However, as on-field experiments increasingly demonstrate the benefits of microorganisms, we anticipate the greater use of these sustainable alternatives in the near future.

4. Conclusions and Future Prospects

Plant–microorganism interactions represent a promising avenue for advancing agriculture and food security while minimizing the environmental impact caused by chemical fertilizers and pesticides. Beneficial microorganisms are essential in this quest, forming symbiotic relationships with plants to improve nutrient cycling, soil health, and plant resilience against various stresses. From mycorrhizal fungi extending the reach of plant roots for nutrient uptake to nitrogen-fixing bacteria enhancing nutrient availability, the multifaceted functions of these microorganisms offer many alternatives for addressing the challenges of modern agriculture. Integrating microbial inoculants into precision farming practices can optimize resource use and crop performance. Additionally, expanding these technologies to a broader range of crops and adapting them to diverse climates and soils will further their application and benefits.
Supportive regulatory frameworks and policies are crucial for widespread adoption, including developing standards and guidelines for microbial inoculants’ production and application. Public awareness and education initiatives are essential to inform farmers and consumers about the benefits of these technologies for sustainable agriculture.
However, to fully harness the potential of beneficial microorganisms in agricultural systems, we must address key aspects of microbial ecology. Despite the significant strides that have been made in understanding their mechanisms of action and agronomic applications, challenges such as competition with indigenous microbial strains and environmental variability persist. To overcome these barriers, interdisciplinary approaches that integrate microbiology, agronomy, and environmental science are required to optimize the efficacy and sustainability of microbial inoculants.
In conclusion, by promoting the natural symbiotic relationships that have evolved over millennia, we expect to reduce the reliance on chemical inputs while we enhance soil health and crop yields in a way that is environmentally sound and economically viable. As we continue to unlock the secrets of plant–microbe interaction, we will strive towards future agriculture that is both resilient and sustainable.

Author Contributions

Conceptualization, E.L. and L.S.; writing—review and editing, L.L.-X., G.C., B.V. and A.I.G.-H.; supervision, E.L. and L.S.; funding acquisition, B.V. and G.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Diputación de Castellón, Project “Fijación de nitrógeno en cítricos” 24I094 and Universitat Jaume I code UJI-B2022-30.

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 conflicts of interest.

References

  1. Compant, S.; Samad, A.; Faist, H.; Sessitsch, A. A review on the plant microbiome: Ecology, functions, and emerging trends in microbial application. J. Adv. Res. 2019, 19, 29–37. [Google Scholar] [CrossRef] [PubMed]
  2. Misra, M.; Sachan, A.; Sachan, S.G. Current Aspects and Applications of Biofertilizers for Sustainable Agriculture. In Plant Microbiomes for Sustainable Agriculture; Springer: New York, NY, USA, 2020; pp. 445–473. [Google Scholar]
  3. Sarkar, D.; Rakshit, A.; Al-Turki, A.I.; Sayyed, R.Z.; Datta, R. Connecting Bio-Priming Approach with Integrated Nutrient Management for Improved Nutrient Use Efficiency in Crop Species. Agriculture 2021, 11, 372. [Google Scholar] [CrossRef]
  4. Zilber-Rosenberg, I.; Rosenberg, E. Role of microorganisms in the evolution of animals and plants: The hologenome theory of evolution. FEMS Microbiol. Rev. 2008, 32, 723–735. [Google Scholar] [CrossRef] [PubMed]
  5. Santoyo, G. How plants recruit their microbiome? New insights into beneficial interactions. J. Adv. Res. 2022, 40, 45–58. [Google Scholar] [CrossRef] [PubMed]
  6. Johnston-Monje, D.; Vergara, L.I.; Lopez-Mejia, J.; White, J.F. Plant microbiomes as contributors to agricultural terroir. Front. Agron. 2023, 5, 1216520. [Google Scholar] [CrossRef]
  7. Dhaliwal, S.S.; Sharma, V.; Mandal, A.; Naresh, R.K.; Verma, G. Improving soil micronutrient availability under organic farming. In Advances in Organic Farming: Agronomic Soil Management Practices; Woodhead Publishing: Cambridge, UK, 2021; pp. 93–114. [Google Scholar]
  8. Dastogeer, K.M.G.; Tumpa, F.H.; Sultana, A.; Akter, M.A.; Chakraborty, A. Plant microbiome–an account of the factors that shape community composition and diversity. Curr. Plant Biol. 2020, 23, 100161. [Google Scholar] [CrossRef]
  9. Vorholt, J.A.; Vogel, C.; Carlström, C.I.; Müller, D.B. Establishing Causality: Opportunities of Synthetic Communities for Plant Microbiome Research. Cell Host Microbe 2017, 22, 142–155. [Google Scholar] [CrossRef] [PubMed]
  10. Farvardin, A.; González-hernández, A.I.; Llorens, E.; García-agustín, P.; Scalschi, L.; Vicedo, B. The apoplast: A key player in plant survival. Antioxidants 2020, 9, 604. [Google Scholar] [CrossRef] [PubMed]
  11. Schulz, B.; Boyle, C. The endophytic continuum. Mycol. Res. 2005, 109, 661–686. [Google Scholar] [CrossRef]
  12. Eroğlu, G.; Cabral, C.; Ravnskov, S.; Bak Topbjerg, H.; Wollenweber, B. Arbuscular mycorrhiza influences carbon-use efficiency and grain yield of wheat grown under pre- and post-anthesis salinity stress. Plant Biol. 2020, 22, 863–871. [Google Scholar] [CrossRef]
  13. Liu, Y.; Hou, W.; Jin, J.; Christensen, M.J.; Gu, L.; Cheng, C.; Wang, J. Epichloë gansuensis increases the tolerance of achnatherum inebrians to low-p stress by modulating amino acids metabolism and phosphorus utilization efficiency. J. Fungi 2021, 7, 390. [Google Scholar] [CrossRef] [PubMed]
  14. del Orozco-Mosqueda, M.C.; Santoyo, G. Plant-microbial endophytes interactions: Scrutinizing their beneficial mechanisms from genomic explorations. Curr. Plant Biol. 2021, 25, 100189. [Google Scholar] [CrossRef]
  15. Thamer, S.; Schädler, M.; Bonte, D.; Ballhorn, D.J. Dual benefit from a belowground symbiosis: Nitrogen fixing rhizobia promote growth and defense against a specialist herbivore in a cyanogenic plant. Plant Soil 2011, 341, 209–219. [Google Scholar] [CrossRef]
  16. Prakash, J.; Mishra, S. Role of beneficial soil microbes in alleviating climatic stresses in plants. In Microbiome under Changing Climate: Implications and Solutions; Woodhead Publishing: Cambridge, UK, 2022; pp. 29–68. [Google Scholar]
  17. Soni, S.K.; Manhas, R.; Jakhar, Y.; Sharma, A.; Soni, R. Biofertilizers for Sustainable Agriculture: Current Trends and Future Perspective. In Genomic, Proteomics, and Biotechnology; CRC Press: Boca Raton, FL, USA, 2022; pp. 331–356. [Google Scholar]
  18. Marasco, R.; Rolli, E.; Ettoumi, B.; Vigani, G.; Mapelli, F.; Borin, S.; Abou-Hadid, A.F.; El-Behairy, U.A.; Sorlini, C.; Cherif, A.; et al. A Drought Resistance-Promoting Microbiome Is Selected by Root System under Desert Farming. PLoS ONE 2012, 7, e48479. [Google Scholar] [CrossRef] [PubMed]
  19. Saeki, E.K.; Kobayashi, R.K.T.; Nakazato, G. Quorum sensing system: Target to control the spread of bacterial infections. Microb. Pathog. 2020, 142, 104068. [Google Scholar] [CrossRef] [PubMed]
  20. Compant, S.; Clément, C.; Sessitsch, A. Plant growth-promoting bacteria in the rhizo- and endosphere of plants: Their role, colonization, mechanisms involved and prospects for utilization. Soil Biol. Biochem. 2010, 42, 669–678. [Google Scholar] [CrossRef]
  21. Hayat, R.; Ali, S.; Amara, U.; Khalid, R.; Ahmed, I. Soil beneficial bacteria and their role in plant growth promotion: A review. Ann. Microbiol. 2010, 60, 579–598. [Google Scholar] [CrossRef]
  22. Santos, M.S.; Nogueira, M.A.; Hungria, M. Microbial inoculants: Reviewing the past, discussing the present and previewing an outstanding future for the use of beneficial bacteria in agriculture. AMB Express 2019, 9, 205. [Google Scholar] [CrossRef]
  23. Baldani, J.I.; Baldani, V.L.D. History on the biological nitrogen fixation research in graminaceous plants: Special emphasis on the Brazilian experience. An. Acad. Bras. Cienc. 2005, 77, 549–579. [Google Scholar] [CrossRef]
  24. da Silva, K.; Nóbrega, R.S.A.; Lima, A.S.; Barberi, A.; de Moreira, F.M.S. Density and diversity of diazotrophic bacteria isolated from Amazonian soils using N-free semi-solid media. Sci. Agric. 2011, 68, 518–525. [Google Scholar] [CrossRef]
  25. Kaschuck, G.; Hungria, M. Diversity and importance of diazotrophic bacteria to agricultural sustainability in the tropics. In Diversity and Benefits of Microorganisms from the Tropics; Springer International Publishing: New York, NY, USA, 2017; pp. 269–292. [Google Scholar]
  26. Hungria, M.; Nogueira, M.A.; Araujo, R.S. Co-inoculation of soybeans and common beans with rhizobia and azospirilla: Strategies to improve sustainability. Biol. Fertil. Soils 2013, 49, 791–801. [Google Scholar] [CrossRef]
  27. Hungria, M.; Nogueira, M.A.; Araujo, R.S. Inoculation of Brachiaria spp. with the plant growth-promoting bacterium Azospirillum brasilense: An environment-friendly component in the reclamation of degraded pastures in the tropics. Agric. Ecosyst. Environ. 2016, 221, 125–131. [Google Scholar] [CrossRef]
  28. Cocking, E.C. Endophytic colonization of plant roots by nitrogen-fixing bacteria. Plant Soil 2003, 252, 169–175. [Google Scholar] [CrossRef]
  29. Siqueira, J.O. Inter-Relação Fertilidade, Biologia do solo e Nutrição de Plantas; Sociedade Brasileira de Ciência do Solo: Lavras, Brazil, 1999. [Google Scholar]
  30. Monteiro, R.A.; Balsanelli, E.; Wassem, R.; Marin, A.M.; Brusamarello-Santos, L.C.C.; Schmidt, M.A.; Tadra-Sfeir, M.Z.; Pankievicz, V.C.S.; Cruz, L.M.; Chubatsu, L.S.; et al. Herbaspirillum-plant interactions: Microscopical, histological and molecular aspects. Plant Soil 2012, 356, 175–196. [Google Scholar] [CrossRef]
  31. Hurek, T.; Handley, L.L.; Reinhold-Hurek, B.; Piché, Y. Azoarcus grass endophytes contribute fixed nitrogen to the plant in an unculturable state. Mol. Plant-Microbe Interact. 2002, 15, 233–242. [Google Scholar] [CrossRef] [PubMed]
  32. Welbaum, G.E.; Sturz, A.V.; Dong, Z.; Nowak, J. Managing soil microorganisms to improve productivity of agro-ecosystems. Crit. Rev. Plant Sci. 2004, 23, 175–193. [Google Scholar] [CrossRef]
  33. Liu, H.; Zhang, L.; Meng, A.; Zhang, J.; Xie, M.; Qin, Y.; Faulk, D.C.; Zhang, B.; Yang, S.; Qiu, L. Isolation and molecular identification of endophytic diazotrophs from seeds and stems of three cereal crops. PLoS ONE 2017, 12, e0187383. [Google Scholar] [CrossRef] [PubMed]
  34. Montañez, A.; Abreu, C.; Gill, P.R.; Hardarson, G.; Sicardi, M. Biological nitrogen fixation in maize (Zea mays L.) by 15N isotope-dilution and identification of associated culturable diazotrophs. Biol. Fertil. Soils 2009, 45, 253–263. [Google Scholar] [CrossRef]
  35. Rout, M.E.; Chrzanowski, T.H. The invasive Sorghum halepense harbors endophytic N2-fixing bacteria and alters soil biogeochemistry. Plant Soil 2009, 315, 163–172. [Google Scholar] [CrossRef]
  36. Davis, S.C.; Parton, W.J.; Dohleman, F.G.; Smith, C.M.; Del Grosso, S.; Kent, A.D.; DeLucia, E.H. Comparative biogeochemical cycles of bioenergy crops reveal nitrogen-fixation and low greenhouse gas emissions in a Miscanthus × giganteus agro-ecosystem. Ecosystems 2010, 13, 144–156. [Google Scholar] [CrossRef]
  37. Videira, S.S.; de Pereira e Silva, M.C.; de Souza Galisa, P.; Dias, A.C.F.; Nissinen, R.; Divan, V.L.B.; van Elsas, J.D.; Baldani, J.I.; Salles, J.F. Culture-independent molecular approaches reveal a mostly unknown high diversity of active nitrogen-fixing bacteria associated with Pennisetum purpureum-a bioenergy crop. Plant Soil 2013, 373, 737–754. [Google Scholar] [CrossRef]
  38. DeLuca, T.H.; Zackrisson, O.; Gundale, M.J.; Nilsson, M.C. Ecosystem feedbacks and nitrogen fixation in boreal forests. Science 2008, 320, 1181. [Google Scholar] [CrossRef] [PubMed]
  39. Carrell, A.A.; Frank, A.C. Pinus flexilis and Picea engelmannii share a simple and consistent needle endophyte microbiota with a potential role in nitrogen fixation. Front. Microbiol. 2014, 5, 333. [Google Scholar] [CrossRef]
  40. Moyes, A.B.; Kueppers, L.M.; Pett-Ridge, J.; Carper, D.L.; Vandehey, N.; O’Neil, J.; Frank, A.C. Evidence for foliar endophytic nitrogen fixation in a widely distributed subalpine conifer. New Phytol. 2016, 210, 657–668. [Google Scholar] [CrossRef] [PubMed]
  41. Terakado-Tonooka, J.; Ohwaki, Y.; Yamakawa, H.; Tanaka, F.; Yoneyama, T.; Fujihara, S. Expressed nifH Genes of Endophytic Bacteria Detected in Field-Grown Sweet Potatoes (Ipomoea batatas L.). Microbes Environ. 2008, 23, 89–93. [Google Scholar] [CrossRef] [PubMed]
  42. Jimenez-Salgado, T.; Fuentes-Ramirez, L.E.; Tapia-Hernandez, A.; Mascarua-Esparza, M.A.; Martinez-Romero, E.; Caballero-Mellado, J. Coffea arabica L., a new host plant for Acetobacter diazotrophicus, and isolation of other nitrogen-fixing acetobacteria. Appl. Environ. Microbiol. 1997, 63, 3676–3683. [Google Scholar] [CrossRef] [PubMed]
  43. Kandel, S.L.; Herschberger, N.; Kim, S.H.; Doty, S.L. Diazotrophic Endophytes of Poplar and Willow for Growth Promotion of Rice Plants in Nitrogen-Limited Conditions. Crop Sci. 2015, 55, 1765–1772. [Google Scholar] [CrossRef]
  44. Woźniak, M.; Gałązka, A.; Tyśkiewicz, R.; Jaroszuk-ściseł, J. Endophytic bacteria potentially promote plant growth by synthesizing different metabolites and their phenotypic/physiological profiles in the biolog gen iii microplateTM test. Int. J. Mol. Sci. 2019, 20, 5283. [Google Scholar] [CrossRef]
  45. Puri, A.; Padda, K.P.; Chanway, C.P. Nitrogen-Fixation by Endophytic Bacteria in Agricultural Crops: Recent Advances. In Nitrogen in Agriculture—Updates; InTech: Rijeka, Croatia, 2018. [Google Scholar]
  46. Rana, K.L.; Kour, D.; Kaur, T.; Devi, R.; Yadav, A.N.; Yadav, N.; Dhaliwal, H.S.; Saxena, A.K. Endophytic microbes: Biodiversity, plant growth-promoting mechanisms and potential applications for agricultural sustainability. Antonie van Leeuwenhoek 2020, 113, 1075–1107. [Google Scholar] [CrossRef]
  47. Hungria, M.; Campo, R.J.; Souza, E.M.; Pedrosa, F.O.; Hungria, M.; Campo, R.J.; Souza, E.M.; Pedrosa, F.O. Inoculation with selected strains of Azospirillum brasilense and A. lipoferum improves yields of maize and wheat in Brazil. Plant Soil 2010, 331, 413–425. [Google Scholar] [CrossRef]
  48. Bechtaoui, N.; Raklami, A.; Benidire, L.; Tahiri, A.I.; Göttfert, M.; Oufdou, K. Effects of PGPR co-inoculation on growth, phosphorus nutrition and phosphatase/phytase activities of faba bean under different phosphorus availability conditions. Pol. J. Environ. Stud. 2020, 29, 1557–1565. [Google Scholar] [CrossRef]
  49. Thiebaut, F.; de Urquiaga, M.C.O.; Rosman, A.C.; da Silva, M.L.; Hemerly, A.S. The Impact of Non-Nodulating Diazotrophic Bacteria in Agriculture: Understanding the Molecular Mechanisms That Benefit Crops. Int. J. Mol. Sci. 2022, 23, 11301. [Google Scholar] [CrossRef] [PubMed]
  50. Khan, A.A.; Jilani, G.; Akhtar, M.S.; Naqvi, S.M.S.; Rasheed, M. Phosphorus solubilizing bacteria: Occurrence, mechanisms and their role in crop production. J. Agric. Biol. Sci. 2009, 1, 48–58. [Google Scholar]
  51. Abd-Elsalam, K.A.; Mohamed, H.I. Synthesis and application of bacterial secondary metabolites in agroecosystems: A note from the editors. In Bacterial Secondary Metabolites: Synthesis and Applications in Agroecosystem; Elsevier: Amsterdam, The Netherlands, 2024; pp. 1–14. [Google Scholar]
  52. Pan, L.; Cai, B. Phosphate-Solubilizing Bacteria: Advances in Their Physiology, Molecular Mechanisms and Microbial Community Effects. Microorganisms 2023, 11, 2904. [Google Scholar] [CrossRef]
  53. Pantigoso, H.A.; Manter, D.K.; Fonte, S.J.; Vivanco, J.M. Root exudate-derived compounds stimulate the phosphorus solubilizing ability of bacteria. Sci. Rep. 2023, 13, 4050. [Google Scholar] [CrossRef]
  54. Habibi, S.; Djedidi, S.; Ohkama-Ohtsu, N.; Sarhadi, W.A.; Kojima, K.; Rallos, R.V.; Ramirez, M.D.A.; Yamaya, H.; Sekimoto, H.; Yokoyama, T. Isolation and Screening of Indigenous Plant Growth-promoting Rhizobacteria from Different Rice Cultivars in Afghanistan Soils. Microbes Environ. 2019, 34, 347. [Google Scholar] [CrossRef] [PubMed]
  55. Udaondo, Z.; Duque, E.; Daddaoua, A.; Caselles, C.; Roca, A.; Pizarro-Tobias, P.; Ramos, J.L. Developing robust protein analysis profiles to identify bacterial acid phosphatases in genomes and metagenomic libraries. Environ. Microbiol. 2020, 22, 3561–3571. [Google Scholar] [CrossRef] [PubMed]
  56. Wang, M.; Zheng, Q.; Shen, Q.; Guo, S. The Critical Role of Potassium in Plant Stress Response. Int. J. Mol. Sci. 2013, 14, 7370. [Google Scholar] [CrossRef]
  57. Meena, V.S.; Maurya, B.R.; Verma, J.P. Does a rhizospheric microorganism enhance K+ availability in agricultural soils? Microbiol. Res. 2014, 169, 337–347. [Google Scholar] [CrossRef]
  58. Etesami, H.; Emami, S.; Alikhani, H.A. Potassium solubilizing bacteria (KSB): Mechanisms, promotion of plant growth, and future prospects—A review. J. Soil Sci. Plant Nutr. 2017, 17, 897–911. [Google Scholar] [CrossRef]
  59. Lace, B.; Ott, T. Commonalities and Differences in Controlling Multipartite Intracellular Infections of Legume Roots by Symbiotic Microbes. Plant Cell Physiol. 2018, 59, 661–672. [Google Scholar] [CrossRef] [PubMed]
  60. Denison, R.F.; Bledsoe, C.; Kahn, M.; O’gara, F.; Simms, E.L.; Thomashow, L.S. Cooperation in the rhizosphere and the “free rider” problem. Ecology 2003, 84, 838–845. [Google Scholar] [CrossRef]
  61. Ujvári, G.; Turrini, A.; Avio, L.; Agnolucci, M. Possible role of arbuscular mycorrhizal fungi and associated bacteria in the recruitment of endophytic bacterial communities by plant roots. Mycorrhiza 2021, 31, 527–544. [Google Scholar] [CrossRef] [PubMed]
  62. Bhunjun, C.S.; Phukhamsakda, C.; Hyde, K.D.; McKenzie, E.H.C.; Saxena, R.K.; Li, Q. Do all fungi have ancestors with endophytic lifestyles? Fungal Divers. 2023, 125, 73–98. [Google Scholar] [CrossRef]
  63. Frank, B. Über die auf Wurzelsymbiose Beruhende Ernährung Gewisser Bäume durch Unterirdische Pilze. Berichte Der Dtsch. Bot. Ges. 1885, 3, 128–145. [Google Scholar]
  64. Zuccaro, A.; Lahrmann, U.; Langen, G. Broad compatibility in fungal root symbioses. Curr. Opin. Plant Biol. 2014, 20, 135–145. [Google Scholar] [CrossRef]
  65. Parniske, M. Arbuscular mycorrhiza: The mother of plant root endosymbioses. Nat. Rev. Microbiol. 2008, 6, 763–775. [Google Scholar] [CrossRef]
  66. Jeffries, P.; Gianinazzi, S.; Perotto, S.; Turnau, K.; Barea, J.M. The contribution of arbuscular mycorrhizal fungi in sustainable maintenance of plant health and soil fertility. Biol. Fertil. Soils 2003, 37, 1–16. [Google Scholar] [CrossRef]
  67. López-Ráez, J.A.; Pozo, M.J.; García-Garrido, J.M. Strigolactones: A cry for help in the rhizosphere. Botany 2011, 89, 513–522. [Google Scholar] [CrossRef]
  68. Poveda, J.; Díaz-González, S.; Díaz-Urbano, M.; Velasco, P.; Sacristán, S. Fungal endophytes of Brassicaceae: Molecular interactions and crop benefits. Front. Plant Sci. 2022, 13, 932288. [Google Scholar] [CrossRef]
  69. Miransari, M. Contribution of arbuscular mycorrhizal symbiosis to plant growth under different types of soil stress. Plant Biol. 2010, 12, 563–569. [Google Scholar] [CrossRef]
  70. Antoine, S.; Hériché, M.; Boussageon, R.; Noceto, P.A.; van Tuinen, D.; Wipf, D.; Courty, P.E. A historical perspective on mycorrhizal mutualism emphasizing arbuscular mycorrhizas and their emerging challenges. Mycorrhiza 2021, 31, 637–653. [Google Scholar]
  71. Paszkowski, U. Mutualism and parasitism: The yin and yang of plant symbioses. Curr. Opin. Plant Biol. 2006, 9, 364–370. [Google Scholar] [CrossRef]
  72. Chang, X.; Young, B.; Vaccaro, N.; Strickland, R.; Goldstein, W.; Struwe, L.; White, J.F.; Chang, X.; Young, B.; Vaccaro, N.; et al. Endophyte symbiosis evolutionary development, and impacts of plant agriculture. Grass Res. 2023, 3, 18. [Google Scholar] [CrossRef]
  73. Chen, M.; Arato, M.; Borghi, L.; Nouri, E.; Reinhardt, D. Beneficial services of arbuscular mycorrhizal fungi—From ecology to application. Front. Plant Sci. 2018, 9, 408113. [Google Scholar] [CrossRef]
  74. Salvioli, A.; Bonfante, P. Systems biology and “omics” tools: A cooperation for next-generation mycorrhizal studies. Plant Sci. 2013, 203–204, 107–114. [Google Scholar] [CrossRef]
  75. Singh, A.K.; Zhu, X.; Chen, C.; Wu, J.; Yang, B.; Zakari, S.; Jiang, X.J.; Singh, N.; Liu, W. The role of glomalin in mitigation of multiple soil degradation problems. Crit. Rev. Environ. Sci. Technol. 2022, 52, 1604–1638. [Google Scholar] [CrossRef]
  76. Porcel, R.; Aroca, R.; Ruiz-Lozano, J.M. Salinity stress alleviation using arbuscular mycorrhizal fungi. A review. Agron. Sustain. Dev. 2012, 32, 181–200. [Google Scholar] [CrossRef]
  77. Grassi, A.; Pagliarani, I.; Avio, L.; Cristani, C.; Rossi, F.; Turrini, A.; Giovannetti, M.; Agnolucci, M. Bioprospecting for plant resilience to climate change: Mycorrhizal symbionts of European and American beachgrass (Ammophila arenaria and Ammophila breviligulata) from maritime sand dunes. Mycorrhiza 2024, 1–13. [Google Scholar] [CrossRef]
  78. Boomsma, C.R.; Vyn, T.J. Maize drought tolerance: Potential improvements through arbuscular mycorrhizal symbiosis? Field Crops Res. 2008, 108, 14–31. [Google Scholar] [CrossRef]
  79. Wu, Q.S.; Srivastava, A.K.; Zou, Y.N. AMF-induced tolerance to drought stress in citrus: A review. Sci. Hortic. 2013, 164, 77–87. [Google Scholar] [CrossRef]
  80. Evelin, H.; Kapoor, R.; Giri, B. Arbuscular mycorrhizal fungi in alleviation of salt stress: A review. Ann. Bot. 2009, 104, 1263–1280. [Google Scholar] [CrossRef] [PubMed]
  81. Saboor, A.; Ali, M.A.; Hussain, S.; El Enshasy, H.A.; Hussain, S.; Ahmed, N.; Gafur, A.; Sayyed, R.Z.; Fahad, S.; Danish, S.; et al. Zinc nutrition and arbuscular mycorrhizal symbiosis effects on maize (Zea mays L.) growth and productivity. Saudi J. Biol. Sci. 2021, 28, 6339–6351. [Google Scholar] [CrossRef] [PubMed]
  82. Albuquerque da Silva Campos, M. Bioprotection by arbuscular mycorrhizal fungi in plants infected with Meloidogyne nematodes: A sustainable alternative. Crop Prot. 2020, 135, 105203. [Google Scholar] [CrossRef]
  83. Guerrero-Galán, C.; Calvo-Polanco, M.; Zimmermann, S.D. Ectomycorrhizal symbiosis helps plants to challenge salt stress conditions. Mycorrhiza 2019, 29, 291–301. [Google Scholar] [CrossRef] [PubMed]
  84. Thoms, D.; Liang, Y.; Haney, C.H. Maintaining symbiotic homeostasis: How do plants engage with beneficial microorganisms while at the same time restricting pathogens? Mol. Plant-Microbe Interact. 2021, 34, 462–469. [Google Scholar] [CrossRef] [PubMed]
  85. Ghahremani, M.; MacLean, A.M. Home sweet home: How mutualistic microbes modify root development to promote symbiosis. J. Exp. Bot. 2021, 72, 2275–2287. [Google Scholar] [CrossRef]
  86. Choi, J.; Summers, W.; Paszkowski, U. Mechanisms Underlying Establishment of Arbuscular Mycorrhizal Symbioses. Annu. Rev. Phytopathol. 2018, 56, 135–160. [Google Scholar] [CrossRef]
  87. Duc, N.H.; Vo, H.T.N.; van Doan, C.; Hamow, K.Á.; Le, K.H.; Posta, K. Volatile organic compounds shape belowground plant–fungi interactions. Front. Plant Sci. 2022, 13, 1046685. [Google Scholar] [CrossRef]
  88. Bouwmeester, H.J.; Roux, C.; Lopez-Raez, J.A.; Bécard, G. Rhizosphere communication of plants, parasitic plants and AM fungi. Trends Plant Sci. 2007, 12, 224–230. [Google Scholar] [CrossRef]
  89. Rengel, Z.; Marschner, P. Nutrient availability and management in the rhizosphere: Exploiting genotypic differences. New Phytol. 2005, 168, 305–312. [Google Scholar] [CrossRef] [PubMed]
  90. Llorens, E.; López-Moral, A.; Agustí-Brisach, C. Root Exudates Metabolic Profiling Confirms Distinct Defense Mechanisms Between Cultivars and Treatments with Beneficial Microorganisms and Phosphonate Salts Against Verticillium Wilt in Olive Trees. Phytopathology 2024. [Google Scholar] [CrossRef] [PubMed]
  91. Liu-Xu, L.; Vicedo, B.; García-Agustín, P.; Llorens, E. Advances in endophytic fungi research: A data analysis of 25 years of achievements and challenges. J. Plant Interact. 2022, 17, 244–266. [Google Scholar] [CrossRef]
  92. Mandyam, K.G.; Jumpponen, A. Mutualism-parasitism paradigm synthesized from results of root-endophyte models. Front. Microbiol. 2014, 5, 123328. [Google Scholar]
  93. Hyde, K.; Soytong, K. the fungal endophyte dilemma. Fungal Divers. 2008, 33, 163–173. [Google Scholar]
  94. Aly, A.H.; Debbab, A.; Proksch, P. Fungal endophytes: Unique plant inhabitants with great promises. Appl. Microbiol. Biotechnol. 2011, 90, 1829–1845. [Google Scholar] [CrossRef] [PubMed]
  95. Hardoim, P.R.; van Overbeek, L.S.; Berg, G.; Pirttilä, A.M.; Compant, S.; Campisano, A.; Döring, M.; Sessitsch, A. The Hidden World within plants: Ecological and evolutionary considerations for defining functioning of microbial endophytes. Microbiol. Mol. Biol. Rev. 2015, 79, 293–320. [Google Scholar] [CrossRef] [PubMed]
  96. Schardl, C.L.; Leuchtmann, A.; Spiering, M.J. Symbioses of grasses with seedborne fungal endophytes. Annu. Rev. Plant Biol. 2004, 55, 315–340. [Google Scholar] [CrossRef] [PubMed]
  97. Arnold, A.E. Understanding the diversity of foliar endophytic fungi: Progress, challenges, and frontiers. Fungal Biol. Rev. 2007, 21, 51–66. [Google Scholar] [CrossRef]
  98. Guerre, P. Ergot alkaloids produced by endophytic fungi of the genus Epichloë. Toxins 2015, 7, 773–790. [Google Scholar] [CrossRef]
  99. Belesky, D.P.; Bacon, C.W. Tall fescue and associated mutualistic toxic fungal endophytes in agroecosystems. Toxin Rev. 2009, 28, 102–117. [Google Scholar] [CrossRef]
  100. Ting, A.S.Y.; Meon, S.; Kadir, J.; Radu, S.; Singh, G. Endophytic microorganisms as potential growth promoters of banana. BioControl 2008, 53, 541–553. [Google Scholar] [CrossRef]
  101. Hamayun, M.; Afzal Khan, S.; Ahmad, N.; Tang, D.S.; Kang, S.M.; Na, C.I.; Sohn, E.Y.; Hwang, Y.H.; Shin, D.H.; Lee, B.H.; et al. Cladosporium sphaerospermum as a new plant growth-promoting endophyte from the roots of Glycine max (L.) Merr. World J. Microbiol. Biotechnol. 2009, 25, 627–632. [Google Scholar] [CrossRef]
  102. Khan, A.L.; Lee, I.J. Endophytic Penicillium funiculosum LHL06 secretes gibberellin that reprograms Glycine max L. growth during copper stress. BMC Plant Biol. 2013, 13, 86. [Google Scholar] [CrossRef] [PubMed]
  103. Woo, S.L.; Hermosa, R.; Lorito, M.; Monte, E. Trichoderma: A multipurpose, plant-beneficial microorganism for eco-sustainable agriculture. Nat. Rev. Microbiol. 2022, 21, 312–326. [Google Scholar] [CrossRef]
  104. Hagag, A.; Abdelwahab, M.F.; Abd El-kader, A.M.; Fouad, M.A. The endophytic Aspergillus strains: A bountiful source of natural products. J. Appl. Microbiol. 2022, 132, 4150–4169. [Google Scholar] [CrossRef]
  105. Weiß, M.; Waller, F.; Zuccaro, A.; Selosse, M.A. Sebacinales—One thousand and one interactions with land plants. New Phytol. 2016, 211, 20–40. [Google Scholar] [CrossRef]
  106. Vurukonda, S.S.K.P.; Giovanardi, D.; Stefani, E. Plant Growth Promoting and Biocontrol Activity of Streptomyces spp. as Endophytes. Int. J. Mol. Sci. 2018, 19, 952. [Google Scholar] [CrossRef] [PubMed]
  107. Hossain, M.M.; Sultana, F.; Islam, S. Plant Growth-Promoting Fungi (PGPF): Phytostimulation and Induced Systemic Resistance. Plant-Microbe Interact. Agro-Ecol. Perspect. 2017, 2, 135–191. [Google Scholar]
  108. Attia, M.S.; Abdelaziz, A.M.; Al-Askar, A.A.; Arishi, A.A.; Abdelhakim, A.M.; Hashem, A.H. Plant Growth-Promoting Fungi as Biocontrol Tool against Fusarium Wilt Disease of Tomato Plant. J. Fungi 2022, 8, 775. [Google Scholar] [CrossRef]
  109. Rodriguez, R.J.; White, J.F.; Arnold, A.E.; Redman, R.S. Fungal endophytes: Diversity and functional roles: Tansley review. New Phytol. 2009, 182, 314–330. [Google Scholar] [CrossRef] [PubMed]
  110. Yan, L.; Zhu, J.; Zhao, X.; Shi, J.; Jiang, C.; Shao, D. Beneficial effects of endophytic fungi colonization on plants. Appl. Microbiol. Biotechnol. 2019, 103, 3327–3340. [Google Scholar] [CrossRef] [PubMed]
  111. Thilini Chethana, K.W.; Jayawardena, R.S.; Chen, Y.J.; Konta, S.; Tibpromma, S.; Abeywickrama, P.D.; Gomdola, D.; Balasuriya, A.; Xu, J.; Lumyong, S.; et al. Diversity and Function of Appressoria. Pathogens 2021, 10, 746. [Google Scholar] [CrossRef] [PubMed]
  112. Redkar, A.; Sabale, M.; Zuccaro, A.; Di Pietro, A. Determinants of endophytic and pathogenic lifestyle in root colonizing fungi. Curr. Opin. Plant Biol. 2022, 67, 102226. [Google Scholar] [CrossRef] [PubMed]
  113. Galindo-Solís, J.M.; Fernández, F.J. Endophytic Fungal Terpenoids: Natural Role and Bioactivities. Microorganisms 2022, 10, 339. [Google Scholar] [CrossRef] [PubMed]
  114. Hamilton, C.E.; Gundel, P.E.; Helander, M.; Saikkonen, K. Endophytic mediation of reactive oxygen species and antioxidant activity in plants: A review. Fungal Divers. 2012, 54, 1–10. [Google Scholar] [CrossRef]
  115. Varma, A.; Bakshi, M.; Lou, B.; Hartmann, A.; Oelmueller, R. Piriformospora indica: A Novel Plant Growth-Promoting Mycorrhizal Fungus. Agric. Res. 2012, 1, 117–131. [Google Scholar] [CrossRef]
  116. Lahrmann, U.; Zuccaro, A. Opprimo ergo sum--evasion and suppression in the root endophytic fungus Piriformospora indica. Mol. Plant. Microbe. Interact. 2012, 25, 727–737. [Google Scholar] [CrossRef] [PubMed]
  117. Behie, S.W.; Bidochka, M.J. Nutrient transfer in plant-fungal symbioses. Trends Plant Sci. 2014, 19, 734–740. [Google Scholar] [CrossRef]
  118. Vega, F.E. The use of fungal entomopathogens as endophytes in biological control: A review. Mycologia 2018, 110, 4–30. [Google Scholar] [CrossRef]
  119. Rai, N.; Gupta, P.; Keshri, P.K.; Verma, A.; Mishra, P.; Kumar, D.; Kumar, A.; Santosh, S.K.; Singh, K.; Gautam, V.; et al. Fungal Endophytes: An Accessible Source of Bioactive Compounds with Potential Anticancer Activity. Appl. Biochem. Biotechnol. 2010, 194, 3296–3319. [Google Scholar] [CrossRef] [PubMed]
  120. Nisa, H.; Kamili, A.N.; Nawchoo, I.A.; Shafi, S.; Shameem, N.; Bandh, S.A. Fungal endophytes as prolific source of phytochemicals and other bioactive natural products: A review. Microb. Pathog. 2015, 82, 50–59. [Google Scholar] [CrossRef] [PubMed]
  121. Elvira Sánchez-Fernández, R.; Lorena Sánchez-Ortiz, B.; Monserrat Sandoval-Espinosa, Y.K.; Ulloa-Benítez, Á.; Armendáriz-Guillén, B.; Claudia García-Méndez, M.; Lydia Macías-Rubalcava, M. Hongos endófitos: Fuente potencial de metabolitos secundarios bioactivos con utilidad en agricultura y medicina. Tip 2013, 16, 132–146. [Google Scholar] [CrossRef]
  122. Saikkonen, K.; Young, C.A.; Helander, M.; Schardl, C.L. Endophytic Epichloë species and their grass hosts: From evolution to applications. Plant Mol. Biol. 2016, 90, 665–675. [Google Scholar] [CrossRef]
  123. Van Deynze, A.; Zamora, P.; Delaux, P.M.; Heitmann, C.; Jayaraman, D.; Rajasekar, S.; Graham, D.; Maeda, J.; Gibson, D.; Schwartz, K.D.; et al. Nitrogen fixation in a landrace of maize is supported by a mucilage-associated diazotrophic microbiota. PLoS Biol. 2018, 16, e2006352. [Google Scholar] [CrossRef] [PubMed]
  124. Serna-Cock, L.; Arias-García, C.; Valencia Hernandez, L.J. Effect of Biofertilization on the growth of potted sugarcane plants (Saccharum officinarum). Biotecnol. Sect. Agropecu. Agroindustrial 2011, 9, 85–95. [Google Scholar]
  125. Antunes, J.E.L.; De Freitas, A.D.S.; Oliveira, L.M.S.; De Lyra, M.D.C.C.P.; Fonseca, M.A.C.; Santos, C.E.R.S.; De Oliveira, J.P.; De Araújo, A.S.F.; Figueiredo, M.V.B. Sugarcane inoculated with endophytic diazotrophic bacteria: Effects on yield, biological nitrogen fixation and industrial characteristics. An. Acad. Bras. Cienc. 2019, 91, e20180990. [Google Scholar] [CrossRef] [PubMed]
  126. Babalola, O.O. Beneficial bacteria of agricultural importance. Biotechnol. Lett. 2010, 32, 1559–1570. [Google Scholar] [CrossRef] [PubMed]
  127. Glick, B.R. Plant Growth-Promoting Bacteria: Mechanisms and Applications. Scientifica 2012, 2012, 963401. [Google Scholar] [CrossRef]
  128. Hayat, Q.; Hayat, S.; Irfan, M.; Ahmad, A. Effect of exogenous salicylic acid under changing environment: A review. Environ. Exp. Bot. 2010, 68, 14–25. [Google Scholar] [CrossRef]
  129. Bargaz, A.; Lyamlouli, K.; Chtouki, M.; Zeroual, Y.; Dhiba, D. Soil Microbial Resources for Improving Fertilizers Efficiency in an Integrated Plant Nutrient Management System. Front. Microbiol. 2018, 9, 1606. [Google Scholar] [CrossRef] [PubMed]
  130. Gupta, M.; Bisht, S.; Singh, B.; Gulati, A.; Tewari, R. Enhanced biomass and steviol glycosides in Stevia rebaudiana treated with phosphate-solubilizing bacteria and rock phosphate. Plant Growth Regul. 2011, 65, 449–457. [Google Scholar] [CrossRef]
  131. Javeed, H.M.R.; Qamar, R.; Ur Rehman, A.; Ali, M.; Rehman, A.; Farooq, M.; Zamir, S.I.; Nadeem, M.; Cheema, M.A.; Shehzad, M.; et al. Improvement in soil characteristics of sandy loam soil and grain quality of spring maize by using phosphorus solublizing bacteria. Sustainability 2019, 11, 7049. [Google Scholar] [CrossRef]
  132. Suleman, M.; Yasmin, S.; Rasul, M.; Yahya, M.; Atta, B.M.; Mirza, M.S. Phosphate solubilizing bacteria with glucose dehydrogenase gene for phosphorus uptake and beneficial effects on wheat. PLoS ONE 2018, 13, e0204408. [Google Scholar] [CrossRef] [PubMed]
  133. Shirmohammadi, E.; Alikhani, H.A.; Pourbabaei, A.A.; Etesami, H. Improved Phosphorus (P) Uptake and Yield of Rainfed Wheat Fed with P Fertilizer by Drought-Tolerant Phosphate-Solubilizing Fluorescent Pseudomonads Strains: A Field Study in Drylands. J. Soil Sci. Plant Nutr. 2020, 20, 2195–2211. [Google Scholar] [CrossRef]
  134. Duarah, I.; Deka, M.; Saikia, N.; Deka Boruah, H.P. Phosphate solubilizers enhance NPK fertilizer use efficiency in rice and legume cultivation. 3 Biotech 2011, 1, 227–238. [Google Scholar] [CrossRef] [PubMed]
  135. Barra, P.J.; Pontigo, S.; Delgado, M.; Parra–Almuna, L.; Duran, P.; Valentine, A.J.; Jorquera, M.A.; de la Mora, M.L. Phosphobacteria inoculation enhances the benefit of P–fertilization on Lolium perenne in soils contrasting in P–availability. Soil Biol. Biochem. 2019, 136, 107516. [Google Scholar] [CrossRef]
  136. Verma, V.; Ravindran, P.; Kumar, P.P. Plant hormone-mediated regulation of stress responses. BMC Plant Biol. 2016, 16, 86. [Google Scholar] [CrossRef] [PubMed]
  137. Hou, L.; Zhang, X.; Feng, G.; Li, Z.; Zhang, Y.; Cao, N. Arbuscular mycorrhizal enhancement of phosphorus uptake and yields of maize under high planting density in the black soil region of China. Sci. Rep. 2021, 11, 1100. [Google Scholar] [CrossRef]
  138. Zhang, L.; Shi, N.; Fan, J.; Wang, F.; George, T.S.; Feng, G. Arbuscular mycorrhizal fungi stimulate organic phosphate mobilization associated with changing bacterial community structure under field conditions. Environ. Microbiol. 2018, 20, 2639–2651. [Google Scholar] [CrossRef]
  139. Bouhraoua, D.; Aarab, S.; Laglaoui, A.; Bakkali, M.; Arakrak, A. Phosphate Solubilizing Bacteria Efficiency on Mycorrhization and Growth of Peanut in the Northwest of Morocco. Am. J. Microbiol. Res. 2015, 3, 176–180. [Google Scholar]
  140. de Novais, C.B.; Sbrana, C.; da Conceição Jesus, E.; Rouws, L.F.M.; Giovannetti, M.; Avio, L.; Siqueira, J.O.; Saggin Júnior, O.J.; da Silva, E.M.R.; de Faria, S.M. Mycorrhizal networks facilitate the colonization of legume roots by a symbiotic nitrogen-fixing bacterium. Mycorrhiza 2020, 30, 389–396. [Google Scholar] [CrossRef] [PubMed]
  141. Chen, H.; Renault, S.; Markham, J. The effect of Frankia and multiple ectomycorrhizal fungil species on Alnus growing in low fertility soil. Symbiosis 2020, 80, 207–215. [Google Scholar] [CrossRef]
  142. Ayaz, M.; Li, C.-H.; Ali, Q.; Zhao, W.; Chi, Y.-K.; Shafiq, M.; Ali, F.; Yu, X.-Y.; Yu, Q.; Zhao, J.-T.; et al. Bacterial and Fungal Biocontrol Agents for Plant Disease Protection: Journey from Lab to Field, Current Status, Challenges, and Global Perspectives. Molecules 2023, 28, 6735. [Google Scholar] [CrossRef] [PubMed]
  143. Shternshis, M.V.; Belyaev, A.A.; Shpatova, T.V.; Lelyak, A.A. Influence of Bacillus spp. on strawberry gray-mold causing agent and host plant resistance to disease. Contemp. Probl. Ecol. 2015, 8, 390–396. [Google Scholar] [CrossRef]
  144. Jayakumar, A.; Krishna, A.; Mohan, M.; Nair, I.C.; Radhakrishnan, E.K. Plant Growth Enhancement, Disease Resistance, and Elemental Modulatory Effects of Plant Probiotic Endophytic Bacillus sp. Fcl1. Probiotics Antimicrob. Proteins 2019, 11, 526–534. [Google Scholar] [CrossRef] [PubMed]
  145. Figueredo, M.S.; Tonelli, M.L.; Ibáñez, F.; Morla, F.; Cerioni, G.; del Carmen Tordable, M.; Fabra, A. Induced systemic resistance and symbiotic performance of peanut plants challenged with fungal pathogens and co-inoculated with the biocontrol agent Bacillus sp. CHEP5 and Bradyrhizobium sp. SEMIA6144. Microbiol. Res. 2017, 197, 65–73. [Google Scholar] [CrossRef] [PubMed]
  146. Anderson, A.J.; Kim, Y.C. Biopesticides produced by plant-probiotic Pseudomonas chlororaphis isolates. Crop Prot. 2018, 105, 62–69. [Google Scholar] [CrossRef]
  147. Kang, B.R.; Anderson, A.J.; Kim, Y.C. The Plant Pathology Journal Hydrogen Cyanide Produced by Pseudomonas chlororaphis O6 Exhibits Nematicidal Activity against Meloidogyne hapla. Plant Pathol. J. 2018, 34, 35–43. [Google Scholar] [CrossRef]
  148. Chutulo, E.C.; Chalannavar, R.K. Endophytic Mycoflora and Their Bioactive Compounds from Azadirachta Indica: A Comprehensive Review. J. Fungi 2018, 4, 42. [Google Scholar] [CrossRef]
  149. Mousa, W.K.; Raizada, M.N. The diversity of anti-microbial secondary metabolites produced by fungal endophytes: An interdisciplinary perspective. Front. Microbiol. 2013, 4, 44840. [Google Scholar] [CrossRef] [PubMed]
  150. War, A.R.; Paulraj, M.G.; Ahmad, T.; Buhroo, A.A.; Hussain, B.; Ignacimuthu, S.; Sharma, H.C. Mechanisms of plant defense against insect herbivores. Plant Signal. Behav. 2012, 7, 1306–1320. [Google Scholar] [CrossRef] [PubMed]
  151. Abdulhadi, S.Y.; Hasan, G.Q.; Gergees, R.N. Molecular detection and antimicrobial activity of Endophytic fungi isolated from a medical plant Rosmarinus officinalis. Ann. Trop. Med. Public Health 2020, 23. [Google Scholar] [CrossRef]
  152. Rashid, T.S. Bioactive metabolites from tomato endophytic fungi with antibacterial activity against tomato bacterial spot disease. Rhizosphere 2021, 17, 100292. [Google Scholar] [CrossRef]
  153. Qin, X.; Zhao, X.; Huang, S.; Deng, J.; Li, X.; Luo, Z.; Zhang, Y. Pest management via endophytic colonization of tobacco seedlings by the insect fungal pathogen Beauveria bassiana. Pest Manag. Sci. 2021, 77, 2007–2018. [Google Scholar] [CrossRef] [PubMed]
  154. Hartley, S.E.; Eschen, R.; Horwood, J.M.; Gange, A.C.; Hill, E.M. Infection by a foliar endophyte elicits novel arabidopside-based plant defence reactions in its host, C irsium arvense. New Phytol. 2015, 205, 816–827. [Google Scholar] [CrossRef] [PubMed]
  155. Tian, B.; Xie, J.; Fu, Y.; Cheng, J.; Li, B.O.; Chen, T.; Zhao, Y.; Gao, Z.; Yang, P.; Barbetti, M.J. A cosmopolitan fungal pathogen of dicots adopts an endophytic lifestyle on cereal crops and protects them from major fungal diseases. ISME J. 2020, 14, 3120–3135. [Google Scholar] [CrossRef] [PubMed]
  156. Martínez-Arias, C.; Sobrino-Plata, J.; Gil, L.; Rodríguez-Calcerrada, J.; Martín, J.A. Priming of plant defenses against Ophiostoma novo-ulmi by Elm (Ulmus minor Mill.) fungal endophytes. J. Fungi 2021, 7, 687. [Google Scholar] [CrossRef] [PubMed]
  157. Zhang, F.-M.; He, W.; Wu, C.-Y.; Sun, K.; Zhang, W.; Dai, C.-C. Phomopsis liquidambaris inoculation induces resistance in peanut to leaf spot and root rot. BioControl 2020, 65, 475–488. [Google Scholar] [CrossRef]
  158. Llorens, E.; Scalschi, L.; Sharon, O.; Vicedo, B.; Sharon, A.; García-Agustín, P. Jasmonic acid pathway is required in the resistance induced by Acremonium sclerotigenum in tomato against Pseudomonas syringae. Plant Sci. 2022, 318, 111210. [Google Scholar] [CrossRef]
  159. Sallam, N.; Ali, E.F.; Seleim, M.A.A.; Khalil Bagy, H.M.M. Endophytic fungi associated with soybean plants and their antagonistic activity against Rhizoctonia solani. Egypt. J. Biol. Pest Control 2021, 31, 54. [Google Scholar] [CrossRef]
  160. Abdou, R.; Alqahtani, A.M.; Attia, G.H. Bioactive Metabolites of Aspergillus neoniger, an Endophyte of the Medicinal Plant Ficus carica. Indian J. Pharm. Sci. 2021, 83, 101–109. [Google Scholar] [CrossRef]
  161. Soumare, A.; Diedhiou, A.G.; Thuita, M.; Hafidi, M.; Ouhdouch, Y.; Gopalakrishnan, S.; Kouisni, L. Exploiting Biological Nitrogen Fixation: A Route Towards a Sustainable Agriculture. Plants 2020, 9, 1011. [Google Scholar] [CrossRef] [PubMed]
  162. White, C.; Sayer, J.A.; Gadd, G.M. Microbial solubilization and immobilization of toxic metals: Key biogeochemical processes for treatment of contamination. FEMS Microbiol. Rev. 1997, 20, 503–516. [Google Scholar] [CrossRef] [PubMed]
  163. Kang, J.; Amoozegar, A.; Hesterberg, D.; Osmond, D.L. Phosphorus leaching in a sandy soil as affected by organic and inorganic fertilizer sources. Geoderma 2011, 161, 194–201. [Google Scholar] [CrossRef]
  164. Rosado, A.S.; De Azevedo, F.S.; Da Cruz, D.W.; Van Elsas, J.D.; Seldin, L. Phenotypic and genetic diversity of Paenibacillus azotofixans strains isolated from the rhizoplane or rhizosphere soil of different grasses. J. Appl. Microbiol. 1998, 84, 216–226. [Google Scholar] [CrossRef]
  165. Johri, J.K.; Surange, S.; Nautiyal, C.S. Occurrence of salt, pH, and temperature-tolerant, phosphate-solubilizing bacteria in alkaline soils. Curr. Microbiol. 1999, 39, 89–93. [Google Scholar] [CrossRef] [PubMed]
  166. Timofeeva, A.; Galyamova, M.; Sedykh, S. Prospects for Using Phosphate-Solubilizing Microorganisms as Natural Fertilizers in Agriculture. Plants 2022, 11, 2119. [Google Scholar] [CrossRef]
  167. Arora, R.; Gupta, M.; Singh, H. Fortification of Fertilizers with Organic Amendments and Bio-Inoculants for Augmenting Yield and Quality in Papaya. Agric. Res. J. 2021, 58, 1065–1070. [Google Scholar] [CrossRef]
Figure 1. Location and role of the microbiome in different parts of the plant. Figure adapted from Johnston-Monje et al. [6] and Dhaliwal et al. [7].
Figure 1. Location and role of the microbiome in different parts of the plant. Figure adapted from Johnston-Monje et al. [6] and Dhaliwal et al. [7].
Horticulturae 10 00621 g001
Figure 2. Potential benefits of interaction between symbiotic microorganisms and plants.
Figure 2. Potential benefits of interaction between symbiotic microorganisms and plants.
Horticulturae 10 00621 g002
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

Liu-Xu, L.; González-Hernández, A.I.; Camañes, G.; Vicedo, B.; Scalschi, L.; Llorens, E. Harnessing Green Helpers: Nitrogen-Fixing Bacteria and Other Beneficial Microorganisms in Plant–Microbe Interactions for Sustainable Agriculture. Horticulturae 2024, 10, 621. https://doi.org/10.3390/horticulturae10060621

AMA Style

Liu-Xu L, González-Hernández AI, Camañes G, Vicedo B, Scalschi L, Llorens E. Harnessing Green Helpers: Nitrogen-Fixing Bacteria and Other Beneficial Microorganisms in Plant–Microbe Interactions for Sustainable Agriculture. Horticulturae. 2024; 10(6):621. https://doi.org/10.3390/horticulturae10060621

Chicago/Turabian Style

Liu-Xu, Luisa, Ana Isabel González-Hernández, Gemma Camañes, Begonya Vicedo, Loredana Scalschi, and Eugenio Llorens. 2024. "Harnessing Green Helpers: Nitrogen-Fixing Bacteria and Other Beneficial Microorganisms in Plant–Microbe Interactions for Sustainable Agriculture" Horticulturae 10, no. 6: 621. https://doi.org/10.3390/horticulturae10060621

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