Next Article in Journal
Socialization of Providencia stuartii Enables Resistance to Environmental Insults
Previous Article in Journal
Molecular Epidemiology and Evolutionary Dynamics of Human Influenza Type-A Viruses in Africa: A Systematic Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Rhizosphere Signaling: Insights into Plant–Rhizomicrobiome Interactions for Sustainable Agronomy

1
Institute of Industrial Biotechnology, Government College University, Lahore 54000, Pakistan
2
CHEMBIOPRO Chimie et Biotechnologie des Produits Naturels, Faculté des Sciences et Technologies, Université de la Réunion, F-97490 Sainte-Clotilde, Ile de La Réunion, France
3
CHEMBIOPRO Chimie et Biotechnologie des Produits Naturels, ESIROI Département Agroalimentaire, Université de la Réunion, F-97490 Sainte-Clotilde, Ile de La Réunion, France
*
Authors to whom correspondence should be addressed.
Microorganisms 2022, 10(5), 899; https://doi.org/10.3390/microorganisms10050899
Submission received: 21 March 2022 / Revised: 19 April 2022 / Accepted: 21 April 2022 / Published: 25 April 2022
(This article belongs to the Section Environmental Microbiology)

Abstract

:
Rhizospheric plant–microbe interactions have dynamic importance in sustainable agriculture systems that have a reduced reliance on agrochemicals. Rhizosphere signaling focuses on the interactions between plants and the surrounding symbiotic microorganisms that facilitate the development of rhizobiome diversity, which is beneficial for plant productivity. Plant–microbe communication comprises intricate systems that modulate local and systemic defense mechanisms to mitigate environmental stresses. This review deciphers insights into how the exudation of plant secondary metabolites can shape the functions and diversity of the root microbiome. It also elaborates on how rhizosphere interactions influence plant growth, regulate plant immunity against phytopathogens, and prime the plant for protection against biotic and abiotic stresses, along with some recent well-reported examples. A holistic understanding of these interactions can help in the development of tailored microbial inoculants for enhanced plant growth and targeted disease suppression.

1. Introduction

The rhizosphere is a most captivating environment, which harbors a variety of microorganisms that are deeply involved in plant–microbe communication. This high-density niche allows plants to interact with associated microorganisms through chemical signals that are produced in response to specific stimuli, which in turn activate many regulatory mechanisms [1]. Rhizospheric regions possess higher bacterial activity than non-rhizospheric regions. The composition of a microbiome is determined by different biotic and abiotic factors, i.e., the climate, type of soil, and chemical signals that are produced by the plant and its associated microbes [2,3].
Plants and microbes have diverse interactions that involve close interactions, either positive or negative, including mutualism (symbiosis), parasitism, and commensalism [4]. Positive interactions include those with microorganisms, e.g., rhizobia, plant growth-promoting rhizobacteria (PGPR), and mycorrhiza, which result in beneficial outcomes, such as growth promotion, nutrient accessibility, and protection against abiotic and biotic environmental stresses [5,6,7]. On the other hand, plant interactions with microbial pathogens result in negative outcomes, i.e., plant diseases [8,9]. The interaction between both partners depends on specialized signaling molecules or chemical signals that are also significant in cooperative, as well as competitive, microbial behavior [10,11]. The chemical cues or secondary metabolites act as mediators in plant–microbe and microbe–microbe communication and also trigger plant responses [12].
In the past decade, progress has been made in the understanding of the types of chemical signals that are responsible for controlling the activities of plants and associated microbes [13,14,15]. The most studied signaling compounds are N-Acyl homoserine lactones (N-AHLs), which are produced by a variety of bacterial taxa and regulate quorum sensing and pathogenicity within a bacterial population [16,17,18]. Likewise, plant roots secrete a variety of metabolites as exudates, including photosynthetically derived carbon compounds, e.g., organic acids, vitamins, flavonoids, polysaccharides, amino acids, and sugars. These root exudates create an enriched environment for the rhizomicrobiome to interact and increase diversity based on the composition of the exudates [15,19,20]. Moreover, plants release secondary metabolites against pathogens and insects that act as defensive signals [21,22]. Plants use adaptive strategies to enhance the defensive capacity of their innate resistance to biotic and abiotic factors by interacting with beneficial microbes [23,24]. To unravel the process of microbial interaction with plants, an understanding of the types of chemical communication between all members is necessary. Thus, the known microbial community and their interactions could help in the optimal use of beneficial microbes for better plant growth.
Some of the literature on different signaling molecules that participate in the development of interactions within the rhizosphere for enhanced plant growth has been reviewed [15,25,26,27,28]. However, much still needs to be explored in terms of the significance of these interactions as far as microbial or chemical diversity and the understanding of signaling molecules are concerned. This review aims to enlighten our understanding of rhizosphere signaling in plant–microbe communication, both cooperative and competitive, and its significance in plant productivity and the development of sustainable agriculture systems. Nonetheless, the signaling compounds have a prodigious potential to escalate plant functions when they are understood in depth.

2. Rhizosphere: A Pool of Plant–Microbe Signaling

The rhizosphere serves as a hotspot for diverse microbial activity. It is an intricate ecosystem comprising nutrient-rich soil that surrounds the plant roots, which provides a pool for plant–microbe communication. The term “Rhizomicrobiome” is defined as a microbial community that is present in the rhizosphere [29]. A variety of microorganisms reside within the rhizosphere, including bacteria, fungi, nematodes, protists, and invertebrates [11]. Most of the microbiome studies within the context of rhizosphere signaling have been focused on the bacteria and fungi that make up the major portion of the rhizosphere microbiome. However, viruses that infect bacteria, which are known as phages, have an influence on the dynamics of the rhizosphere microbiome. The most abundant known phages in the soil virome include Microviridae, Siphoviridae, and Podoviridae [30]. The diversity and amount of identified viruses in the rhizosphere is much less, i.e., 2700 less than bacterial abundance, which amounts to around 1010 species per g of soil [31,32]. The complete biological entity comprising a plant and its associated microbial community, i.e., the symbionts (facultative and obligate) and microbes that have parasitic relationships with the host plant, is referred to as a “holobiont” [33]. The cascade of intricate chemical signals develops a communication within the rhizosphere that facilitates various mechanisms in a holobiont, i.e., root–root interactions, biofilm formation, nutrient acquisition, microbiota development, and resistance against pathogens [34,35,36]. Different types of interactions between microbes and plants are depicted in Figure 1.
According to the “rhizosphere effect”, this region is associated with a lesser microbial diversity but a higher bacterial abundance compared to bulk soil [37]. The recruitment of microbes within the rhizosphere directly depends on the soil properties, soil type, and plant metabolites [38]. Different plant genotypes and physicochemical soil properties develop specific environments for the selection of a promising microbiome [10,34]. Plant-derived metabolites or root exudates play a significant role in the shaping of the rhizomicrobiome by modifying the soil chemistry around the plant roots and serving as substrates for the growth of specific microbiota [39]. The root exudates differ both in quality and quantity depending on the type, nutritional status, and growth stage of the plant [40]. These root exudates, including both primary and secondary metabolites, contain 10–16% of the total plant nitrogen and 11% of the photosynthetically fixed carbon that actively tune the microbiota in the microbial reservoir that is within the vicinity of the plant roots [41]. Another factor that affects the rhizomicrobiome is the cellular response that is shown either by the microbes or the plants, which results in transformation, catabolism, and resistance to the chemical that is being sensed [11]. The release of root exudates into the rhizosphere by plants selects the desired microbial community by attracting it while deterring harmful communities, which in turn allows plants to be adaptable. This selection process for microbiota is known as “niche colonization” [42,43].
Signaling in the rhizosphere can be categorized into two major types based on the direction of the communication, i.e., the inter- and intraspecies microbial signaling and the interkingdom signaling between microbes and plants.

2.1. Inter- or Intraspecies Signaling among Microorganisms

Microbes in the rhizomicrobiome interact with each other by producing signaling molecules that adjust their gene expression [11]. Inter- or intraspecies communication among microbes occur via the quorum sensing mechanism, which involves cell density-dependent coordination. Quorum sensing (QS) is the cell-to-cell signaling process that takes place through the synthesis, release, and detection of chemical signals [44,45]. These chemical signals are also known as autoinducers (AIs) and they regulate the gene expression of some bacterial functions upon the recognition of the signal by the recipient, i.e., biofilm formation, adhesion, motility [46,47,48], propagation, virulence, metabolism [49], and symbiotic association [1].
Quorum sensing signals in bacteria are categorized into two groups: acyl-homoserine lactone (AHLs or AI-1), which is found in Gram-negative bacteria, and autoinducer peptides (AIPs), which are found in Gram-positive bacteria; and autoinducer type 2 (AI-2), which has the traits of both AHLs and AIPs and is found in both Gram-negative and Gram-positive bacteria [50,51]. N-acyl homoserine lactone (N-AHL) has been majorly reported in various Gram-negative bacteria, including Burkholderia sp., Pseudomonas syringae, Pseudomonas putida, Pseudomonas chlororaphis, Erwinia sp., and Serratia sp. (Table 1) [52,53]. As well as AHLs, a diverse range of signals have been reported in Gram-negative bacteria, i.e., fatty acid methyl esters, 2-alkyl-quinolones, furanone, and γ-butyrolactones [16,50,54].
Another type of QS bacterial signals is antibiotics, which have been reported for intra- or interspecies communication [67]. A recent study revealed some antibiotic signals in a bacterial strain that resided in the tobacco rhizosphere, Lysobacter capsica. The strain possessed biocontrol potential due to its production of antifungal and antibiotic signals, i.e., cyclic lipodepsipeptides and cyclic and polycyclic tetramate macrolactams [68]. Fungi also produce quorum sensing signals that communicate with bacterial species, e.g., γ-heptalactone, tyrosol, γ-butyrolactone, dodecanol, and farnesol [69]. Fungal and bacterial interaction through QS signals creates a competition for existence among the associated microbes in the rhizosphere in terms of infecting the host [70].
Another important class of signaling molecules is volatile organic compounds (VOCs). Microbial VOCs are synthesized and released for long-distance communication among a microbial community and also in microbe–plant interaction [71]. VOCs are low molecular weight lipophilic compounds (100–500 Da) that are produced by various bacterial and fungal species through distinct metabolic pathways that are specific to the species genotype [72]. The VOCs of bacterial origin include alkanes, ketones, alkene, terpenoids, and sulfurs [73,74]. VOCs play an important role in microbe–microbe communication by serving as antimicrobial QS signaling molecules and influencing microbial activity, i.e., virulence, stress resistance, and biofilm formation [75,76]. These attributes are exemplified as a low concentration of nitric oxide (NO) influences the biofilm formation in bacterial species such as P. aeruginosa, B. licheniformis, and S. marcescens [77]. Apart from this, VOC signals have been reported to regulate plant growth (root architecture and hormonal signaling) and plant immunity against biotic and abiotic stresses [71]. Additional studies will further unravel other signaling mechanisms that are associated with microbial VOCs.
Hence, these intricate signaling mechanisms among rhizospheric microorganisms have a key role in the shaping of the rhizomicrobiome by recruiting specific microbes through inter- or intraspecies communication.

2.2. Interkingdom Signaling

Microbes and plants interact with each other via interkingdom signaling, which can influence plant growth by either inducing or suppressing gene expression. Interkingdom signaling can be subdivided into two categories based on the stimulus direction: microbe–plant signaling and plant–microbe signaling.

2.2.1. Microbe–Plant Signaling

In microbe–plant signaling, microbes produce and emit signals that induce symbiotic interactions with the plant. Microbial signals that are of rhizosphere origin can trigger definite changes in the plant transcriptome. As in plants, phytohormones are also produced by PGPR. These plant growth-stimulating signals can regulate the developmental processes of plants and can also provide plants with resistance to abiotic and biotic stresses. Many PGPRs, including Azospirillum spp., Bacillus amyloliquefaciens, B. muralis, B. thuringiensis, Rhizobium spp., and Pseudomonas spp., have been reported to produce a diverse group of phytohormones, e.g., auxin, abscisic acid, salicylic acid, cytokinin, gibberellin, and strigolactones [78,79,80,81]. A recent study in Nature reported three newly isolated strains of Phoma spp. in the rhizosphere of Pinus tabulaeformis. These strains secreted stress the resistance substance abscisic acid (ABA) when under drought stress, which triggered drought resistance mechanisms in the pine tree and also stimulated its antioxidant activities [82].
Most of the literature on microbe–plant communication has been focused on beneficial interactions, such as the induction of plant growth and plant defenses against biotic and abiotic stresses. Rhizospheric microbes that have plant-friendly associations include mycorrhiza, rhizobia, plant growth-promoting bacteria, and fungi (PGPR or PGPF) [11]. Microbial signals are recognized by plants as microbe-associated molecular patterns (MAMPs), i.e., flagellin, chitin, and lipopolysaccharides, via pattern recognition receptors (PRRs) that trigger a local defense through a hormonal signaling network, which in turn produces immune responses [83,84,85].
Quorum sensing signals have also been reported in interkingdom communication. Bacterial QS signals that are perceived by plants elicit various plant responses, e.g., AHLs (N-butanoyl homoserine lactone and N-hexanoyl homoserine lactone), which aid in the development of symbiotic association with plants and enhance root growth by modifying the hormonal levels in the plant [46,86]. Another bacterial QS molecule, DSF, stimulates innate immunity (the detection of harmful microbes via PRRs) in different plants [87]. Some other examples of QS sensing in interkingdom communication is presented in Table 1. As well as causing physiological changes in plants, bacterial QS molecules also provoke the plants into secreting molecules that mimic the QS molecules of pathogenic microbes [88,89]. For example, the p-coumaric acid that is produced by garlic [90], the isothiocyanate sulforaphane that is produced by broccoli [91], the curcumin that is produced by turmeric [92], and the patulin that is produced by fruits (apple, banana, pear, grape, etc.) [93]. Studies on AHL mimicry have shown that the mimicry molecule of bacterial AHL (AHL analog) specifically interferes with QS regulatory pathways and can either stimulate or inhibit the gene expression of the original QS receptor or host, as illustrated in Figure 2 [94,95]. This type of interference in microbial QS signaling for attenuating pathogenicity is known as quorum quenching (QQ) [44,96,97]. A recent study identified a quorum quenching mechanism in a novel bacterial strain, Acinetobacter sp. strain XN-10, which degraded the QS molecules of the AHL family that was acting against the QS-mediated pathogenicity. The strain suppressed the pathogenicity of Pectobacterium carotovorum subsp. carotovorum (Pcc), thereby protecting tissue maceration in potatoes, cabbage, and carrots [98]. Another study reported a quorum quenching defense mechanism against Pseudomonas aeruginosa infection in plants. Upon the interaction of the plant with P. aeruginosa, the plant-secreted QQ molecule (rosmarinic acid) mimicked and competed with the QS pathogenic signal (C4-HSL) and stimulated QS-mediated responses that provide disease protection to the plant, i.e., virulence factors and biofilm formation [88].
Rhizospheric bacteria have been reported to produce antimicrobial signaling compounds. These compounds stimulate systemic resistance in plants against phytopathogens by altering hormonal pathways, e.g., the di-acetyl phloroglucinol (DAPG) that is produced by Pseudomonas sp. [99]. It is common that microbially produced VOCs can induce plant growth and resistance to biotic stress [72,100]. Plants perceive and respond to various VOC signals of PGPR or PGPF origins, such as undecanone and heptanol. For example, 2-heptanol and 2-undecanone, which are produced by B. subtilis and B. amyloliquefaciens, promote the growth of Arabidopsis thaliana when cultivated with these PGPR strains [101,102].

2.2.2. Plant–Microbe Signaling

Plants serve as residences for microbial communities, which include mutualists, commensals, and parasitic microorganisms. Plants respond to the microbial signals by secreting chemical compounds, which are known as root exudates. These include low molecular weight compounds (organic acids, sugar, aliphatic acids, fatty acids, amino acids, flavonoids, and secondary metabolites) and high molecular weight compounds (proteins and mucilage) [20,34,103]. The plant–microbe interaction through the secretion of phytochemicals affects the biology of the rhizosphere [104]. The secreted chemicals attract rhizospheric microbes toward the plant roots and develop either pathogenic or symbiotic interactions with them. The most studied plant–microbe interaction is the symbiotic relationship of legumes with nitrogen-fixing bacteria. The cascade of signals that is produced upon plant–microbe interaction forms root nodules. Rhizospheric microbes feed on these nodules and in turn provide the plant with an available form of nitrogen. Plant signals for nodule formation have been extensively studied over the last decade [15,105].
A recent work demonstrated pathogenic plant–microbe interaction through a broccoli plant induced defense system, which released root exudates upon interaction with fungal species. The root exudates (isothiocyanates (ITC) and glucosinolates (GLS)) showed antifungal potential against the rhizospheric microbes Pseudomonas syringae, Sphingomonas suberifaciens, and Fusarium oxysporum [106]. The chemical complexity and specificity of the root exudates depend on the genotype of the plant. The chemical signals of root exudates attract specific microorganisms for interaction [34,35]. For example, cucumber secretes citric acid, which attracts Bacillus amyloliquefaciens, whereas bananas secrete fumaric acid, which attracts B. subtilis N11, but both interactions result in biofilm formation [107]. Therefore, specific root exudates recruit specific microbial communities within the rhizosphere. Phytochemicals are also capable of stimulating the QS signaling mechanism in microbes, e.g., the flavonoids that are produced by legumes upregulate the gene expression of AHL genes in rhizobia [108].
Plants also release volatile organic compounds (VOCs) from different parts: leaves, flowers, fruits, and roots [109]. These include terpenoids, fatty acids, phenylpropanoids, and amino acids that constitute approx. 1% of the total plant secondary metabolites. VOCs can easily cross plant membranes and be released into the atmosphere or soil. In soil, they attract root colonizing pathogens and inhibit their growth [110].
Plant secretions, including root exudates, volatiles, and strigolactones, are widespread signaling compounds that bind with bacterial receptor proteins and elicit a response in microbes that regulates gene expression. The signaling molecules that have been identified so far are known to influence the developmental and defense processes of plants. These phytochemical compounds are responsible for shaping or adapting the rhizomicrobiome due to their specificity and complexity. However, many of these chemical compounds and their target functions are yet to be explored. The rhizosphere signaling mechanisms that are involved in the interactions between microorganisms and between the microorganisms and plants are represented in Figure 3.

3. Significance of Signaling in Plant–Rhizomicrobiome Interaction

The plant–microbe signaling in the rhizosphere contributes toward sustainable agronomy. Some features of rhizosphere signaling that have significant importance are elaborated here. A summary of the significant factors of rhizosphere signaling is shown in Figure 4.

3.1. Nutrient Acquisition for Phytostimulation

A plant can acquire nutrients from the rhizomicrobiome in two ways: either through symbiotic or non-symbiotic interactions with rhizospheric microorganisms.

3.1.1. Signal-Mediated Symbiosis

Plants are known to develop synergistic interactions with the associated rhizomicrobiome. Nutrient acquisition is a crucial consequence of these interactions as once the symbionts have established the association, a cascade of signals results in the continuous exchange of nutrients between both symbionts. The most studied interkingdom signal-mediated interactions are the mutualistic associations between rhizobia and legumes and between AMF and non-vascular plants. These interactions have been extensively studied throughout the last decade; however, research is still ongoing with regards to signal-mediated symbiosis.
  • Rhizobia–legume symbiotic interactions
Signal-mediated symbiosis plays a key role in nutrient acquisition for plants. Rhizobia provide a reduced form of nitrogen, i.e., ammonia, to plants and in exchange, the plants provide dicarboxylates to the rhizobia [111].
The symbiotic interactions between rhizobia and legumes are initiated after the secretion of phytochemicals by the roots of the host plant, i.e., flavonoid derivatives [112]. Flavonoids have been well studied regarding their role in legume–rhizobia symbiosis. They can accumulate auxin in the root tissues, which facilitates the nodulation process [113]. The flavonoids act as a chemical attractant for rhizobia and guide the rhizobial cells into activating the nodulation genes (nod, noe, and nol), which in turn synthesize nodulation (Nod) factors, i.e., lipo-chitooligosaccharides (LCOs) [114]. LCOs are major signaling molecules for the initiation of nodule formation, which is perceived by receptors that are present in the root epidermis of the host plant (known as kinases). This results in nodule formation in the roots of the legumes [15,115,116]. The legume secretions are rhizobia-specific and attract specific rhizobial species. Likewise, the rhizobia response is specific to the flavonoids, i.e., distinct LCOs diffuse through rhizobia and activate certain nodulation genes that are specific to legume species in order to develop an accurate symbiotic interaction [117,118]. LCOs can promote plant growth by modulating legume root architecture, which in turn enhances nutrient acquisition. This process has been implicated in the increase in the number of root hairs for increased nutrient uptake [119]. This behavior, including an increase in root surface area, length, and hair number, was observed in Arabidopsis thaliana that was inoculated with Bradyrhizobium japonicum [120]. As with flavonoids, some other signaling molecules have also been reported to activate nod genes, e.g., chalcones, betaines, vanillin [121], jasmonate, and aldonic acid [122]. Thus, these signaling compounds regulate the gene expression of particular partner symbionts for successful symbiosis.
The neighboring microbes of rhizobia in a phytomicrobiome assist in nodule formation by producing synergistic consortia that play an important role in the alleviation of abiotic stress, nutrient uptake, and disease resistance [118]. The co-inoculation of PGPR has shown drastic plant growth-promoting effects, e.g., Mesorhizobium cicero that was co-inoculated with Bacillus sp. and Enterobacter aerogenes demonstrated significantly enhanced root nodulation, uptake of nitrogen and phosphorous, and total protein content in the chickpea plant [123].
  • Mycorrhizal symbiotic interactions
Mycorrhizal symbiosis involves the interactions between arbuscular mycorrhizal fungi (AMF) and land plants. The signaling pathways are quite similar to those in rhizobia–legume symbiosis. Plant secretions of strigolactones, which are known as branching factors (BFs), send signals to the AMF that, upon recognition, release mycorrhizal lipo-chitooligosaccharides (Myc factors) that are responsible for the development of the symbiotic interactions between plants and AMF [124,125]. As well as acting as AMF stimuli, strigolactones are plant hormones that are capable of interfering with auxin transport [126]. Intriguingly, AMF penetrate through the epidermal cells of plant roots after perceiving plant signals. There, they start to differentiate into highly branched structures called arbuscles. AMF colonize, either intercellularly or intracellularly, within the cortical tissues that have a key role in the nutrient exchange between both symbionts [127,128]. A complex cascade of signals enhances the nutrient uptake of the roots, such as nitrates, amino acids, immobile ortho-phosphates, starch, and calcium, using specialized transporters. This improves the metabolic pathways of the plant while the host plant fulfills the carbohydrate requirements of the AMF in return [129,130,131].
Recently, studies have reported a tripartite interaction between three symbionts: AFM–rhizobia–plant (ARP). This tripartite interaction increases nitrogen fixation, helps the plant to survive in drought conditions [132,133], and enhances root soluble sugar content and dry weight, as well as increasing the total number of nodules [133,134]. Some of the literature also considered tripartite symbiosis to be an efficient interaction for plant growth [98,135,136,137]. Further research is required to understand the gene expression profiles and metabolic events that are associated with ARP tripartite interactions. Meta-genomics would provide better ways to monitor and understand the role of each partner that is involved in the symbioses within the rhizosphere.

3.1.2. Nutrient Acquisition without Symbiosis

Plants can acquire nutrients without symbiosis with rhizobia or AMF. Various microbes meet nutrient requirements via the production of certain metabolites [19,138,139]. During iron deficiency, certain microbes colonize the roots of a plant in iron deficit soil and produce chelating compounds to sequester iron from the soil. Bacterial colonization induces the gene expression that is responsible for iron uptake in plants in response to iron deficiency [139,140,141]. The phosphate requirements of the arabidopsis plant are provided by endophytic bacteria Colletotrichum tofieldiae via the translocation of phosphate through the roots [142]. The rhizospheric and endophytic consortium of Bacillus sp. and Pseudomonas sp. Has been shown to significantly enhance the phosphate solubilization efficiency of wheat cultivars that were growing in phosphate deficient soil [143]. Similarly, microbes other than rhizobia meet the nitrogen demands of non-leguminous plants [19,144]. The decomposition of organic matter by rhizospheric microbes enhances plant productivity and soil fertility by providing nutrients to plants. For example, lignocellulolytic fungi, such as, Pleurotus ostreatus, Phanerochaete, and Trichoderma harzianum, and bacteria, including Pseudomonas sp., Sporocytophaga sp., Cytophaga sp., and Streptomyces sp., degrade plant biomass and release nutrients into the soil, which are taken up by the plants that are growing in nutrient-poor soil [145,146]. The soil phages or virus-like particles (VLPs) affect the availability of nutrients to plants by modulating mutations in microbial phylotypes or interfering with the structure and diversity of the bacterial population. These effects generate mutations in rhizodeposits and aid in nutrient uptake (N, S, and P) by the plants. Despite the lower abundance of the lysogenic virus community than the bacterial community, viruses play their role in the resilience of the rhizosphere microbiome [30].

3.2. Regulation of Plant Immunity against Phytopathogens

Biotic stresses cause a 30% loss in crop yield and a 15% loss in food worldwide due to various pathogen attacks, i.e., bacteria, fungi, viroids, viruses, protists, nematodes, and insects [147]. PGPR induce an innate immune response in plants against the phytopathogens in a sequential manner, i.e.: (i) perceive the stress stimulus; (ii) regulate defense responses by activating defense-related signaling pathways; and (iii) defense priming that prepares the plant to respond against the stress after recognition by the PGPR [148,149]. The perception of the stimulus by the plant triggers a local immune defense response that is translated into a systemic defense response, which is mediated by hormonal signaling pathways [150].

3.2.1. Pattern-Triggered Immunity (PTI)

Plant immunity is initiated by host–pathogen interactions, in which plant receptors and pattern recognition receptors (PRRs) recognize the pathogen- or microbe-associated molecular patterns (PAMPs or MAMPs) that induce the resistance mechanisms. After PAMP recognition, the plant triggers signaling modules, i.e., mitogen-activated protein kinases (MAPKs) and calcium-dependent protein kinases (CDPKs), which further triggers a signal cascade that is responsible for the activation of specific transcription factors that lead to the induction of multiple intracellular defense responses. The induction of the signal cascade is either hormone-dependent (MAPKs and CDPKs) or -independent [12,151,152]. The plant signals that make up the signal cascade include methyl jasmonate (MeJA), azelaic acid (AzA), pipecolic acid (Pip), and salicylic acid (SA) signals that regulate defense genes [153,154]. Some phytohormones, such as auxins, cytokinins, gibberellins, and abscisic acid, also take part in this signal cascade [28]. This type of defense response is known as MAMP- or PAMP-triggered immunity [155,156,157] or more generally as pattern-triggered immunity (PTI) [158], which is depicted in Figure 5.
The induction of PTI stimulates intracellular defense responses, including stomatal closure, the deposition of callose, and the induction of reactive oxygen species (ROS) and antimicrobial metabolites (Figure 5) [158,159]. Several AHLs and diffusible signal factors (DSFs) have been identified as quorum sensing signals in phytopathogenic interactions. In the host–pathogen interaction between Arabidopsis thaliana and Sclerotinia sclerotirum, small sRNAs were abundantly identified in the signal pathways [160]. Plants produce secondary metabolites upon pathogen interaction, which are known as “phytoanticipins” or “phytoalexins” (Figure 5) [157,161]. The biosynthesis of these inducible compounds is regulated by phytohormone signaling pathways, phosphorylation, and defense-related genes [162]. Capsidiol production in Nicotiana benthamiana has been shown to be mediated by an ethylene signaling pathway that provides defense against Phytophthora infestans [163].

3.2.2. Effector-Triggered Immunity (ETI)

In response to PTI signaling, pathogens resist the plant defense mechanisms by producing effectors that prevent their detection by the host plant. This response mechanism is known as effector-triggered susceptibility (ETS) [83,155,164]. Upon ETS activation, the plant in turn develops another defense layer, which is known as effector-triggered immunity (ETI). ETI is activated upon the detection of pathogen effectors through a specialized intracellular protein, which is known as NB-LRR (nucleotide-binding and leucine-rich repeat domains) [83,156,165]. NB-LRR is encoded by a plant resistance gene (R gene) that mediates resistance to specific phytopathogens, i.e., viruses, oomycetes, fungi, nematodes, and bacteria. It recognizes the specific pathogen effectors either by direct binding or by changing its protein binding site during pathogen–effector binding [165], as depicted in Figure 5. The virulence specificity of pathogen effectors has been as yet undiscovered because several PRRs can recognize specific pathogens and pathogens have evolved effectors that dislocate the defense signaling of plants [166]. This protective mechanism is called the hypersensitive response (HR) [167]. For example, Phytophthora infestans produces the effector protein (PexRD2), which increases its susceptibility to Nicotiana benthamiana by suppressing the host’s defense signaling [168]. Effector proteins have also been reported to attenuate hormonal signaling, i.e., salicylic acid and jasmonic acid signals induce P. infestans colonization and infection in plants [169].

3.3. Defense Priming

Plants use adaptive strategies to enhance the defensive capacity of their innate resistance. Warning signals or stimuli, such as pathogens, abiotic stresses, chemicals, PGPR, and arthropods, trigger their innate resistance and prime the plants against secondary stresses. This type of enhanced defense response state is known as priming. During this phase, plants change at transcriptional, physiological, metabolic, and epigenetic levels [23]. Primed plants show induced and more competent responses against stresses [170,171]. Defense priming can be sustained throughout the life of a plant and can be transferred on to the next generation. Defense priming is mediated by two induced signaling resistance pathways: systemic acquired resistance and induced systemic resistance.
  • Systemic acquired resistance (SAR)
PTI and ETI trigger long-distance defense signals that induce resistance against the future localized infection via broad-spectrum phytopathogens, which involves the activation of proteins and pathogenesis-related (PR) genes. This type of long-distance immune priming that is induced by a pathogen is known as systemic acquired resistance (SAR). It is a salicylic acid (SA)-dependent pathway [83,153,172]. Salicylic acid is produced by various rhizobacteria that are responsible for activating PR genes, PTI, and ETI in SAR signaling [173].
SAR pathways are regulated by the key protein NPR1 (non-repressor of PR genes), which interacts with transcriptional co-factors that in turn induce the expression of pathogen-related proteins that are associated with systemic resistance [174]. SA binds with NPR1 and induces conformational changes in its structure to further expose it to transcriptional factors for PR gene activation (Figure 6) [175]. Jasmonic acid (JA) pathways in SAR protect the plant from necrotrophic pathogens by activating a regulatory protein: COI1 in the form of a ubiquitin complex. The activated protein complex participates in the defense of the plant [174].
SA pathways have been reported to induce the biosynthesis of flavonoids, i.e., pro-anthocyanidins and catechin, which inhibit the proliferation of foliar rust fungus Melampsora larici-populina in poplar trees [176]. SA and methyl jasmonate (MeJA) have been shown to mediate the induction of defense-related genes and proteins against the blight disease-causing pathogen Xanthomonas axonopodis pv. manihotis in cassava plants [177]. Trichoderma harzianum has also been shown to induce systemic resistance via ET, SA, and JA pathways against the cucumber mosaic virus in tomato plants [178]. Trichoderma viride has been reported to provide resistance against Phytophthora infestans infection via SA signaling in potato plants [179].
  • Induced systemic resistance (ISR)
The induction of defense signaling by a beneficial rhizomicrobiome, such as PGPR against phytopathogens and pests, is known as induced systemic resistance (ISR), which is an SA-independent pathway [83,180]. This defense signaling pathway also increases plants’ tolerance of abiotic and biotic stresses and is therefore also known as induced systemic tolerance [181,182]. The core difference between ISR and SAR is the dependency of SAR on the SA synthesis and accumulation by PGPR that activates SA signaling pathways, whereas ISR is SA-independent and includes ET and JA signaling pathways (Figure 6) [14].
Beneficial microbes do not possess virulence genes but they do have the potential to sensitize plants by exploiting them to perceive MAMPs that lead to the activation of ISR [183]. Defense priming by rhizobacteria overlaps under biotic and abiotic stresses.
ISR is regulated by the signaling pathways of phytohormones, including ethylene (ET) and jasmonic acid (JA) [184,185]. The reactive oxygen species (ROS) and reactive nitrogen oxygen species (RNOS) form a complex defense signaling network by stimulating the biosynthesis of JA, ET, and SA [186]. As an example, JA and ET were shown to regulate a defense response by stimulating the lipid transfer protein gene against Fusarium graminearum [187]. Another defense response that is induced by JA and ET includes the RSL1 (recombinant Solanum lycopersicum) protein that triggers the gene expression of ET and JA against phytopathogens, e.g., Alternaria alternata and Botrytis cinerea [188]. ISR expression is usually determined by the type of bacterial strain and the plant in which resistance is induced [189]. PGPR consortiums have also been reported to induce ISR. For instance, a PGPR consortium was shown to trigger ISR in the cucumber plant by increasing its antioxidants and root vigor index while also regulating the photosynthetic machinery [190]. Moreover, the co-inoculation of AMF with Pseudomonas sp. reportedly enhanced the antioxidant machinery in the leaves and the phosphatase activity in the roots of lettuce plants under drought stress [191].
PGPR products that elicit defense priming include VOCs, LOCs, AHLs, ACC deaminase, antibiotics, siderophores, and lipopolypeptides [192,193]. The literature has reported that Bacillus amyloliquefaciens produced lipopolypeptides that primed the lettuce plant against Rhizoctonia solani [194]. AHLs from Acidovorax radicus N35 reportedly primed the defense response by stimulating the production of flavonoids, i.e., lutonarin and saponarin, in barley [195]. Furthermore, a siderophore-producing Pseudomonas putida mutant strain was shown to have an efficient biocontrol ability over the siderophore-deficient Pseudomonas aeruginosa mutant strain in controlling Fusarium wilt in tomato plants [196].
Many rhizobacteria have been reported to have parasitic traits due to the production of cell wall degrading enzymes, e.g., Stenotrophomonas maltophilia secretes proteases that have a parasitic action on Bursaphelenchus xylophilus (nematode) and P. ultimum (oomycetes) [197,198]. Some phages play a role in controlling the infection of a plant through bacterial phytopathogens. A recent study that was published in Nature reported that a combination of phages lessened the infection of tomato plants that was caused by Ralstonia solanacearum [199].

3.4. Regulation of Plant Defenses against Abiotic Stresses

Plants experience about 70% damage due to various abiotic stresses, including temperature, drought, nutrient deficiency, salinity, and heavy metal toxicity. A rhizomicrobiome manages to ameliorate abiotic stresses through induced systemic tolerance via microbe–plant signaling, which involves biochemical and physiological changes in the plants [200]. Many PGPR have been reported to ameliorate abiotic stresses by adopting defense-related mechanisms, i.e., biofilm formation or the production of phytohormones, antibiotics, siderophores, hydrolytic enzymes, and quorum quenching compounds [201,202]. Rhizobacteria that possess the genetic capability to mitigate abiotic stresses include Caulobacter, Rhizobia, Serratia, Flavobacterium, Erwinia, Chromobacterium, Burkholderia, Methylobacterium, Trichoderma, Micrococcus, and Pseudomonas [203,204,205].
Plants manage to respond to abiotic stresses through a complex signal cascade that is activated following the perception of stress stimuli by the receptors or sensors that are in the plant cell walls. The sensed signals are then translated into intracellular signals through secondary messengers, i.e., calcium ions, nitric oxide, sugars, cyclic nucleotides, inositol phosphate, and ROS. These secondary messengers transduce the signaling pathways of the stress signals [206,207]. As a result, various physiological and metabolic responses become activated during different stages of plant development, such as the synthesis of jasmonic acid, ethylene, salicylic acid, abscisic acid [208,209], flavonoids, phenolic compounds [210], antioxidants, and osmolytes, as well as the activation of certain transcription factors for the expression of stress genes [211,212].
PGPR-produced phytohormones, mainly ethylene (ET), jasmonate (JT), salicylic acid (SA), abscisic acid (ABA), cytokinin, auxin, and gibberellins, regulate abiotic stress responses in plants. These hormones interact either synergistically or antagonistically to adjust their biosynthesis according to the stress response [213,214]. ABA biosynthesis genes are triggered upon osmotic stress to enhance its production, which is responsible for maintaining water balance through the closure of stomata [215]. ABA acts synergistically with SA and antagonistically with ET to regulate osmotic stress, stomatal movement, and leaf senescence under drought conditions [216]. The exact mechanisms of bacterially produced phytohormones need to be explored further. However, some of the reported hormonal signaling pathways of rhizo-microorganisms in different host plants are presented in Table 2. PGPR can effectively stimulate antioxidant activity in plants under drought and salinity conditions. As a result, plants increase their antioxidant signaling, i.e., the production of redox enzymes and phenolic-like compounds (glycine, proline, and betaine) [217].
Enzymatic pathways, including phosphatases and protein kinases, i.e., calcium-dependent protein kinases (CDPKs), mitogen-activated protein kinases (MAPKs), and receptor-like kinases, have been reported to regulate dephosphorylation and phosphorylation in the management of abiotic stresses in plants [241]. Plants can deregulate the salt-sensitive gene expression under salinity stress via calcium ion signaling pathways [242]. PGPR-mediated stress resistance is involved in these signaling pathways that modulate the transcription factors in order to enhance the expression of stress-responsive genes [232]. For instance, Bacillus amyloliquefaciens has been shown to ameliorate the salt stress response in Oryza sativa by upregulating MAPK and CDPK pathways [243]. Another study showed an improved salinity stress resistance in wheat plants following inoculation with bacterial strains, i.e., Enterobacter cloacae, Pseudomonas fluorescens, Pseudomonas putida, and Serratia ficaria, which elevated the Na+/K+ ratio [244].

4. Conclusions and Future Perspectives

The interaction intricate chemical signaling between plants and their associated microbiome deciphers the communication dynamics that are underway within the rhizosphere and their outcomes in terms of plant productivity and development. Over the last decade, quorum sensing has been studied in the gene expression of both beneficial and harmful interactions between inter- and intramicrobial species and in the interkingdom signaling between plants and PGPR, i.e., nitrogen-fixing bacteria and rhizobia. Most of the plant metabolites that have been reported to date have been characterized by the Arabidopsis thaliana plant model. The recent literature suggests that other plants need to be investigated for the analysis of more plant metabolites and chemical cues. Expanding the research on rhizosphere communities will aid in the discovery of new chemical cues and their potential to enhance plant productivity in terms of symbiosis, resistance against environmental stresses, and immunity or defense mechanisms toward phytopathogens. Based on experimental evidence, there is no doubt that plants create an environment that recruits a specific rhizomicrobiome and shape microbial associations that are beneficial for their growth. Therefore, the escalating pressure of the demand for high crop production has opened up new avenues for research on rhizosphere signaling to reduce dependency on synthetic processes and agrochemicals. The alterations in the metabolic pathways of plants, microbes, rhizodeposits, and signaling molecules could be effective in the development of a rhizomicrobiome that is beneficial for effective plant productivity and resistance toward biotic and abiotic stresses, which could ultimately lead to more sustainable agriculture. Moreover, future research can be expanded by using novel “multi-omics techniques”, which encompass genomics, proteomics, transcriptomics, phenomics, metagenomics, metabolomics, metatranscriptomics, metaproteomics, and metagenomics and could reveal multi-layered information about new chemical cues, cellular mechanisms, signaling mechanisms, and plant genes. Furthermore, future research on rhizosphere engineering is essential to dissect the targeted outcomes of the beneficial plant–microbe communications that could broaden the application of sustainable agriculture.

Author Contributions

Conceptualization, F.J., H.M. and M.F.; validation, H.M. and L.D.; writing—original draft preparation, F.J.; writing—review and editing, H.M., F.J., L.D. and M.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

Laurent Dufossé deeply acknowledges the strong financial support from the Conseil Régional de Bretagne and the Conseil Régional de La Réunion for research activities that are dedicated to microbial biotechnology.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

PGPRPlant growth-promoting rhizobacteria
N-AHLN-Acyl homoserine lactone
AIsAutoinducers
QSQuorum sensing
QQQuorum quenching
VOCVolatile organic compound
MAMPMicrobe-associated molecular pattern
PRRPattern recognition receptor
DSFDiffusible factor
ITCIsothiocyanate
GLSGlucosinolates
AMFArbuscular mycorrhizal fungus
NodNodulation
LCOsLipo-chitooligosaccharides
BFBranching factor
PAMPPathogen-associated molecular pattern
MAPKsMitogen-activated protein kinases
CDPKsCalcium-dependent protein kinases
PTIPattern-triggered immunity
ROSReactive oxygen species
ETSEffector-triggered susceptibility
ETIEffector-triggered immunity
NB-LRRsNucleotide binding and leucine-rich repeat domains
HRHypersensitive response
SASalicylic acid
NPR1Non-expressor of pathogenesis-related genes
SARSystemic acquired resistance
JAJasmonic acid
ETEthylene
MeJAMethyl jasmonate
AzAAzelaic acid
PipPipecolic acid
ISRInduced systemic resistance
ACC1-Aminocyclopropane-1-carboxylic acid
ABAAbscisic acid
AOXAlternative oxidases
COXCytochrome oxidase
GAGibberellin
SOSSalt overly sensitive

References

  1. Checcucci, A.; Marchetti, M. The Rhizosphere Talk Show: The Rhizobia on Stage. Front. Agron. 2020, 2, 591494. [Google Scholar] [CrossRef]
  2. Hartmann, A.; Schmid, M.; Van Tuinen, D.; Berg, G. Plant-Driven selection of microbes. Plant Soil 2009, 321, 235–257. [Google Scholar] [CrossRef]
  3. Bais, H.P.; Weir, T.L.; Perry, L.G.; Gilroy, S.; Vivanco, J.M. The role of root exudates in rhizosphere interactions with plants and other organisms. Annu. Rev. Plant Biol. 2006, 57, 233–266. [Google Scholar] [CrossRef] [Green Version]
  4. Martin, B.D.; Schwab, E. Current Usage of Symbiosis and Associated Terminology. Int. J. Biol. 2013, 5, 32. [Google Scholar] [CrossRef]
  5. Zolla, G.; Badri, D.V.; Bakker, M.; Manter, D.K.; Vivanco, J.M. Soil microbiomes vary in their ability to confer drought tolerance to Arabidopsis. Appl. Soil Ecol. 2013, 68, 1–9. [Google Scholar] [CrossRef]
  6. Selvakumar, G.; Panneerselvam, P.; Ganeshamurthy, A.N. Bacterial mediated alleviation of abiotic stress in crops. In Bacteria in Agrobiology: Stress Management; Springer: Berlin/Heidelberg, Germany, 2012; pp. 205–224. [Google Scholar] [CrossRef]
  7. Mendes, R.; Kruijt, M.; de Bruijn, I.; Dekkers, E.; Van Der Voort, M.; Schneider, J.H.; Piceno, Y.M.; DeSantis, T.Z.; Andersen, G.L.; Bakker, P.A.; et al. Deciphering the rhizosphere microbiome for disease-suppressive bacteria. Science 2011, 332, 1097–1100. [Google Scholar] [CrossRef]
  8. Sarma, B.K.; Yadav, S.K.; Singh, S.; Singh, H.B. Microbial consortium-mediated plant defense against phytopathogens: Readdressing for enhancing efficacy. Soil Biol. Biochem. 2015, 87, 25–33. [Google Scholar] [CrossRef]
  9. Venturi, V.; Fuqua, C. Chemical signalling between plants and plant-pathogenic bacteria. Annu. Rev. Phytopathol. 2013, 51, 17–37. [Google Scholar] [CrossRef]
  10. Lareen, A.; Burton, F.; Schäfer, P. Plant root-microbe communication in shaping root microbiomes. Plant Mol. Biol. 2016, 90, 575–587. [Google Scholar] [CrossRef] [Green Version]
  11. Venturi, V.; Keel, C. Signaling in the rhizosphere. Trends Plant Sci. 2016, 21, 187–198. [Google Scholar] [CrossRef]
  12. Zipfel, C.; Oldroyd, G.E.D. Plant signalling in symbiosis and immunity. Nature 2017, 543, 328–336. [Google Scholar] [CrossRef]
  13. Chagas, F.O.; Pessotti, R.D.C.; Caraballo-Rodríguez, A.M.; Pupo, M.T. Chemical signaling involved in plant–microbe interactions. Chem. Soc. Rev. 2018, 47, 1652–1704. [Google Scholar] [CrossRef]
  14. Bukhat, S.; Imran, A.; Javaid, S.; Shahid, M.; Majeed, A.; Naqqash, T. Communication of plants with microbial world: Exploring the regulatory networks for PGPR mediated defense signaling. Microbiol. Res. 2020, 238, 126486. [Google Scholar] [CrossRef]
  15. Oldroyd, G.E.D. Speak, friend and enter: Signalling systems that promote beneficial symbiotic associations in plants. Nat. Rev. Microbiol. 2013, 11, 252–263. [Google Scholar] [CrossRef]
  16. Borges, A.; Simões, M. Quorum Sensing Inhibition by marine bacteria. Mar. Drugs 2019, 17, 427. [Google Scholar] [CrossRef] [Green Version]
  17. Cheng, F.; Ma, A.; Zhuang, G.; Fray, R. Exogenous N-acyl-homoserine lactones enhance the expression of flagella of Pseudomonas syringae and activate defence responses in plants. Mol. Plant Pathol. 2018, 19, 104–115. [Google Scholar] [CrossRef]
  18. Schenk, S.T.; Hernández-Reyes, C.; Samans, B.; Stein, E.; Neumann, C.; Schikora, M.; Reichelt, M.; Mithöfer, A.; Becker, A.; Kogel, K.H.; et al. N-acyl-homoserine lactone primes plants for cell wall reinforcement and induces resistance to bacterial pathogens via the salicylic acid/oxylipin pathway. Plant Cell 2014, 26, 2708–2723. [Google Scholar] [CrossRef] [Green Version]
  19. Jacoby, R.; Peukert, M.; Succurro, A.; Koprivova, A.; Kopriva, S. The Role of Soil Microorganisms in plant mineral nutrition—Current Knowledge and Future Directions. Front. Plant Sci. 2017, 8, 1617. [Google Scholar] [CrossRef] [Green Version]
  20. Bais, H.P.; Park, S.-W.; Weir, T.L.; Callaway, R.M.; Vivanco, J.M. How plants communicate using the underground information superhighway. Trends Plant Sci. 2004, 9, 26–32. [Google Scholar] [CrossRef]
  21. Li, Y.; Kong, D.; Fu, Y.; Sussman, M.R.; Wu, H. The effect of developmental and environmental factors on secondary metabolites in medicinal plants. Plant Physiol. Biochem. 2020, 148, 80–89. [Google Scholar] [CrossRef]
  22. Rudrappa, T.; Czymmek, K.J.; Paré, P.W.; Bais, H.P. Root-Secreted malic acid recruits beneficial soil bacteria. Plant Physiol. 2008, 148, 1547–1556. [Google Scholar] [CrossRef] [Green Version]
  23. Mauch-Mani, B.; Baccelli, I.; Luna, E.; Flors, V. Defense priming: An adaptive part of induced resistance. Annu. Rev. Plant Biol. 2017, 68, 485–512. [Google Scholar] [CrossRef] [Green Version]
  24. Hilker, M.; Schmülling, T. Stress priming, memory, and signalling in plants. Plant Cell Environ. 2019, 42, 753–761. [Google Scholar] [CrossRef]
  25. Bakker, P.A.H.M.; Berendsen, R.L.; Doornbos, R.F.; Wintermans, P.C.A.; Pieterse, C.M.J. The rhizosphere revisited: Root microbiomics. Front. Plant Sci. 2013, 4, 165. [Google Scholar] [CrossRef] [Green Version]
  26. Miller, J.B.; Oldroyd, G.E.D. The role of diffusible signals in the establishment of rhizobial and mycorrhizal symbioses. In Signalling and Communication in Plant Symbiosis; Springer: Berlin/Heidelberg, Germany, 2011; pp. 1–30. [Google Scholar] [CrossRef]
  27. Berendsen, R.L.; Pieterse, C.M.J.; Bakker, P.A.H.M. The rhizosphere microbiome and plant health. Trends Plant Sci. 2012, 17, 478–486. [Google Scholar] [CrossRef]
  28. Pieterse, C.M.J.; Leon-Reyes, A.; Van der Ent, S.; Van Wees, S.C.M. Networking by small-molecule hormones in plant immunity. Nat. Chem. Biol. 2009, 5, 308–316. [Google Scholar] [CrossRef] [Green Version]
  29. Olanrewaju, O.S.; Ayangbenro, A.S.; Glick, B.R.; Babalola, O.O.; Ayangbenro, A. Plant health: Feedback effect of root exudates-rhizobiome interactions. Appl. Microbiol. Biotechnol. 2019, 103, 1155–1166. [Google Scholar] [CrossRef] [Green Version]
  30. Pratama, A.A.; Terpstra, J.; de Oliveria, A.L.M.; Salles, J.F. The Role of Rhizosphere Bacteriophages in Plant Health. Trends Microbiol. 2020, 28, 709–718. [Google Scholar] [CrossRef]
  31. Sharma, R.S.; Nayak, S.; Malhotra, S.; Karmakar, S.; Sharma, M.; Raiping, S.; Mishra, V. Rhizosphere provides a new paradigm on the prevalence of lysogeny in the environment. Soil Tillage Res. 2019, 195, 104368. [Google Scholar] [CrossRef]
  32. Li, Y.; Sun, H.; Yang, W.; Chen, G.; Xu, H. Dynamics of Bacterial and Viral Communities in Paddy Soil with Irrigation and Urea Application. Viruses 2019, 11, 347. [Google Scholar] [CrossRef] [Green Version]
  33. Simon, J.-C.; Marchesi, J.; Mougel, C.; Selosse, M.-A. Host-Microbiota interactions: From holobiont theory to analysis. Microbiome 2019, 7, 5. [Google Scholar] [CrossRef]
  34. Sasse, J.; Martinoia, E.; Northen, T. Feed Your Friends: Do plant exudates shape the root microbiome? Trends Plant Sci. 2018, 23, 25–41. [Google Scholar] [CrossRef] [Green Version]
  35. Mommer, L.; Kirkegaard, J.; van Ruijven, J. Root–Root Interactions: Towards a rhizosphere framework. Trends Plant Sci. 2016, 21, 209–217. [Google Scholar] [CrossRef]
  36. Rosier, A.; Bishnoi, U.; Lakshmanan, V.; Sherrier, D.J.; Bais, H.P. A perspective on inter-kingdom signaling in plant–beneficial microbe interactions. Plant Mol. Biol. 2016, 90, 537–548. [Google Scholar] [CrossRef]
  37. Hein, J.W.; Wolfe, G.V.; Blee, K.A. Comparison of Rhizosphere Bacterial Communities in Arabidopsis thaliana mutants for systemic Acquired Resistance. Microb. Ecol. 2008, 55, 333–343. [Google Scholar] [CrossRef]
  38. Schlaeppi, K.; Bulgarelli, D. The Plant Microbiome at Work. Mol. Plant-Microbe Interact. 2015, 28, 212–217. [Google Scholar] [CrossRef] [Green Version]
  39. Yang, C.-H.; Crowley, D.E. Rhizosphere microbial community structure in relation to root location and plant iron nutritional status. Appl. Environ. Microbiol. 2000, 66, 345–351. [Google Scholar] [CrossRef] [Green Version]
  40. Malusà, E.; Pinzari, F.; Canfora, L. Efficacy of biofertilizers: Challenges to improve crop production. In Microbial Inoculants in Sustainable Agricultural Productivity; Springer: New Delhi, India, 2016; pp. 17–40. [Google Scholar] [CrossRef]
  41. Jones, P.; Garcia, B.; Furches, A.; Tuskan, G.A.; Jacobson, D. Plant host-associated mechanisms for microbial selection. Front. Plant Sci. 2019, 10, 862. [Google Scholar] [CrossRef] [Green Version]
  42. Mendes, R.; Garbeva, P.; Raaijmakers, J.M. The rhizosphere microbiome: Significance of plant beneficial, plant pathogenic, and human pathogenic microorganisms. FEMS Microbiol. Rev. 2013, 37, 634–663. [Google Scholar] [CrossRef]
  43. Bai, Y.; Müller, D.B.; Srinivas, G.; Garrido-Oter, R.; Potthoff, E.; Rott, M.; Dombrowski, N.; Münch, P.C.; Spaepen, S.; Remus-Emsermann, M.; et al. Functional overlap of the Arabidopsis leaf and root microbiota. Nature 2015, 528, 364–369. [Google Scholar] [CrossRef]
  44. Hong, K.W.; Koh, C.L.; Sam, C.-K.; Yin, W.-F.; Chan, K.-G. Quorum Quenching Revisited—From Signal Decays to Signalling Confusion. Sensors 2012, 12, 4661–4696. [Google Scholar] [CrossRef] [Green Version]
  45. Helman, Y.; Chernin, L. Silencing the mob: Disrupting quorum sensing as a means to fight plant disease. Mol. Plant Pathol. 2015, 16, 316–329. [Google Scholar] [CrossRef]
  46. Schikora, A.; Schenk, S.T.; Hartmann, A. Beneficial effects of bacteria-plant communication based on quorum sensing molecules of the N-acyl-homoserine lactone group. Plant Mol. Biol. 2016, 90, 605–612. [Google Scholar] [CrossRef]
  47. Hassan, R.; Shaaban, M.I.; Bar, F.A.; El-Mahdy, A.; Shokralla, S. Quorum sensing inhibiting activity of Streptomyces coelicoflavus Isolated from Soil. Front. Microbiol. 2016, 7, 659. [Google Scholar] [CrossRef] [Green Version]
  48. Chu, W. Role of the quorum-sensing system in biofilm formation and virulence of Aeromonas hydrophila. Afr. J. Microbiol. Res. 2011, 5, 5819–5825. [Google Scholar] [CrossRef] [Green Version]
  49. Atkinson, S.; Williams, P. Quorum sensing and social networking in the microbial world. J. R. Soc. Interface 2009, 6, 959–978. [Google Scholar] [CrossRef] [Green Version]
  50. Vendeville, A.; Winzer, K.; Heurlier, K.; Tang, C.M.; Hardie, K. Making ‘sense’ of metabolism: Autoinducer-2, LUXS and pathogenic bacteria. Nat. Rev. Genet. 2005, 3, 383–396. [Google Scholar] [CrossRef]
  51. Winzer, K. Bacterial cell-to-cell communication: Sorry, can’t talk now—Gone to lunch! Curr. Opin. Microbiol. 2002, 5, 216–222. [Google Scholar] [CrossRef]
  52. Zhang, W.; Li, C. Exploiting Quorum Sensing Interfering Strategies in Gram-negative bacteria for the enhancement of environmental applications. Front. Microbiol. 2016, 6, 1535. [Google Scholar] [CrossRef]
  53. Ferluga, S.; Steindler, L.; Venturi, V. N-acyl homoserine lactone quorum sensing in Gram-negative rhizobacteria. In Secondary Metabolites in Soil Ecology; Springer: Berlin/Heidelberg, Germany, 2008; pp. 69–90. [Google Scholar] [CrossRef]
  54. Sindhu, S.S.; Sehrawat, A.; Sharma, R.; Dahiya, A.; Khandelwal, A. Belowground microbial crosstalk and rhizosphere biology. In Plant-Microbe Interactions in Agro-Ecological Perspectives; Springer: Singapore, 2017; pp. 695–752. [Google Scholar] [CrossRef]
  55. Aigle, B.; Corre, C. Waking up Streptomyces Secondary Metabolism by Constitutive Expression of Activators or Genetic Disruption of Repressors. Methods Enzymol. 2012, 517, 343–366. [Google Scholar] [CrossRef]
  56. Ulrich, K.; Kube, M.; Becker, R.; Schneck, V.; Ulrich, A. Genomic analysis of the endophytic Stenotrophomonas strain 169 reveals features related to plant-growth promotion and stress tolerance. Front. Microbiol. 2021, 12, 1542. [Google Scholar] [CrossRef]
  57. Xiong, Q.; Liu, D.; Zhang, H.; Dong, X.; Zhang, G.; Liu, Y.; Zhang, R. Quorum sensing signal autoinducer-2 promotes root colonization of Bacillus velezensis SQR9 by affecting biofilm formation and motility. Appl. Microbiol. Biotechnol. 2020, 104, 7177–7185. [Google Scholar] [CrossRef]
  58. Zhao, Q.; Yang, X.-Y.; Li, Y.; Liu, F.; Cao, X.-Y.; Jia, Z.-H.; Song, S.-S. N-3-oxo-hexanoyl-homoserine lactone, a bacterial quorum sensing signal, enhances salt tolerance in Arabidopsis and wheat. Bot. Stud. 2020, 61, 8–12. [Google Scholar] [CrossRef]
  59. Jung, B.K.; Khan, A.R.; Hong, S.-J.; Park, G.-S.; Park, Y.-J.; Kim, H.-J.; Jeon, H.-J.; Khan, M.A.; Waqas, M.; Lee, I.-J.; et al. Quorum sensing activity of the plant growth-promoting rhizobacterium Serratia glossinae GS2 isolated from the sesame (Sesamum indicum L.) rhizosphere. Ann. Microbiol. 2017, 67, 623–632. [Google Scholar] [CrossRef]
  60. Liu, X.; Bimerew, M.; Ma, Y.; Müller, H.; Ovadis, M.; Eberl, L.; Berg, G.; Chernin, L. Quorum-Sensing signaling is required for production of the antibiotic pyrrolnitrin in a rhizospheric biocontrol strain of Serratia plymuthica. FEMS Microbiol. Lett. 2007, 270, 299–305. [Google Scholar] [CrossRef] [Green Version]
  61. Barriuso, J.; Solano, B.R.; Fray, R.G.; Cámara, M.; Hartmann, A.; Mañero, F.J.G. Transgenic tomato plants alter quorum sensing in plant growth-promoting rhizobacteria. Plant Biotechnol. J. 2008, 6, 442–452. [Google Scholar] [CrossRef]
  62. Ryu, C.-M.; Choi, H.K.; Lee, C.-H.; Murphy, J.F.; Lee, J.-K.; Kloepper, J.W. Modulation of Quorum Sensing in Acyl-homoserine Lactone-Producing or -Degrading Tobacco Plants Leads to Alteration of Induced Systemic Resistance Elicited by the Rhizobacterium Serratia marcescens 90–166. Plant Pathol. J. 2013, 29, 182–192. [Google Scholar] [CrossRef] [Green Version]
  63. Song, S.; Jia, Z.; Xu, J.; Zhang, Z.; Bian, Z. N-butyryl-homoserine lactone, a bacterial quorum-sensing signaling molecule, induces intracellular calcium elevation in Arabidopsis root cells. Biochem. Biophys. Res. Commun. 2011, 414, 355–360. [Google Scholar] [CrossRef]
  64. Imran, A.; Saadalla, M.J.A.; Khan, S.-U.; Mirza, M.S.; Malik, K.A.; Hafeez, F.Y. Ochrobactrum sp. Pv2Z2 exhibits multiple traits of plant growth promotion, biodegradation and N-acyl-homoserine-lactone quorum sensing. Ann. Microbiol. 2014, 64, 1797–1806. [Google Scholar] [CrossRef]
  65. Alavi, P.; Müller, H.; Cardinale, M.; Zachow, C.; Sanchez, M.B.; Martinez, J.L.; Berg, G. The DSF Quorum Sensing System Controls the Positive Influence of Stenotrophomonas maltophilia on Plants. PLoS ONE 2013, 8, e67103. [Google Scholar] [CrossRef]
  66. De Sordi, L.; Mühlschlegel, F.A. Quorum sensing and fungal-bacterial interactions inCandida albicans: A communicative network regulating microbial coexistence and virulence. FEMS Yeast Res. 2009, 9, 990–999. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Andersson, D.I.; Hughes, D. Microbiological effects of sublethal levels of antibiotics. Nat. Rev. Microbiol. 2014, 12, 465–478. [Google Scholar] [CrossRef] [PubMed]
  68. Brescia, F.; Marchetti-Deschmann, M.; Musetti, R.; Perazzolli, M.; Pertot, I.; Puopolo, G. The rhizosphere signature on the cell motility, biofilm formation and secondary metabolite production of a plant-associated Lysobacter strain. Microbiol. Res. 2020, 234, 126424. [Google Scholar] [CrossRef] [PubMed]
  69. Hartmann, A.; Schikora, A. Quorum Sensing of Bacteria and Trans-Kingdom Interactions of N-Acyl Homoserine Lactones with Eukaryotes. J. Chem. Ecol. 2012, 38, 704–713. [Google Scholar] [CrossRef] [PubMed]
  70. Clinton, A.; Rumbaugh, K.P. Interspecies and Interkingdom Signaling via Quorum Signals. Isr. J. Chem. 2016, 56, 265–272. [Google Scholar] [CrossRef]
  71. Bitas, V.; Kim, H.-S.; Bennett, J.W.; Kang, S. Sniffing on Microbes: Diverse Roles of Microbial Volatile Organic Compounds in Plant Health. Mol. Plant-Microbe Interact. 2013, 26, 835–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Kanchiswamy, C.N.; Malnoy, M.; Maffei, M. Bioprospecting bacterial and fungal volatiles for sustainable agriculture. Trends Plant Sci. 2015, 20, 206–211. [Google Scholar] [CrossRef] [PubMed]
  73. Kai, M.; Effmert, U.; Piechulla, B. Bacterial-Plant-Interactions: Approaches to Unravel the Biological Function of Bacterial Volatiles in the Rhizosphere. Front. Microbiol. 2016, 7, 108. [Google Scholar] [CrossRef] [Green Version]
  74. Tyc, O.; Zweers, H.; de Boer, W.; Garbeva, P. Volatiles in Inter-Specific Bacterial Interactions. Front. Microbiol. 2015, 6, 1412. [Google Scholar] [CrossRef] [Green Version]
  75. Audrain, B.; Farag, M.A.; Ryu, C.-M.; Ghigo, J.-M. Role of bacterial volatile compounds in bacterial biology. FEMS Microbiol. Rev. 2015, 39, 222–233. [Google Scholar] [CrossRef] [Green Version]
  76. Schmidt, R.; Cordovez, V.; de Boer, W.; Raaijmakers, J.; Garbeva, P. Volatile affairs in microbial interactions. ISME J. 2015, 9, 2329–2335. [Google Scholar] [CrossRef] [Green Version]
  77. Barraud, N.; Schleheck, D.; Klebensberger, J.; Webb, J.S.; Hassett, D.J.; Rice, S.A.; Kjelleberg, S. Nitric Oxide Signaling in Pseudomonas aeruginosa Biofilms Mediates Phosphodiesterase Activity, Decreased Cyclic Di-GMP Levels, and Enhanced Dispersal. J. Bacteriol. 2009, 191, 7333–7342. [Google Scholar] [CrossRef] [Green Version]
  78. Taghavi, S.; Garafola, C.; Monchy, S.; Newman, L.; Hoffman, A.; Weyens, N.; Barac, T.; Vangronsveld, J.; van der Lelie, D. Genome Survey and Characterization of Endophytic Bacteria Exhibiting a Beneficial Effect on Growth and Development of Poplar Trees. Appl. Environ. Microbiol. 2009, 75, 748–757. [Google Scholar] [CrossRef] [Green Version]
  79. Duca, D.; Lorv, J.; Patten, C.L.; Rose, D.; Glick, B.R. Indole-3-acetic acid in plant–microbe interactions. Antonie Van Leeuwenhoek 2014, 106, 85–125. [Google Scholar] [CrossRef]
  80. Fahad, S.; Hussain, S.; Bano, A.; Saud, S.; Hassan, S.; Shan, D.; Khan, F.A.; Khan, F.; Chen, Y.; Wu, C.; et al. Potential role of phytohormones and plant growth-promoting rhizobacteria in abiotic stresses: Consequences for changing environment. Environ. Sci. Pollut. Res. 2015, 22, 4907–4921. [Google Scholar] [CrossRef] [PubMed]
  81. Raheem, A.; Shaposhnikov, A.; Belimov, A.A.; Dodd, I.C.; Ali, B. Auxin production by rhizobacteria was associated with improved yield of wheat (Triticum aestivum L.) under drought stress. Arch. Agron. Soil Sci. 2018, 64, 574–587. [Google Scholar] [CrossRef]
  82. Zhou, X.R.; Dai, L.; Xu, G.F.; Wang, H.S. A strain of Phoma species improves drought tolerance of Pinus tabulaeformis. Sci. Rep. 2021, 11, 7637. [Google Scholar] [CrossRef] [PubMed]
  83. Pieterse, C.M.J.; Zamioudis, C.; Berendsen, R.L.; Weller, D.M.; Van Wees, S.C.M.; Bakker, P.A.H.M. Induced Systemic Resistance by Beneficial Microbes. Annu. Rev. Phytopathol. 2014, 52, 347–375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Vos, I.A.; Pieterse, C.; Van Wees, S. Costs and benefits of hormone-regulated plant defences. Plant Pathol. 2013, 62, 43–55. [Google Scholar] [CrossRef]
  85. Zamioudis, C.; Pieterse, C.M.J. Modulation of Host Immunity by Beneficial Microbes. Mol. Plant-Microbe Interact. 2012, 25, 139–150. [Google Scholar] [CrossRef] [Green Version]
  86. Von Rad, U.; Klein, I.; Dobrev, P.I.; Kottova, J.; Zazimalova, E.; Fekete, A.; Hartmann, A.; Schmitt-Kopplin, P.; Durner, J. Response of Arabidopsis thaliana to N-hexanoyl-dl-homoserine-lactone, a bacterial quorum sensing molecule produced in the rhizosphere. Planta 2008, 229, 73–85. [Google Scholar] [CrossRef] [PubMed]
  87. Kakkar, A.; Nizampatnam, N.R.; Kondreddy, A.; Pradhan, B.B.; Chatterjee, S. Xanthomonas campestriscell–cell signalling molecule DSF (diffusible signal factor) elicits innate immunity in plants and is suppressed by the exopolysaccharide xanthan. J. Exp. Bot. 2015, 66, 6697–6714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Corral-Lugo, A.; Daddaoua, A.; Ortega, A.; Espinosa-Urgel, M.; Krell, T. Rosmarinic acid is a homoserine lactone mimic produced by plants that activates a bacterial quorum-sensing regulator. Sci. Signal. 2016, 9, ra1. [Google Scholar] [CrossRef] [PubMed]
  89. Hartmann, A.; Rothballer, M.; Hense, B.A.; Schröder, P. Bacterial quorum sensing compounds are important modulators of microbe-plant interactions. Front. Plant Sci. 2014, 5, 131. [Google Scholar] [CrossRef] [Green Version]
  90. Jakobsen, T.H.; van Gennip, M.; Phipps, R.K.; Shanmugham, M.S.; Christensen, L.D.; Alhede, M.; Skindersoe, M.E.; Rasmussen, T.B.; Friedrich, K.; Uthe, F.; et al. Ajoene, a Sulfur-Rich Molecule from Garlic, Inhibits Genes Controlled by Quorum Sensing. Antimicrob. Agents Chemother. 2012, 56, 2314–2325. [Google Scholar] [CrossRef] [Green Version]
  91. Ganin, H.; Rayo, J.; Amara, N.; Levy, N.; Krief, P.; Meijler, M.M. Sulforaphane and erucin, natural isothiocyanates from broccoli, inhibit bacterial quorum sensing. MedChemComm 2013, 4, 175–179. [Google Scholar] [CrossRef]
  92. Rudrappa, T.; Bais, H.P. Curcumin, a Known Phenolic from Curcuma longa, Attenuates the Virulence of Pseudomonas aeruginosa PAO1 in Whole Plant and Animal Pathogenicity Models. J. Agric. Food Chem. 2008, 56, 1955–1962. [Google Scholar] [CrossRef]
  93. Rasmussen, T.B.; Bjarnsholt, T.; Skindersoe, M.E.; Hentzer, M.; Kristoffersen, P.; Köte, M.; Nielsen, J.; Eberl, L.; Givskov, M. Screening for Quorum-Sensing Inhibitors (QSI) by Use of a Novel Genetic System, the QSI Selector. J. Bacteriol. 2005, 187, 1799–1814. [Google Scholar] [CrossRef] [Green Version]
  94. Bauer, W.D.; Mathesius, U. Plant responses to bacterial quorum sensing signals. Curr. Opin. Plant Biol. 2004, 7, 429–433. [Google Scholar] [CrossRef]
  95. Gao, M.; Teplitski, M.; Robinson, J.B.; Bauer, W.D. Production of Substances by Medicago truncatula that Affect Bacterial Quorum Sensing. Mol. Plant-Microbe Interact. 2003, 16, 827–834. [Google Scholar] [CrossRef] [Green Version]
  96. Kusari, P.; Kusari, S.; Spiteller, M.; Kayser, O. Implications of endophyte-plant crosstalk in light of quorum responses for plant biotechnology. Appl. Microbiol. Biotechnol. 2015, 99, 5383–5390. [Google Scholar] [CrossRef] [PubMed]
  97. Teplitski, M.; Mathesius, U.; Rumbaugh, K.P. Perception and Degradation of N-Acyl Homoserine Lactone Quorum Sensing Signals by Mammalian and Plant Cells. Chem. Rev. 2011, 111, 100–116. [Google Scholar] [CrossRef] [PubMed]
  98. Zhang, W.; Luo, Q.; Zhang, Y.; Fan, X.; Ye, T.; Mishra, S.; Bhatt, P.; Zhang, L.; Chen, S. Quorum Quenching in a Novel Acinetobacter sp. XN-10 Bacterial Strain against Pectobacterium carotovorum subsp. carotovorum. Microorganisms 2020, 8, 1100. [Google Scholar] [CrossRef] [PubMed]
  99. Weller, D.M.; Mavrodi, D.V.; Van Pelt, J.A.; Pieterse, C.M.J.; Van Loon, L.C.; Bakker, P.A.H.M. Induced Systemic Resistance in Arabidopsis thaliana Against Pseudomonas syringae pv. tomato by 2,4-Diacetylphloroglucinol-Producing Pseudomonas fluorescens. Phytopathology 2012, 102, 403–412. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Groenhagen, U.; Baumgartner, R.; Bailly, A.; Gardiner, A.; Eberl, L.; Schulz, S.; Weisskopf, L. Production of Bioactive Volatiles by Different Burkholderia ambifaria Strains. J. Chem. Ecol. 2013, 39, 892–906. [Google Scholar] [CrossRef] [PubMed]
  101. Kai, M.; Haustein, M.; Molina, F.; Petri, A.; Scholz, B.; Piechulla, B. Bacterial volatiles and their action potential. Appl. Microbiol. Biotechnol. 2009, 81, 1001–1012. [Google Scholar] [CrossRef]
  102. Ryu, C.-M.; Farag, M.A.; Hu, C.-H.; Reddy, M.S.; Wei, H.-X.; Paré, P.W.; Kloepper, J.W. Bacterial volatiles promote growth in Arabidopsis. Proc. Natl. Acad. Sci. USA 2003, 100, 4927–4932. [Google Scholar] [CrossRef] [Green Version]
  103. Bertin, C.; Yang, X.; Weston, L.A. The role of root exudates and allelochemicals in the rhizosphere. Plant Soil 2003, 256, 67–83. [Google Scholar] [CrossRef]
  104. Zhang, Y.; Ruyter-Spira, C.; Bouwmeester, H.J. Engineering the plant rhizosphere. Curr. Opin. Biotechnol. 2015, 32, 136–142. [Google Scholar] [CrossRef]
  105. Downie, J.A. The roles of extracellular proteins, polysaccharides and signals in the interactions of rhizobia with legume roots. FEMS Microbiol. Rev. 2010, 34, 150–170. [Google Scholar] [CrossRef]
  106. Rios, J.; Pascual, J.; Guillen, M.; Lopez-Martinez, A.; Carvajal, M. Influence of foliar Methyl-jasmonate biostimulation on exudation of glucosinolates and their effect on root pathogens of broccoli plants under salinity condition. Sci. Hortic. 2021, 282, 110027. [Google Scholar] [CrossRef]
  107. Zhang, N.; Wang, D.; Liu, Y.; Li, S.; Shen, Q.; Zhang, R. Effects of different plant root exudates and their organic acid components on chemotaxis, biofilm formation and colonization by beneficial rhizosphere-associated bacterial strains. Plant Soil 2014, 374, 689–700. [Google Scholar] [CrossRef]
  108. Pérez-Montaño, F.; Alias-Villegas, C.; Bellogín, R.A.; del Cerro, P.; Espuny, M.R.; Jiménez-Guerrero, I.; López-Baena, F.J.; Ollero, F.; Cubo, T. Plant growth promotion in cereal and leguminous agricultural important plants: From microorganism capacities to crop production. Microbiol. Res. 2014, 169, 325–336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Dudareva, N.; Negre, F.; Nagegowda, D.A.; Orlova, I. Plant Volatiles: Recent Advances and Future Perspectives. Crit. Rev. Plant Sci. 2006, 25, 417–440. [Google Scholar] [CrossRef]
  110. Rasmann, S.; Köllner, T.G.; Degenhardt, J.; Hiltpold, I.; Toepfer, S.; Kuhlmann, U.; Gershenzon, J.; Turlings, T.C.J. Recruitment of entomopathogenic nematodes by insect-damaged maize roots. Nature 2005, 434, 732–737. [Google Scholar] [CrossRef] [PubMed]
  111. Poole, P.; Ramachandran, V.; Terpolilli, J. Rhizobia: From saprophytes to endosymbionts. Nat. Rev. Genet. 2018, 16, 291–303. [Google Scholar] [CrossRef]
  112. Janczarek, M.; Rachwał, K.; Marzec, A.; Grządziel, J.; Palusińska-Szysz, M. Signal molecules and cell-surface components involved in early stages of the legume–rhizobium interactions. Appl. Soil Ecol. 2015, 85, 94–113. [Google Scholar] [CrossRef]
  113. Hassan, S.; Mathesius, U. The role of flavonoids in root-rhizosphere signalling: Opportunities and challenges for improving plant-microbe interactions. J. Exp. Bot. 2012, 63, 3429–3444. [Google Scholar] [CrossRef] [Green Version]
  114. Lerouge, P.; Roche, P.; Faucher, C.; Maillet, F.; Truchet, G.; Promé, J.C.; Dénarié, J. Symbiotic host-specificity of Rhizobium meliloti is determined by a sulphated and acylated glucosamine oligosaccharide signal. Nature 1990, 344, 781–784. [Google Scholar] [CrossRef]
  115. Oldroyd, G.E.; Murray, J.D.; Poole, P.S.; Downie, J.A. The Rules of Engagement in the Legume-Rhizobial Symbiosis. Annu. Rev. Genet. 2011, 45, 119–144. [Google Scholar] [CrossRef]
  116. Limpens, E.; van Zeijl, A.; Geurts, R. Lipochitooligosaccharides Modulate Plant Host Immunity to Enable Endosymbioses. Annu. Rev. Phytopathol. 2015, 53, 311–334. [Google Scholar] [CrossRef]
  117. Smith, D.L.; Gravel, V.; Yergeau, E. Editorial: Signaling in the Phytomicrobiome. Front. Plant Sci. 2017, 8, 611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Smith, D.L.; Praslickova, D.; Ilangumaran, G. Inter-organismal signaling and management of the phytomicrobiome. Front. Plant Sci. 2015, 6, 722. [Google Scholar] [CrossRef] [Green Version]
  119. Oláh, B.; Brière, C.; Bécard, G.; Dénarié, J.; Gough, C. Nod factors and a diffusible factor from arbuscular mycorrhizal fungi stimulate lateral root formation in Medicago truncatula via the DMI1/DMI2 signalling pathway. Plant J. 2005, 44, 195–207. [Google Scholar] [CrossRef]
  120. Khan, A.L.; Hamayun, M.; Kim, Y.-H.; Kang, S.-M.; Lee, J.-H.; Lee, I.-J. Gibberellins producing endophytic Aspergillus fumigatus sp. LH02 influenced endogenous phytohormonal levels, isoflavonoids production and plant growth in salinity stress. Process Biochem. 2011, 46, 440–447. [Google Scholar] [CrossRef]
  121. Cooper, J. Early interactions between legumes and rhizobia: Disclosing complexity in a molecular dialogue. J. Appl. Microbiol. 2007, 103, 1355–1365. [Google Scholar] [CrossRef] [PubMed]
  122. Mabood, F.; Souleimanov, A.; Khan, W.; Smith, D. Jasmonates induce Nod factor production by Bradyrhizobium japonicum. Plant Physiol. Biochem. 2006, 44, 759–765. [Google Scholar] [CrossRef] [PubMed]
  123. Benjelloun, I.; Alami, I.T.; El Khadir, M.; Douira, A.; Udupa, S. Co-Inoculation of Mesorhizobium ciceri with Either Bacillus sp. or Enterobacter aerogenes on Chickpea Improves Growth and Productivity in Phosphate-Deficient Soils in Dry Areas of a Mediterranean Region. Plants 2021, 10, 571. [Google Scholar] [CrossRef]
  124. Schmitz, A.M.; Harrison, M.J. Signaling events during initiation of arbuscular mycorrhizal symbiosis. J. Integr. Plant Biol. 2014, 56, 250–261. [Google Scholar] [CrossRef]
  125. Ruyter-Spira, C.; Al-Babili, S.; van der Krol, S.; Bouwmeester, H. The biology of strigolactones. Trends Plant Sci. 2013, 18, 72–83. [Google Scholar] [CrossRef]
  126. Waldie, T.; McCulloch, H.; Leyser, O. Strigolactones and the control of plant development: Lessons from shoot branching. Plant J. 2014, 79, 607–622. [Google Scholar] [CrossRef] [PubMed]
  127. Kaschuk, G.; Leffelaar, P.A.; Giller, K.E.; Alberton, O.; Hungria, M.; Kuyper, T.W. Responses of legumes to rhizobia and arbuscular mycorrhizal fungi: A meta-analysis of potential photosynthate limitation of symbioses. Soil Biol. Biochem. 2010, 42, 125–127. [Google Scholar] [CrossRef]
  128. Phour, M.; Sehrawat, A.; Sindhu, S.S.; Glick, B.R. Interkingdom signaling in plant-rhizomicrobiome interactions for sustainable agriculture. Microbiol. Res. 2020, 241, 126589. [Google Scholar] [CrossRef] [PubMed]
  129. Vangelisti, A.; Natali, L.; Bernardi, R.; Sbrana, C.; Turrini, A.; Hassani-Pak, K.; Hughes, D.; Cavallini, A.; Giovannetti, M.; Giordani, T. Transcriptome changes induced by arbuscular mycorrhizal fungi in sunflower (Helianthus annuus L.) roots. Sci. Rep. 2018, 8, 4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  130. Maillet, F.; Poinsot, V.; André, O.; Puech-Pagès, V.; Haouy, A.; Gueunier, M.; Cromer, L.; Giraudet, D.; Formey, D.; Niebel, A.; et al. Fungal lipochitooligosaccharide symbiotic signals in arbuscular mycorrhiza. Nature 2011, 469, 58–63. [Google Scholar] [CrossRef]
  131. MacLean, A.M.; Bravo, A.; Harrison, M.J. Plant Signaling and Metabolic Pathways Enabling Arbuscular Mycorrhizal Symbiosis. Plant Cell 2017, 29, 2319–2335. [Google Scholar] [CrossRef]
  132. Ben Laouane, R.; Meddich, A.; Bechtaoui, N.; Oufdou, K.; Wahbi, S. Effects of Arbuscular Mycorrhizal Fungi and Rhizobia Symbiosis on the Tolerance of Medicago Sativa to Salt Stress. Gesunde Pflanz. 2019, 71, 135–146. [Google Scholar] [CrossRef]
  133. Hao, Z.; Xie, W.; Jiang, X.; Wu, Z.; Zhang, X.; Chen, B. Arbuscular Mycorrhizal Fungus Improves Rhizobium–Glycyrrhiza Seedling Symbiosis under Drought Stress. Agronomy 2019, 9, 572. [Google Scholar] [CrossRef] [Green Version]
  134. Tang, T.-T.; Xie, M.-M.; Chen, S.-M.; Zhang, S.-M.; Wu, Q.-S. Effects of Arbuscular Mycorrhizal Fungi and Rhizobia on Physiological Activities in White Clover (Trifolium repens). Biotechnology 2019, 18, 49–54. [Google Scholar] [CrossRef]
  135. Sui, X.-L.; Zhang, T.; Tian, Y.-Q.; Xue, R.-J.; Li, A.-R. A neglected alliance in battles against parasitic plants: Arbuscular mycorrhizal and rhizobial symbioses alleviate damage to a legume host by root hemiparasitic Pedicularis species. New Phytol. 2019, 221, 470–481. [Google Scholar] [CrossRef] [Green Version]
  136. Sakamoto, K.; Ogiwara, N.; Kaji, T.; Sugimoto, Y.; Ueno, M.; Sonoda, M.; Matsui, A.; Ishida, J.; Tanaka, M.; Totoki, Y.; et al. Transcriptome analysis of soybean (Glycine max) root genes differentially expressed in rhizobial, arbuscular mycorrhizal, and dual symbiosis. J. Plant Res. 2019, 132, 541–568. [Google Scholar] [CrossRef] [PubMed]
  137. Selvaraj, A.; Thangavel, K.; Uthandi, S. Arbuscular mycorrhizal fungi (Glomus intraradices) and diazotrophic bacterium (Rhizobium BMBS) primed defense in blackgram against herbivorous insect (Spodoptera litura) infestation. Microbiol. Res. 2020, 231, 126355. [Google Scholar] [CrossRef] [PubMed]
  138. Fabiańska, I.; Gerlach, N.; Almario, J.; Bucher, M. Plant-mediated effects of soil phosphorus on the root-associated fungal microbiota in Arabidopsis thaliana. New Phytol. 2019, 221, 2123–2137. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  139. Verbon, E.H.; Trapet, P.L.; Stringlis, I.; Kruijs, S.; Bakker, P.A.; Pieterse, C.M. Iron and Immunity. Annu. Rev. Phytopathol. 2017, 55, 355–375. [Google Scholar] [CrossRef] [PubMed]
  140. Martínez-Medina, A.; Van Wees, S.C.; Pieterse, C.M. Airborne signals from Trichoderma fungi stimulate iron uptake responses in roots resulting in priming of jasmonic acid-dependent defences in shoots of Arabidopsis thaliana and Solanum lycopersicum. Plant Cell Environ. 2017, 40, 2691–2705. [Google Scholar] [CrossRef] [Green Version]
  141. Zamioudis, C.; Korteland, J.; Van Pelt, J.A.; Hamersveld, M.; Dombrowski, N.; Bai, Y.; Hanson, J.; Van Verk, M.C.; Ling, H.; Schulze-Lefert, P.; et al. Rhizobacterial volatiles and photosynthesis-related signals coordinate MYB 72 expression in Arabidopsis roots during onset of induced systemic resistance and iron-deficiency responses. Plant J. 2015, 84, 309–322. [Google Scholar] [CrossRef] [Green Version]
  142. Hiruma, K.; Gerlach, N.; Sacristan, S.; Nakano, R.T.; Hacquard, S.; Kracher, B.; Neumann, U.; Ramirez, 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] [Green Version]
  143. 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]
  144. Martin, F.M.; Uroz, S.; Barker, D.G. Ancestral alliances: Plant mutualistic symbioses with fungi and bacteria. Science 2017, 356, eaad4501. [Google Scholar] [CrossRef]
  145. Singh, S.; Nain, L. Microorganisms in the Conversion of Agricultural Wastes to Compost. Proc. Indian Natl. Sci. Acad. 2014, 80, 473. [Google Scholar] [CrossRef]
  146. Ahmed, S.; Rahman, S.; Hasan, M.; Paul, N.; Sajib, A.A. Microbial degradation of lignocellulosic biomass: Discovery of novel natural lignocellulolytic bacteria. BioTechnologia 2018, 99, 137–146. [Google Scholar] [CrossRef]
  147. Haggag, W.M.; Abouziena, H.F.; Abd-El-Kreem, F.; El Habbasha, S. Agriculture biotechnology for management of multiple biotic and abiotic environmental stress in crops. J. Chem. Pharm. Res. 2015, 7, 882–889. [Google Scholar]
  148. Balmer, A.; Pastor, V.; Gamir, J.; Flors, V.; Mauch-Mani, B. The ‘prime-ome’: Towards a holistic approach to priming. Trends Plant Sci. 2015, 20, 443–452. [Google Scholar] [CrossRef] [PubMed]
  149. Selosse, M.-A.; Bessis, A.; Pozo, M.J. Microbial priming of plant and animal immunity: Symbionts as developmental signals. Trends Microbiol. 2014, 22, 607–613. [Google Scholar] [CrossRef] [PubMed]
  150. Edenancé, N.; Esánchez-Vallet, A.; Egoffner, D.; Emolina, A. Disease resistance or growth: The role of plant hormones in balancing immune responses and fitness costs. Front. Plant Sci. 2013, 4, 155. [Google Scholar] [CrossRef] [Green Version]
  151. Pandey, S.S.; Singh, S.; Babu, C.S.V.; Shanker, K.; Srivastava, N.K.; Shukla, A.K.; Kalra, A. Fungal endophytes of Catharanthus roseus enhance vindoline content by modulating structural and regulatory genes related to terpenoid indole alkaloid biosynthesis. Sci. Rep. 2016, 6, 26583. [Google Scholar] [CrossRef] [Green Version]
  152. Tena, G.; Boudsocq, M.; Sheen, J. Protein kinase signaling networks in plant innate immunity. Curr. Opin. Plant Biol. 2011, 14, 519–529. [Google Scholar] [CrossRef] [Green Version]
  153. Shah, J.; Zeier, J. Long-distance communication and signal amplification in systemic acquired resistance. Front. Plant Sci. 2013, 4, 30. [Google Scholar] [CrossRef] [Green Version]
  154. Pieterse, C.M.J.; Van der Does, D.; Zamioudis, C.; Leon-Reyes, A.; Van Wees, S.C.M. Hormonal Modulation of Plant Immunity. Annu. Rev. Cell Dev. Biol. 2012, 28, 489–521. [Google Scholar] [CrossRef] [Green Version]
  155. Gimenez-Ibanez, S.; Chini, A.; Solano, R. How Microbes Twist Jasmonate Signaling around Their Little Fingers. Plants 2016, 5, 9. [Google Scholar] [CrossRef] [Green Version]
  156. 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] [PubMed]
  157. Couto, D.; Zipfel, C. Regulation of pattern recognition receptor signalling in plants. Nat. Rev. Immunol. 2016, 16, 537–552. [Google Scholar] [CrossRef]
  158. Li, B.; Meng, X.; Shan, L.; He, P. Transcriptional Regulation of Pattern-Triggered Immunity in Plants. Cell Host Microbe 2016, 19, 641–650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Clay, N.K.; Adio, A.M.; Denoux, C.; Jander, G.; Ausubel, F.M. Glucosinolate Metabolites Required for an Arabidopsis Innate Immune Response. Science 2009, 323, 95–101. [Google Scholar] [CrossRef] [Green Version]
  160. Derbyshire, M.; Mbengue, M.; Barascud, M.; Navaud, O.; Raffaele, S. Small RNAs from the plant pathogenic fungus Sclerotinia sclerotiorum highlight host candidate genes associated with quantitative disease resistance. Mol. Plant Pathol. 2019, 20, 1279–1297. [Google Scholar] [CrossRef] [Green Version]
  161. Baetz, U.; Martinoia, E. Root exudates: The hidden part of plant defense. Trends Plant Sci. 2014, 19, 90–98. [Google Scholar] [CrossRef] [Green Version]
  162. Jeandet, P.; Hébrard, C.; Deville, M.-A.; Cordelier, S.; Dorey, S.; Aziz, A.; Crouzet, J. Deciphering the Role of Phytoalexins in Plant-Microorganism Interactions and Human Health. Molecules 2014, 19, 18033–18056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Rin, S.; Mizuno, Y.; Shibata, Y.; Fushimi, M.; Katou, S.; Sato, I.; Chiba, S.; Kawakita, K.; Takemoto, D. EIN2-Mediated signaling is involved in pre-invasion defense in Nicotiana benthamiana against potato late blight pathogen, Phytophthora infestans. Plant Signal. Behav. 2017, 12, e1300733. [Google Scholar] [CrossRef] [Green Version]
  164. Pel, M.J.C.; Pieterse, C.M.J. Microbial recognition and evasion of host immunity. J. Exp. Bot. 2013, 64, 1237–1248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Jones, J.D.G.; Dangl, J.L. The plant immune system. Nature 2006, 444, 323–329. [Google Scholar] [CrossRef] [Green Version]
  166. Boller, T.; He, S.Y. Innate Immunity in Plants: An Arms Race between Pattern Recognition Receptors in Plants and Effectors in Microbial Pathogens. Science 2009, 324, 742–744. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Hasegawa, M.; Mitsuhara, I.; Seo, S.; Imai, T.; Koga, J.; Okada, K.; Yamane, H.; Ohashi, Y. Phytoalexin Accumulation in the Interaction Between Rice and the Blast Fungus. Mol. Plant-Microbe Interact. 2010, 23, 1000–1011. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  168. King, S.R.; McLellan, H.; Boevink, P.C.; Armstrong, M.R.; Bukharova, T.; Sukarta, O.; Win, J.; Kamoun, S.; Birch, P.R.; Banfield, M.J. Phytophthora infestans RXLR Effector PexRD2 Interacts with Host MAPKKKε to Suppress Plant Immune Signaling. Plant Cell 2014, 26, 1345–1359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Boevink, P.C.; Wang, X.; McLellan, H.; He, Q.; Naqvi, S.; Armstrong, M.R.; Zhang, W.; Hein, I.; Gilroy, E.M.; Tian, Z.; et al. A Phytophthora infestans RXLR effector targets plant PP1c isoforms that promote late blight disease. Nat. Commun. 2016, 7, 10311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Martinez-Medina, A.; Flors, V.; Heil, M.; Mauch-Mani, B.; Pieterse, C.; Pozo, M.J.; Ton, J.; van Dam, N.M.; Conrath, U. Recognizing Plant Defense Priming. Trends Plant Sci. 2016, 21, 818–822. [Google Scholar] [CrossRef] [Green Version]
  171. Bruce, T.J.; Matthes, M.C.; Napier, J.A.; Pickett, J.A. Stressful “memories” of plants: Evidence and possible mechanisms. Plant Sci. 2007, 173, 603–608. [Google Scholar] [CrossRef]
  172. Dempsey, D.A.; Klessig, D.F. SOS—Too many signals for systemic acquired resistance? Trends Plant Sci. 2012, 17, 538–545. [Google Scholar] [CrossRef]
  173. Ádám, A.L.; Nagy, Z.; Kátay, G.; Mergenthaler, E.; Viczián, O. Signals of Systemic Immunity in Plants: Progress and Open Questions. Int. J. Mol. Sci. 2018, 19, 1146. [Google Scholar] [CrossRef] [Green Version]
  174. Beckers, G.J.M.; Spoel, S.H. Fine-Tuning Plant Defence Signalling: Salicylate versus Jasmonate. Plant Biol. 2006, 8, 1–10. [Google Scholar] [CrossRef]
  175. Wu, Y.; Zhang, D.; Chu, J.Y.; Boyle, P.; Wang, Y.; Brindle, I.D.; De Luca, V.; Després, C. The Arabidopsis NPR1 protein is a receptor for the plant defense hormone salicylic acid. Cell Rep. 2012, 1, 639–647. [Google Scholar] [CrossRef] [Green Version]
  176. Ullah, C.; Tsai, C.; Unsicker, S.B.; Xue, L.; Reichelt, M.; Gershenzon, J.; Hammerbacher, A. Salicylic acid activates poplar defense against the biotrophic rust fungus Melampsora larici-populina via increased biosynthesis of catechin and proanthocyanidins. New Phytol. 2019, 221, 960–975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Yoodee, S.; Kobayashi, Y.; Songnuan, W.; Boonchird, C.; Thitamadee, S.; Kobayashi, I.; Narangajavana, J. Phytohormone priming elevates the accumulation of defense-related gene transcripts and enhances bacterial blight disease resistance in cassava. Plant Physiol. Biochem. 2018, 122, 65–77. [Google Scholar] [CrossRef] [PubMed]
  178. Vitti, A.; Pellegrini, E.; Nali, C.; Lovelli, S.; Sofo, A.; Valerio, M.; Scopa, A.; Nuzzaci, M. Trichoderma harzianum T-22 Induces Systemic Resistance in Tomato Infected by Cucumber mosaic virus. Front. Plant Sci. 2016, 7, 1520. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Purwantisari, S.; Priyatmojo, A.; Sancayaningsih, R.P.; Kasiamdari, R.S.; Budihardjo, K. Systemic inducing resistance against late blight by applying antagonist Trichoderma Viride. J. Phys. Conf. Ser. 2018, 1025, 012053. [Google Scholar] [CrossRef] [Green Version]
  180. Doornbos, R.F.; Geraats, B.P.J.; Kuramae, E.E.; Van Loon, L.C.; Bakker, P.A.H.M. Effects of Jasmonic Acid, Ethylene, and Salicylic Acid Signaling on the Rhizosphere Bacterial Community of Arabidopsis thaliana. Mol. Plant-Microbe Interact. 2011, 24, 395–407. [Google Scholar] [CrossRef] [Green Version]
  181. Farag, M.A.; Zhang, H.; Ryu, C.-M. Dynamic Chemical Communication between Plants and Bacteria through Airborne Signals: Induced Resistance by Bacterial Volatiles. J. Chem. Ecol. 2013, 39, 1007–1018. [Google Scholar] [CrossRef] [Green Version]
  182. Yang, J.; Kloepper, J.W.; Ryu, C.-M. Rhizosphere bacteria help plants tolerate abiotic stress. Trends Plant Sci. 2009, 14, 1–4. [Google Scholar] [CrossRef]
  183. Bittel, P.; Robatzek, S. Microbe-associated molecular patterns (MAMPs) probe plant immunity. Curr. Opin. Plant Biol. 2007, 10, 335–341. [Google Scholar] [CrossRef] [Green Version]
  184. Lorenzo, O.; Solano, R. Molecular players regulating the jasmonate signalling network. Curr. Opin. Plant Biol. 2005, 8, 532–540. [Google Scholar] [CrossRef]
  185. Von Dahl, C.C.; Baldwin, I.T. Deciphering the Role of Ethylene in Plant–Herbivore Interactions. J. Plant Growth Regul. 2007, 26, 201–209. [Google Scholar] [CrossRef]
  186. Choudhary, D.K.; Johri, B.N. Interactions of Bacillus spp. and plants—With special reference to induced systemic resistance (ISR). Microbiol. Res. 2009, 164, 493–513. [Google Scholar] [CrossRef] [PubMed]
  187. Sun, Y.; Xiao, J.; Jia, X.; Ke, P.; He, L.; Cao, A.; Wang, H.; Wu, Y.; Gao, X.; Wang, X. The role of wheat jasmonic acid and ethylene pathways in response to Fusarium graminearum infection. Plant Growth Regul. 2016, 80, 69–77. [Google Scholar] [CrossRef]
  188. Singh, V.; Singh, D.; Gautam, J.K.; Nandi, A.K. RSI1/FLD is a positive regulator for defense against necrotrophic pathogens. Physiol. Mol. Plant Pathol. 2019, 107, 40–45. [Google Scholar] [CrossRef]
  189. Kilic-Ekici, O.; Yuen, G.Y. Comparison of strains of Lysobacter enzymogenes and PGPR for induction of resistance against Bipolaris sorokiniana in tall fescue. Biol. Control 2004, 30, 446–455. [Google Scholar] [CrossRef]
  190. Wang, C.-J.; Yang, W.; Wang, C.; Gu, C.; Niu, D.-D.; Liu, H.-X.; Wang, Y.-P.; Guo, J.-H. Induction of Drought Tolerance in Cucumber Plants by a Consortium of Three Plant Growth-Promoting Rhizobacterium Strains. PLoS ONE 2012, 7, e52565. [Google Scholar] [CrossRef] [Green Version]
  191. Kohler, J.; Hernández, J.A.; Caravaca, F.; Roldán, A. Plant-growth-promoting rhizobacteria and arbuscular mycorrhizal fungi modify alleviation biochemical mechanisms in water-stressed plants. Funct. Plant Biol. 2008, 35, 141–151. [Google Scholar] [CrossRef]
  192. Gupta, G.; Parihar, S.S.; Ahirwar, N.K.; Snehi, S.K.; Singh, V. Plant growth promoting rhizobacteria (PGPR): Current and future prospects for development of sustainable agriculture. J. Microb. Biochem. Technol. 2015, 7, 096–102. [Google Scholar]
  193. Beneduzi, A.; Ambrosini, A.; Passaglia, L.M. Plant growth-promoting rhizobacteria (PGPR): Their potential as antagonists and biocontrol agents. Genet. Mol. Biol. 2012, 35, 1044–1051. [Google Scholar] [CrossRef] [Green Version]
  194. Chowdhury, S.P.; Uhl, J.; Grosch, R.; Alquéres, S.; Pittroff, S.; Dietel, K.; Schmitt-Kopplin, P.; Borriss, R.; Hartmann, A. Cyclic Lipopeptides of Bacillus amyloliquefaciens subsp. plantarum Colonizing the Lettuce Rhizosphere Enhance Plant Defense Responses Toward the Bottom Rot Pathogen Rhizoctonia solani. Mol. Plant-Microbe Interact. 2015, 28, 984–995. [Google Scholar] [CrossRef] [Green Version]
  195. Han, S.; Li, D.; Trost, E.; Mayer, K.F.; Vlot, A.C.; Heller, W.; Schmid, M.; Hartmann, A.; Rothballer, M. Systemic Responses of Barley to the 3-hydroxy-decanoyl-homoserine Lactone Producing Plant Beneficial Endophyte Acidovorax radicis N35. Front. Plant Sci. 2016, 7, 1868. [Google Scholar] [CrossRef] [Green Version]
  196. Patil, S.; Bheemaraddi, M.C.; Shivannavar, C.T.; Gaddad, S.M. Biocontrol activity of siderophore producing Bacillus subtilis CTS-G24 against wilt and dry root rot causing fungi in chickpea. IOSR J. Agric. Vet. Sci. 2014, 7, 63–68. [Google Scholar] [CrossRef]
  197. Dunne, C.; Crowley, J.J.; Moënne-Loccoz, Y.; Dowling, D.N.; Bruijn, S.; O’Gara, F. Biological control of Pythium ultimum by Stenotrophomonas maltophilia W81 is mediated by an extracellular proteolytic activity. Microbiology 1997, 143, 3921–3931. [Google Scholar] [CrossRef] [Green Version]
  198. Huang, X.; Liu, J.; Ding, J.; He, Q.; Xiong, R.; Zhang, K. The investigation of nematocidal activity in Stenotrophomonas maltophilia G2 and characterization of a novel virulence serine protease. Can. J. Microbiol. 2009, 55, 934–942. [Google Scholar] [CrossRef] [PubMed]
  199. Wang, X.; Wei, Z.; Yang, K.; Wang, J.; Jousset, A.; Xu, Y.; Shen, Q.; Friman, V.-P. Phage combination therapies for bacterial wilt disease in tomato. Nat. Biotechnol. 2019, 37, 1513–1520. [Google Scholar] [CrossRef] [PubMed]
  200. Shrivastava, P.; Kumar, R. Soil salinity: A serious environmental issue and plant growth promoting bacteria as one of the tools for its alleviation. Saudi J. Biol. Sci. 2015, 22, 123–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  201. Enebe, M.C.; Babalola, O.O. The impact of microbes in the orchestration of plants’ resistance to biotic stress: A disease management approach. Appl. Microbiol. Biotechnol. 2018, 103, 9–25. [Google Scholar] [CrossRef] [Green Version]
  202. Kumar, A.; Verma, J.P. Does plant—Microbe interaction confer stress tolerance in plants: A review? Microbiol. Res. 2018, 207, 41–52. [Google Scholar] [CrossRef]
  203. Meena, K.K.; Sorty, A.M.; Bitla, U.M.; Choudhary, K.; Gupta, P.; Pareek, A.; Singh, D.P.; Prabha, R.; Sahu, P.K.; Gupta, V.K.; et al. Abiotic Stress Responses and Microbe-Mediated Mitigation in Plants: The Omics Strategies. Front. Plant Sci. 2017, 8, 172. [Google Scholar] [CrossRef]
  204. Sorty, A.M.; Meena, K.K.; Choudhary, K.; Bitla, U.M.; Minhas, P.S.; Krishnani, K.K. Effect of Plant Growth Promoting Bacteria Associated with Halophytic Weed (Psoralea corylifolia L.) on Germination and Seedling Growth of Wheat Under Saline Conditions. Appl. Biochem. Biotechnol. 2016, 180, 872–882. [Google Scholar] [CrossRef]
  205. Gopalakrishnan, S.; Sathya, A.; Vijayabharathi, R.; Varshney, R.K.; Gowda, C.L.; Krishnamurthy, L. Plant growth promoting rhizobia: Challenges and opportunities. 3 Biotech 2015, 5, 355–377. [Google Scholar] [CrossRef] [Green Version]
  206. Bhargava, S.; Sawant, K. Drought stress adaptation: Metabolic adjustment and regulation of gene expression. Plant Breed. 2013, 132, 21–32. [Google Scholar] [CrossRef]
  207. Chaves, M.M.; Flexas, J.; Pinheiro, C. Photosynthesis under drought and salt stress: Regulation mechanisms from whole plant to cell. Ann. Bot. 2009, 103, 551–560. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Todaka, D.; Nakashima, K.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Toward understanding transcriptional regulatory networks in abiotic stress responses and tolerance in rice. Rice 2012, 5, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Qin, F.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Achievements and Challenges in Understanding Plant Abiotic Stress Responses and Tolerance. Plant Cell Physiol. 2011, 52, 1569–1582. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  210. Tiwari, S.; Singh, P.; Tiwari, R.; Meena, K.K.; Yandigeri, M.; Singh, D.P.; Arora, D.K. Salt-tolerant rhizobacteria-mediated induced tolerance in wheat (Triticum aestivum) and chemical diversity in rhizosphere enhance plant growth. Biol. Fertil. Soils 2011, 47, 907–916. [Google Scholar] [CrossRef]
  211. Atkinson, N.J.; Lilley, C.J.; Urwin, P.E. Identification of Genes Involved in the Response of Arabidopsis to Simultaneous Biotic and Abiotic Stresses. Plant Physiol. 2013, 162, 2028–2041. [Google Scholar] [CrossRef] [Green Version]
  212. Prasch, C.M.; Sonnewald, U. Simultaneous Application of Heat, Drought, and Virus to Arabidopsis Plants Reveals Significant Shifts in Signaling Networks. Plant Physiol. 2013, 162, 1849–1866. [Google Scholar] [CrossRef]
  213. Nishiyama, R.; Watanabe, Y.; Leyva-Gonzalez, M.A.; Van Ha, C.; Fujita, Y.; Tanaka, M.; Seki, M.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; Herrera-Estrella, L.; et al. Arabidopsis AHP2, AHP3, and AHP5 histidine phosphotransfer proteins function as redundant negative regulators of drought stress response. Proc. Natl. Acad. Sci. USA 2013, 110, 4840–4845. [Google Scholar] [CrossRef] [Green Version]
  214. Peleg, Z.; Blumwald, E. Hormone balance and abiotic stress tolerance in crop plants. Curr. Opin. Plant Biol. 2011, 14, 290–295. [Google Scholar] [CrossRef]
  215. Xiong, L.; Lee, H.; Ishitani, M.; Zhu, J.-K. Regulation of Osmotic Stress-responsive Gene Expression by theLOS6/ABA1 Locus inArabidopsis. J. Biol. Chem. 2002, 277, 8588–8596. [Google Scholar] [CrossRef] [Green Version]
  216. Gazzarrini, S. Cross-talk in Plant Hormone Signalling: What Arabidopsis Mutants Are Telling Us. Ann. Bot. 2003, 91, 605–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Halo, B.A.; Khan, A.L.; Waqas, M.; Al-Harrasi, A.; Hussain, J.; Ali, L.; Adnan, M.; Lee, I.-J. Endophytic bacteria (Sphingomonas sp. LK11) and gibberellin can improve Solanum lycopersicum growth and oxidative stress under salinity. J. Plant Interact. 2015, 10, 117–125. [Google Scholar] [CrossRef] [Green Version]
  218. Barnawal, D.; Bharti, N.; Pandey, S.S.; Pandey, A.; Chanotiya, C.S.; Kalra, A. Plant growth-promoting rhizobacteria enhance wheat salt and drought stress tolerance by altering endogenous phytohormone levels and TaCTR1/TaDREB2 expression. Physiol. Plant. 2017, 161, 502–514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. Kang, S.-M.; Shahzad, R.; Bilal, S.; Khan, A.L.; Park, Y.-G.; Lee, K.-E.; Asaf, S.; Khan, M.A.; Lee, I.-J. Indole-3-acetic-acid and ACC deaminase producing Leclercia adecarboxylata MO1 improves Solanum lycopersicum L. growth and salinity stress tolerance by endogenous secondary metabolites regulation. BMC Microbiol. 2019, 19, 80. [Google Scholar] [CrossRef] [Green Version]
  220. Bharti, N.; Pandey, S.S.; Barnawal, D.; Patel, V.K.; Kalra, A. Plant growth promoting rhizobacteria Dietzia natronolimnaea modulates the expression of stress responsive genes providing protection of wheat from salinity stress. Sci. Rep. 2016, 6, 34768. [Google Scholar] [CrossRef] [Green Version]
  221. Kang, S.-M.; Khan, A.L.; Waqas, M.; You, Y.-H.; Kim, J.-H.; Kim, J.-G.; Hamayun, M.; Lee, I.-J. Plant growth-promoting rhizobacteria reduce adverse effects of salinity and osmotic stress by regulating phytohormones and antioxidants in Cucumis sativus. J. Plant Interact. 2014, 9, 673–682. [Google Scholar] [CrossRef] [Green Version]
  222. Zhao, L.; Zhang, Y.-Q. Effects of phosphate solubilization and phytohormone production of Trichoderma asperellum Q1 on promoting cucumber growth under salt stress. J. Integr. Agric. 2015, 14, 1588–1597. [Google Scholar] [CrossRef] [Green Version]
  223. He, A.; Niu, S.; Yang, D.; Ren, W.; Zhao, L.; Sun, Y.; Meng, L.; Zhao, Q.; Paré, P.W.; Zhang, J. Two PGPR strains from the rhizosphere of Haloxylon ammodendron promoted growth and enhanced drought tolerance of ryegrass. Plant Physiol. Biochem. 2021, 161, 74–85. [Google Scholar] [CrossRef]
  224. Jochum, M.; McWilliams, K.; Borrego, E.J.; Kolomiets, M.V.; Niu, G.; Pierson, E.A.; Jo, Y.-K. Bioprospecting Plant Growth-Promoting Rhizobacteria That Mitigate Drought Stress in Grasses. Front. Microbiol. 2019, 10, 2106. [Google Scholar] [CrossRef]
  225. Saakre, M.; Baburao, T.M.; Salim, A.P.; Ffancies, R.M.; Achuthan, V.P.; Thomas, G.; Sivarajan, S.R. Identification and Characterization of Genes Responsible for Drought Tolerance in Rice Mediated by Pseudomonas fluorescens. Rice Sci. 2017, 24, 291–298. [Google Scholar] [CrossRef]
  226. Begum, N.; Wang, L.; Ahmad, H.; Akhtar, K.; Roy, R.; Khan, M.I.; Zhao, T. Co-inoculation of Arbuscular Mycorrhizal Fungi and the Plant Growth-Promoting Rhizobacteria Improve Growth and Photosynthesis in Tobacco Under Drought Stress by Up-Regulating Antioxidant and Mineral Nutrition Metabolism. Microb. Ecol. 2021, 83, 971–988. [Google Scholar] [CrossRef] [PubMed]
  227. Chatterjee, P.; Kanagendran, A.; Samaddar, S.; Pazouki, L.; Sa, T.-M.; Niinemets, Ü. Influence of Brevibacterium linens RS16 on foliage photosynthetic and volatile emission characteristics upon heat stress in Eucalyptus grandis. Sci. Total Environ. 2020, 700, 134453. [Google Scholar] [CrossRef] [PubMed]
  228. Mukhtar, T.; Rehman, S.U.; Smith, D.; Sultan, T.; Seleiman, M.F.; Alsadon, A.A.; Amna; Ali, S.; Chaudhary, H.J.; Solieman, T.H.I.; et al. Mitigation of Heat Stress in Solanum lycopersicum L. by ACC-deaminase and Exopolysaccharide Producing Bacillus cereus: Effects on Biochemical Profiling. Sustainability 2020, 12, 2159. [Google Scholar] [CrossRef] [Green Version]
  229. Sukkasem, P.; Kurniawan, A.; Kao, T.-C.; Chuang, H.-W. A multifaceted rhizobacterium Bacillus licheniformis functions as a fungal antagonist and a promoter of plant growth and abiotic stress tolerance. Environ. Exp. Bot. 2018, 155, 541–551. [Google Scholar] [CrossRef]
  230. Sarkar, J.; Chakraborty, B.; Chakraborty, U. Plant Growth Promoting Rhizobacteria Protect Wheat Plants Against Temperature Stress Through Antioxidant Signalling and Reducing Chloroplast and Membrane Injury. J. Plant Growth Regul. 2018, 37, 1396–1412. [Google Scholar] [CrossRef]
  231. El-Daim, I.A.A.; Bejai, S.; Meijer, J. Bacillus velezensis 5113 Induced Metabolic and Molecular Reprogramming during Abiotic Stress Tolerance in Wheat. Sci. Rep. 2019, 9, 16282. [Google Scholar] [CrossRef] [Green Version]
  232. Tiwari, S.; Prasad, V.; Chauhan, P.S.; Lata, C. Bacillus amyloliquefaciens Confers Tolerance to Various Abiotic Stresses and Modulates Plant Response to Phytohormones through Osmoprotection and Gene Expression Regulation in Rice. Front. Plant Sci. 2017, 8, 1510. [Google Scholar] [CrossRef] [Green Version]
  233. Caradonia, F.; Francia, E.; Morcia, C.; Ghizzoni, R.; Moulin, L.; Terzi, V.; Ronga, D. Arbuscular Mycorrhizal Fungi and Plant Growth Promoting Rhizobacteria Avoid Processing Tomato Leaf Damage during Chilling Stress. Agronomy 2019, 9, 299. [Google Scholar] [CrossRef] [Green Version]
  234. Su, F.; Jacquard, C.; Villaume, S.; Michel, J.; Rabenoelina, F.; Clément, C.; Barka, E.A.; Dhondt-Cordelier, S.; Vaillant-Gaveau, N. Burkholderia phytofirmans PsJN reduces impact of freezing temperatures on photosynthesis in Arabidopsis thaliana. Front. Plant Sci. 2015, 6, 810. [Google Scholar] [CrossRef] [Green Version]
  235. Pramanik, K.; Mitra, S.; Sarkar, A.; Maiti, T.K. Alleviation of phytotoxic effects of cadmium on rice seedlings by cadmium resistant PGPR strain Enterobacter aerogenes MCC 3092. J. Hazard. Mater. 2018, 351, 317–329. [Google Scholar] [CrossRef]
  236. Zafar-Ul-Hye, M.; Naeem, M.; Danish, S.; Khan, M.J.; Fahad, S.; Datta, R.; Brtnicky, M.; Kintl, A.; Hussain, G.S.; El-Esawi, M.A. Effect of Cadmium-Tolerant Rhizobacteria on Growth Attributes and Chlorophyll Contents of Bitter Gourd under Cadmium Toxicity. Plants 2020, 9, 1386. [Google Scholar] [CrossRef] [PubMed]
  237. Bilal, S.; Shahzad, R.; Imran, M.; Jan, R.; Kim, K.M.; Lee, I.-J. Synergistic association of endophytic fungi enhances Glycine max L. resilience to combined abiotic stresses: Heavy metals, high temperature and drought stress. Ind. Crop. Prod. 2020, 143, 111931. [Google Scholar] [CrossRef]
  238. Mesa-Marín, J.; Del-Saz, N.F.; Rodríguez-Llorente, I.D.; Redondo-Gómez, S.; Pajuelo, E.; Ribas-Carbó, M.; Mateos-Naranjo, E. PGPR Reduce Root Respiration and Oxidative Stress Enhancing Spartina maritima Root Growth and Heavy Metal Rhizoaccumulation. Front. Plant Sci. 2018, 9, 1500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  239. Gururani, M.A.; Upadhyaya, C.P.; Baskar, V.; Venkatesh, J.; Nookaraju, A.; Park, S.W. Plant Growth-Promoting Rhizobacteria Enhance Abiotic Stress Tolerance in Solanum tuberosum Through Inducing Changes in the Expression of ROS-Scavenging Enzymes and Improved Photosynthetic Performance. J. Plant Growth Regul. 2013, 32, 245–258. [Google Scholar] [CrossRef]
  240. Armendariz, A.L.; Talano, M.A.; Nicotra, M.F.O.; Escudero, L.; Breser, M.L.; Porporatto, C.; Agostini, E. Impact of double inoculation with Bradyrhizobium japonicum E109 and Azospirillum brasilense Az39 on soybean plants grown under arsenic stress. Plant Physiol. Biochem. 2019, 138, 26–35. [Google Scholar] [CrossRef]
  241. Singh, R.P.; Usha; Rizvi, S.M.H.; Sonia; Jaiwal, P.K. Genetic Engineering for Enhancing Abiotic Stress Tolerance; Springer: Dordrecht, The Netherlands, 2003; pp. 223–243. [Google Scholar] [CrossRef]
  242. Hrynkiewicz, K.; Zloch, M.; Kowalkowski, T.; Baum, C.; Niedojadło, K.; Buszewski, B. Strain-Specific bioaccumulation and intracellular distribution of Cd2+ in bacteria isolated from the rhizosphere, ectomycorrhizae, and fruitbodies of ectomycorrhizal fungi. Environ. Sci. Pollut. Res. 2015, 22, 3055–3067. [Google Scholar] [CrossRef] [Green Version]
  243. Chauhan, P.S.; Lata, C.; Tiwari, S.; Chauhan, A.S.; Mishra, S.K.; Agrawal, L.; Chakrabarty, D.; Nautiyal, C.S. Transcriptional alterations reveal Bacillus amyloliquefaciens -rice cooperation under salt stress. Sci. Rep. 2019, 9, 11912. [Google Scholar] [CrossRef]
  244. Nadeem, S.M.; Zahir, Z.A.; Naveed, M.; Nawaz, S. Mitigation of salinity-induced negative impact on the growth and yield of wheat by plant growth-promoting rhizobacteria in naturally saline conditions. Ann. Microbiol. 2013, 63, 225–232. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of interactions between rhizo-microorganisms and plants.
Figure 1. Schematic representation of interactions between rhizo-microorganisms and plants.
Microorganisms 10 00899 g001
Figure 2. Schematic representation of quorum sensing (QS) and quorum quenching (QQ) inhibition of signal perception pathways. The binding of QQ agents to the LuxR receptors either inactivates quorum sensing receptors or reduces the quantity of receptors in the QS molecules for targeted gene expression.
Figure 2. Schematic representation of quorum sensing (QS) and quorum quenching (QQ) inhibition of signal perception pathways. The binding of QQ agents to the LuxR receptors either inactivates quorum sensing receptors or reduces the quantity of receptors in the QS molecules for targeted gene expression.
Microorganisms 10 00899 g002
Figure 3. Overview of rhizosphere communication in intra- or interspecies signaling among microorganisms and interkingdom signaling between microbes and plants. Myc, mycorrhizal; LCOs, lipo-chitooligosaccharides; Nod, nodulation; VOCs, volatile organic compounds; Ais, autoinducers; AHLs, N-acyl homoserine lactone; QSMs, quorum sensing molecules.
Figure 3. Overview of rhizosphere communication in intra- or interspecies signaling among microorganisms and interkingdom signaling between microbes and plants. Myc, mycorrhizal; LCOs, lipo-chitooligosaccharides; Nod, nodulation; VOCs, volatile organic compounds; Ais, autoinducers; AHLs, N-acyl homoserine lactone; QSMs, quorum sensing molecules.
Microorganisms 10 00899 g003
Figure 4. Summary of significant factors of rhizosphere signaling. SAR, systemic acquired resistance; ISR, induced systemic resistance; LCOs, lipo-chitooligosaccharides; Nod, nodulation; Myc, mycorrhizal; PRR, pattern recognition receptor; PAMPs, pathogen-associated molecular patterns.
Figure 4. Summary of significant factors of rhizosphere signaling. SAR, systemic acquired resistance; ISR, induced systemic resistance; LCOs, lipo-chitooligosaccharides; Nod, nodulation; Myc, mycorrhizal; PRR, pattern recognition receptor; PAMPs, pathogen-associated molecular patterns.
Microorganisms 10 00899 g004
Figure 5. Schematic representation of plant immunity against phytopathogens through pattern-triggered immunity (PTI) and effector-triggered immunity (ETI). PTI is activated through the recognition of pathogen-associated molecular patterns (PAMPs) by the pattern recognition receptors (PRRs) in plant cell walls. After PAMP recognition, the plant triggers a signal cascade that further induces multiple intracellular defense responses and activates defense genes. In response to PTI, pathogens secrete effectors that are recognized by nucleotide-binding (NB) and leucine-rich repeat (LRR) receptors (NLRs) and develop another defense layer called ETI. Activated NLRs with many intracellular signaling events trigger the hypersensitive response (HR).
Figure 5. Schematic representation of plant immunity against phytopathogens through pattern-triggered immunity (PTI) and effector-triggered immunity (ETI). PTI is activated through the recognition of pathogen-associated molecular patterns (PAMPs) by the pattern recognition receptors (PRRs) in plant cell walls. After PAMP recognition, the plant triggers a signal cascade that further induces multiple intracellular defense responses and activates defense genes. In response to PTI, pathogens secrete effectors that are recognized by nucleotide-binding (NB) and leucine-rich repeat (LRR) receptors (NLRs) and develop another defense layer called ETI. Activated NLRs with many intracellular signaling events trigger the hypersensitive response (HR).
Microorganisms 10 00899 g005
Figure 6. Schematic representation of defense priming: (a) a picture depicting induced systemic resistance (ISR), which is induced by plant growth-promoting rhizobacteria that provide resistance to biotic stresses, and systemic acquired resistance (SAR), which is induced by phytopathogens that provide resistance to biotic and abiotic stresses; (b) a flowchart showing signal transduction pathway of ISR that are followed by jasmonate (JA) and ethylene (ET) plant hormones, which are encoded by the jar1 and etr1 genes, respectively, whereas SAR is followed by salicylic acid (SA), which is encoded by the sid1 and sid2 genes.
Figure 6. Schematic representation of defense priming: (a) a picture depicting induced systemic resistance (ISR), which is induced by plant growth-promoting rhizobacteria that provide resistance to biotic stresses, and systemic acquired resistance (SAR), which is induced by phytopathogens that provide resistance to biotic and abiotic stresses; (b) a flowchart showing signal transduction pathway of ISR that are followed by jasmonate (JA) and ethylene (ET) plant hormones, which are encoded by the jar1 and etr1 genes, respectively, whereas SAR is followed by salicylic acid (SA), which is encoded by the sid1 and sid2 genes.
Microorganisms 10 00899 g006
Table 1. Characterization of quorum sensing signaling molecules that are produced by different rhizomicrobes with respect to the type of communication within the rhizosphere.
Table 1. Characterization of quorum sensing signaling molecules that are produced by different rhizomicrobes with respect to the type of communication within the rhizosphere.
Type of MicrobeRhizo-MicroorganismsQuorum Sensing MoleculesType of CommunicationReference
Gram-positive bacteriaBacillus subtilisComX pheromoneInter- and intraspecies[6]
Streptomyces spp.Gamma-butyrolactones (A-factor) and methylenomycin furans (MMF1)Interspecies[55]
Staphylococcus aureusCircular oligopeptideInterspecies[16]
Stenotrophomonas chelatiphagaDSF (diffuse signal factor)Interkingdom (poplar plant)[56]
Bacillus velezensisAI-2 synthetase (2-methyl-2,3,3,4-tetrahydroxytetrahydrofuran (THMF))Interkingdom (maize plant)[57]
Gram-negative bacteriaBurkholderia sp.N-3-oxo-hexanoyl-homoserine lactone (3OC6-HSL)Interkingdom (wheat and arabidopsis plant)[58]
Serratia glossinaeN-hexanoyl-L-homoserine lactone (m/z 200) and N-octanoyl-L-homoserine lactone (m/z 228)Interkingdom
(rice plant)
[59]
Serratia plymuthicaN-butanoyl-HSL, N-hexanoyl-HSL, and N-3-oxo-hexanoyl-HSL (OHHL)Interkingdom (oil seed rape)[60]
Burkholderia graminis M12 and B. graminis M14N-(3-oxododecanoyl)-L-homoserine lactone (3-oxo-C12-HSL or OC12-HSL (where “O” indicates an oxo substitution at the third carbon atom)) and 3-oxo-C14-HSL (OC14-HSL)Interkingdom (tomato plant)[61]
Serratia marcescensN-3-oxo-hexanoyl-homoserine lactone (3OC6-HSL)Intraspecies and interkingdom (tobacco plant)[62]
Burkholderia sp. and Pseudomonas sp.N-butyryl-homoserine lactone (C4-HSL)Interkingdom (arabidopsis plant)[63]
Ochrobactrum sp.3O-C7-HSL and 3OH-C7-HSLInterkingdom (bean plant)[64]
Stenotrophomonas maltophiliaDSF (diffuse signal factor)Interkingdom (oil seed rape)[65]
Serratia glossinaeN-octanoyl-L-homoserine lactone and N-hexanoyl-L-homoserine lactoneInterkingdom (sesame plant)[59]
FungiCandida albicansTyrosol, γ-butyrolactone, and farnesolIntraspecies[66]
Table 2. Amelioration of abiotic stresses by rhizo-microorganisms using different signaling pathways in different plant species.
Table 2. Amelioration of abiotic stresses by rhizo-microorganisms using different signaling pathways in different plant species.
Stress TypeHost PlantRhizo-MicroorganismsSignaling PathwaysReference
SalinityTriticum aestivumArthrobacter protophormiae (SA3) and Dietzia natronolimnaea (STR1)IAA, ET[218]
Lycopersicum esculentumLeclercia adecarboxylataIAA[219]
Triticum aestivumDietzia natronolimnaeaSOS, ABA[220]
Cucumis sativusBurkholdera cepacia, Promicromonospora sp., and Acinetobacter calcoaceticusGA, ABA, SA[221]
Cucumis sativusTrichoderma asperellumIAA, ABA, GA[222]
DroughtLolium multiflorumBacillus sp. and Pseudomonas sp.ABA[223]
Triticum aestivum, Zea maysBacillus sp. and Enterobacter sp.IAA, SA[224]
Oryza sativaPseudomonas fluorescensABA[225]
Nicotiana tabacumGlomus versiforme (AMF) and Bacillus methylotrophicusABA, IAA[226]
HeatEucalyptus grandisBrevibacterium linensET[227]
Lycopersicum esculentumBacillus cereusET[228]
ArabidopsisBacillus licheniformisJA, ABA[229]
Triticum aestivumBacillus safensis and Ochrobactrum pseudogrignonenseROS[230]
Low temperatureTriticum aestivumBacillus velezensisROS, ABA[231]
Oryza sativaBacillus amyloliquefaciensABA, SA, JA, ET[232]
Solanum lycopersicumFunneliformis mosseae and Paraburkholderia graminisROS[233]
Arabidopsis thalianaBurkholderia phytofirmansROS[234]
Heavy metalsCdOryza sativaEnterobacter aerogenesIAA, ET[235]
Momordica charantiaStenotrophomonas maltophilia and Agrobacterium fabrumET[236]
Ni, Cd, and AlGlycine maxPaecilomyces formosus and Penicillium funiculosumIAA, GA[237]
Cd, Cu, Pb, and NiSpartima maritimaBacillus methylotrophicus, Bacillus licheniformis, and Bacillus aryabhattaiCOX, AOX[238]
Solanum tuberosumBacillus sp.ROS[239]
AsGlycine maxBradyrhizobium japonicum and Azospirillum brasilenseIAA[240]
IAA, indole acetic acid; ET, ethylene; GA, gibbberellin; JA, jasmonic acid; ROS, reactive oxygen species; ABA, abscisic acid; SOS, salt overly sensitive; COX, cytochrome oxidase; AOX, alternative oxidase.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jamil, F.; Mukhtar, H.; Fouillaud, M.; Dufossé, L. Rhizosphere Signaling: Insights into Plant–Rhizomicrobiome Interactions for Sustainable Agronomy. Microorganisms 2022, 10, 899. https://doi.org/10.3390/microorganisms10050899

AMA Style

Jamil F, Mukhtar H, Fouillaud M, Dufossé L. Rhizosphere Signaling: Insights into Plant–Rhizomicrobiome Interactions for Sustainable Agronomy. Microorganisms. 2022; 10(5):899. https://doi.org/10.3390/microorganisms10050899

Chicago/Turabian Style

Jamil, Fatima, Hamid Mukhtar, Mireille Fouillaud, and Laurent Dufossé. 2022. "Rhizosphere Signaling: Insights into Plant–Rhizomicrobiome Interactions for Sustainable Agronomy" Microorganisms 10, no. 5: 899. https://doi.org/10.3390/microorganisms10050899

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