Next Article in Journal
Acknowledgment to the Reviewers of Stresses in 2022
Next Article in Special Issue
Elicitation of Fruit Fungi Infection and Its Protective Response to Improve the Postharvest Quality of Fruits
Previous Article in Journal
Fetal Programming of Brain and Behavior through Ionizing Radiation
Previous Article in Special Issue
Maize Breeding for Low Nitrogen Inputs in Agriculture: Mechanisms Underlying the Tolerance to the Abiotic Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Beneficial Microorganisms as a Sustainable Alternative for Mitigating Biotic Stresses in Crops

by
Ana María García-Montelongo
1,
Amelia C. Montoya-Martínez
1,
Pamela Helue Morales-Sandoval
1,2,
Fannie Isela Parra-Cota
3 and
Sergio de los Santos-Villalobos
1,*
1
Laboratorio de Biotecnología del Recurso Microbiano, Departamento de Ciencias Agronómicas y Veterinarias, Instituto Tecnológico de Sonora (ITSON), 5 de Febrero 818 Sur, Col. Centro, Obregón 85000, Mexico
2
Facultad de Biología, Universidad Veracruzana, Circuito Gonzalo Aguirre Beltrán s/n, Xalapa, Veracruz 91090, Mexico
3
Campo Experimental Norman E. Borlaug, Instituto Nacional De Investigaciones Forestales, Agrícolas y Pecuarias, Norman E. Borlaug s/n, Col. Centro, Obregón 85000, Mexico
*
Author to whom correspondence should be addressed.
Stresses 2023, 3(1), 210-228; https://doi.org/10.3390/stresses3010016
Submission received: 23 November 2022 / Revised: 4 January 2023 / Accepted: 11 January 2023 / Published: 15 January 2023
(This article belongs to the Special Issue Physiological and Molecular Mechanisms of Plant Stress Tolerance)

Abstract

:
Nowadays, population growth, the global temperature increase, and the appearance of emerging diseases in important crops generate uncertainty regarding world food security. The use of agrochemicals has been the “go-to” solution for the control of phytopathogenic microorganisms, such as Magnaporte oryzae, causing blast disease in rice and other cereals; Botrytis cinerea, causing gray mold in over 500 plant species; and Puccinia spp., causing rust in cereals. However, their excessive use has harmed human health, as well as ecosystems (contaminating water, and contributing to soil degradation); besides, phytopathogens can develop resistance to them. The inoculation of plant growth-promoting microorganisms (PGPMs) to crops is a sustainable strategy for increasing the yield and quality of crops and mitigating biotic stresses. Likewise, PGPMs, such as Pseudomonas, Bacillus, and Trichoderma, can trigger a series of signals and reactions in the plant that lead to the induction of systemic resistance, a mechanism by which plants react to microorganism stimulation by activating their defense system, resulting in protection against future pathogen attack. These plant defense mechanisms help to mitigate biotic stresses that threaten global food security. Thus, the study of these mechanisms at molecular, transcriptomic, and metabolomic levels is indispensable to elucidate how stresses affect globally important crops.

1. Introduction

As the world population is estimated to grow to more than 9 billion people by 2050, agriculture and the world’s crop production face significant challenges to increase production and yields. In addition, a shortage of arable land and the impact of climate change are also threats to food safety [1]. Plants experience natural stresses such as temperature change, water deficit, mineral deficiency, pests, and anthropogenic activities (an intensive application of herbicides, pesticides, fungicides, and fertilizers), which cause a loss in crop yield [2].
In plants, stress can be defined as any unfavorable condition or substance that affects or blocks a plant’s metabolism, growth, or development [3] and generally refers to environmental changes [4]. Thus, stress can be divided into two categories: abiotic and biotic. Abiotic stress refers to environmental factors such as physical or chemical, while biotic stress exposes plants to biological units such as diseases and predators [4].
Biotic stress in plants can be caused by one or multiple pathogens. Moreover, a type of phytopathogen can cause disease in different crops; due to this, plants have developed mechanisms of identification, control, and resistance to these diseases [5,6]. The reactions to biotic and abiotic stress negatively affect the crops since it limits their growth, development, yield, and quality [7]. These can hamper the quality and proportion of agricultural products, threatening food security [8].
There are predictions of the significant effects on agricultural production caused by the global temperature increase by the end of the century [9,10]. Climate change can cause the appearance of new diseases and modify the severity of existing ones [11]. Pests and diseases are among the major causes of lost crop production, estimated as between 20% and 40% of losses to yield, with significant consequences for livelihoods, public health, and the environment [12]. The annual yield losses caused by crop pests and diseases are estimated at USD 220 billion [12]. Globally, three of the most important plant diseases caused by phytopathogenic fungi are rice blast disease (caused by Magnaporte oryzae), gray mold (Botrytis cinerea), and rust (Puccinia spp.) [13,14]. The first one is the most destructive disease of rice worldwide; infections caused by M. oryzae lead to the annual destruction of approximately 10–30% of the rice harvested globally. Some analysts estimate that it destroys harvests that could feed 60 million people each year [15]. Gray mold is a disease caused by B. cinerea, having a host range of more than 500 plant species; it is especially damaging to fresh fruits and vegetables, making it the most important post-harvest disease, with annual economic losses easily exceeding USD 10 billion worldwide [16]. Puccinia spp. cause major crop losses, threatening food security and the sustainability of crop production in 31 countries across the world [17]. In wheat, three major rust pathogens (stripe rust (P. graminis f. sp. striiformis), stem rust (P. graminis f. sp. tritici), and leaf rust (P. triticina)) impact wheat production, causing losses of USD 4.3 to 5 billion annually, with resulting yield losses of 6–7 million metric tons per year [18]. However, the use of agrochemicals to maximize crop yields and control pests and diseases has had adverse repercussions on the environment and human health. Thus, sustainable agriculture through beneficial microorganisms is an effective way to satisfy the food demand and maintain functioning and healthy ecosystems [19]. The main objectives of this review are to summarize current knowledge on how different biotic stresses affect globally important crops and threaten world food security, and to focus on how plants combat these stresses through their defense systems.

2. Plants Biotic Stress: Interactions between Plants and Pathogens

Plants have a diversity of microorganisms inhabiting them, both below and aboveground, including endophytes (Pseudomonas, Bacillus, Burkholderia) [20], epiphytes (Pseudomonas, Erwinia, Pantoea) [21], phyllosphere (Methylobacterium, Microbacterium) [22], and/or rhizosphere (Rhizobium, Agrobacterium, Pseudomonas, Stenotrophomonas, Bacillus) [23,24]. These microorganisms can establish close and complex beneficial or pathogenic relationships with plants. The functions of microorganisms in plants are diverse and range from interactions that can be synergistic, antagonistic, and mutualistic, and they are modulated by multiple and complex biotic and abiotic factors [25]. In this sense, some microorganisms cause plant diseases (phytopathogens), generating production and quality losses, making them a considerable threat and some of the biggest obstacles to global food security, such as species of the genera Botrytis, Colletotrichum, Fusarium, Magnaporthe, Mycosphaerella, Phytophthora, Puccinia, Sclerotinia, and Xanthomonas, among many others [13,26,27]. However, several microorganisms have shown a beneficial effect on vegetal growth, development, and overall agricultural production (known as plant growth promoter microorganisms, PGPMs), increasing the interest in integrating them as alternatives to chemical products in sustainable agriculture [25]. Thus, knowing these plant–microorganism interactions, and the factors involved in them, can lead to a better understanding of how plants deal with biotic stresses and how to improve their management [28]. For example, in 2021, Nifakos et al. [29] used a genome mining approach and chemical analysis to elucidate the main mechanism of biocontrol of Botrytis cinerea by Bacillus velezensis Bvel1 in grape berries. They found that preventive treatment with Bvel1 cell culture before B. cinerea inoculation strongly suppressed fungal growth and significantly reduced the incidence of gray mold on red globe grapes. Their analysis showed that Bvel1 is capable of producing and secreting a mixture of bioactive diffusible secondary metabolites (iturin A2, surfactin A-C13 and -C15 isoforms, oxydifficidin, L-dihydroanticapsin, bacillibactin) and specialized metabolites (azelaic acid) that are known to exert a strong antifungal activity and/or trigger host plant defenses against pathogens [29].
Plants are healthy when they carry out their physiological functions to the best of their genetic potential [5]. Microbial pathogens (such as fungi, bacteria, viruses, and/or nematodes) cause plant diseases and biotic stress that limit this potential and cause considerable losses in agriculture, horticulture, and forestry, which is known as a disease [5,30].
Plant disease results from the interaction between the plant, the pathogen, and the environment. This concept is known as the disease triangle and states that for a disease to occur, a susceptible plant host, a virulent pathogen, and the proper environmental conditions are required, as a lack of any of these three factors results in the disease failing to develop (Figure 1) [27].
Phytopathogens generate the natural resistance of plants, which combines the effects of barriers and inducible mechanisms, based on physical (waxy cuticle, active closure of stomatal pores) and biochemical (ROS accumulation, hypersensitive response, PAMP-triggered immunity) defenses [31]. For a pathogenic microorganism to infect a plant, there must be (i) a union to the surface to begin with the degradation of physical and chemical barriers of the host, (ii) a production of toxins, and (iii) an inactivation of the plant defenses [32]. After the pathogen achieves infection successfully, a disease is developed (Table 1) and biotic stress is present.

3. Plants’ Response to Biotic Stress: Systemic Acquired Resistance (SAR)

The first line of defense of plants against pathogens and biotic stress is their physical barriers, such as the waxy cuticle and active closure of stomatal pores, but depending on whether the pathogen is a virus, bacteria, or filamentous microorganism, the pathogen possesses mechanisms to overcome those barriers [44]. Pathogens that can overcome these barriers access the plant’s apoplast, where the pathogen is perceived and recognized through transmembrane proteins called pattern recognition receptors (PRRs), which recognize specific, conserved microbial features named pathogen-associated molecular patterns (PAMPs) [44,45]. Successful pathogens deploy a diversity of effectors that suppress PAMP-triggered immunity (PTI) through susceptibility (S) proteins (effector targets), allowing plant cell infection and resulting in effector-triggered susceptibility (ETS). For pathogen detection, plants have evolved mechanisms that specifically recognize and respond to pathogens’ effectors. This interaction was observed between pathogens carrying single dominant genes (avirulence genes) that caused them to be recognized by plant hosts carrying single dominant resistance (R) genes, leading to the gene-for-gene interaction [46,47]. Thus, when pathogens are recognized and fail to cause disease, they are called avirulent pathogens, the host is called resistant, and the interaction is called incompatible [48]. On the other side, in the absence of gene-for-gene recognition, the pathogen is virulent, the host is susceptible, and the interaction is compatible, resulting in pathogen proliferation within the plant cells and the onset of disease and biotic stress in the plant host [47,48].
PTI is the first active plant response to pathogen recognition. PTI consists of diverse cellular responses, including reactive oxygen and calcium bursts, mitogen-activated protein kinase (MAPK) signaling, plant hormone responses, transcriptional reprogramming, and cell wall fortification [47,49]. One of the earliest physiological responses in PTI is an influx of extracellular Ca2+ in the cytosol (a calcium burst). It induces the opening of other membrane transporters (influx of H+, efflux of K+, Cl, and NO3−), which results in an extracellular alkalinization and depolarization of the plasma membrane [37]. These changes are detected by Ca2+ sensors, such as calmodulin (CaM), calmodulin-like protein (CML), or calcium-dependent protein kinase (CDPK), and it signals the hypersensitive response (HR), a programmed cell death at the site of infection, as well as rapid production of reactive oxygen species (ROS) [50].
The generation and accumulation of ROS include hydrogen peroxide (H2O2), superoxide anion, and hydroxyl radicals [51]. H2O2 is an active signaling molecule and its accumulation results in a variety of cellular responses as hypersensitive cell death [52], and blocks the cell cycle progression [53]. It also functions as a developmental signal for the onset of secondary wall differentiation and is an activator of mitogen-activated protein kinases (MAPKs) [51].
A local HR is often associated with the onset of systemic acquired resistance (SAR) in distal plant tissues, as signals derived from cells undergoing the HR contribute significantly to the induction of defense gene transcription in adjacent cells [54,55]. These include mechanisms such as salicylic acid (SA) accumulation, pathogenesis-related (PR) gene expression, and long-lasting resistance to pathogen infection [56] (Figure 2).
HR cell death leads to the activation of SA signaling throughout the whole plant. After the accumulation of SA, a part is converted to methyl salicylate (MeSA), which works as a phloem-mobile SAR signal [56]. MeSa can quickly spread through systemic tissue but does not induce defense gene expression alone. Once in systemic tissue, MeSa is modified to SA by the methyl esterase SABP2 [57,58]. SA levels rise, activating various presumed defense effector genes, including PR genes [35]. SAR signaling downstream of SA is controlled by the protein nonexpressor of PR gene 1 (NPR1), which, upon SA-induced redox changes in the cell, converts from its oligomeric form to monomers; then, it is transported from the cytosol into the nucleus, where it acts as a transcriptional coactivator for the TGA transcription factors to induce the expression of a large set of PR genes [59,60]. PR proteins accumulate intercellularly in vacuoles and have antimicrobial activity; for example, PR2 protein is a glucanase, PR3 and PR8 are chitinases, PR9 is a peroxidase, and PR1 has been shown to inhibit germination of oomycetes and have antifungal activity [50,59,61]. Disease resistance is given by diverse PR proteins that act together so that the overexpression or silencing of a single one does not significantly affect the resistance or susceptibility of the plant to a range of pathogens [58].
As mentioned before, H2O2, produced during the HR, is an activator of MAPKs, which are conserved protein kinases that establish signaling modules where MAPK kinase kinases (MAPKKKs) activate MAPK kinases (MAPKKs) which in turn activate MAPKs. In plants, upon detecting biotic or abiotic stress, MAPKs participate in the signal transduction to the nucleus, allowing adequate transcriptional reprogramming to occur during defense [49,62]. In plant immunity and defense, transcriptional reprogramming is a highly dynamic and controlled process, considered the main link between signal transduction and the implementation of induced defense mechanisms [49,63].
The other two types of plant defense compounds are phytoalexins and phytoanticipins. The first ones are produced by the plant host as a direct response to pathogen detection, whereas phytoanticipins are produced before an attack and are only converted to their toxic forms post-pathogen perception, as a constitutive chemical barrier against the microbial attack [45,49,64]. All of these mechanisms, signaling, and metabolic pathways act together to stop pathogen infection and mitigate biotic stresses in the host plant; thus, systemic acquired resistance is a mechanism of defense that confers long-lasting protection against a broad spectrum of microorganisms [58].
There are research studies that focused on elucidating the mechanisms of SAR activation in different pathogen–plant host systems. One of them was carried out by Iakimova et al. [65], where they analyzed the morphological features of the HR and the contribution of signaling events resulting from ethylene production, ROS generation, and gene expression of a caspase-1-like protease (VPE) to Erwinia amylovora-induced cell death in spot-inoculated apple leaves. They observed that HR phenotypically resembled both the resistant and susceptible genotypes, expressing a similar pattern of distinguishable micro HR lesions that progressed into confined macro HR lesions, and the morphology of dead cells (protoplast shrinkage and retraction from cell wall) in apple leaves resembled necrotic programmed cell death (PCD). ROS accumulation and elevation of ethylene levels were similar in both cultivars, but in the resistant apple cultivar, an early and late increase in VPE gene expression was detected, suggesting that VPE might be an underlying component of the response to E. amylovora in resistant apple cultivars and that E. amylovora triggers leaf resistance response that is manifested through cell death [65].
Similarly, Moya-Elizondo and Jacobsen [65] demonstrated that Bacillus mycoides reduced the severity of Fusarium crown rot of wheat by 10% compared to water control, by inducing high concentrations of peroxidase and endochitinase, which makes it difficult for Fusarium to penetrate. Thus, Bacillus mycoides can induce SAR in wheat plants [66].

4. Interactions between Plants and Plant Growth-Promoting Microorganisms for Mitigating Biotic Stresses

As mentioned previously, PGPM has a beneficial effect on plant growth and development. Nevertheless, besides growth promotion through diverse mechanisms, PGPM can also protect plants from diseases and control phytopathogens [67], and therefore can alleviate biotic stresses in plants for sustainable agriculture.
PGPM has a diversity of mechanisms to defend and protect plants from pathogens, whether directly or indirectly. Some PGPMs, such as Bacillus, Pseudomonas, and Trichoderma, interact with plants by inducing resistance or priming without directly interacting with the targeted pathogen [67,68,69]. Others act directly by interfering with the pathogen via nutrient and niche competition (as Trichoderma spp. and Pseudomonas spp.), parasitism (Trichoderma spp.), and antibiosis (Bacillus spp., Pseudomonas spp., and Streptomyces spp.) [67,70,71]. The result of these interactions is varied and hardly predictable, and the response depends, to a greater extent, on the interactions of biochemical and physiological compatibility [72]. Soil microorganisms have the particularity of proliferating and developing rapidly, taking advantage of the nutritional variety of the soil. This type of microorganism interacts with the rhizosphere, forming a niche highly dependent on the exudates of the plant roots, such as Azospirillum, Bacillus, Trichoderma, and Rhizobium [69,73,74,75]. As PGPMs and plants interact, they establish specialized relationships whereby gene expression results in the synthesis of metabolites (i.e., flavonoids, strigolactone, malic acid, pectin, citrate, and sugars), which play an essential role in effective communication [76,77].
Plant–microorganism communication can be studied through signalomics, which describes the metabolomics approaches employed to decipher the chemical communications occurring within the rhizosphere [71]. Rhizosphere microorganisms, such as Azospirillum, Bacillus, Trichoderma, and Rhizobium [69,74,75], establish symbiotic interactions with plants, secreting or emitting molecules beneficial to the plant (Table 2). These molecules can be phytohormones, volatile organic compounds (VOCs), and quorum-sensing signals such as N-acyl homoserine lactones (AHLs) [78,79,80].

5. Induced Systemic Resistance (ISR)

A practical method of biological control is the extensive use of non-pathogenic rhizosphere-associated microbial species, known as PGPMs [94,95]. This type of biological control is used as an alternative to synthetic pesticides. These microorganisms have a fundamental role in the ecology of the agroecosystem, promoting plant growth [96] through different mechanisms, such as metabolite production, atmospheric nitrogen fixation, and/or making nutrients assimilable by the plant [97,98]. PGPMs (such as Trichoderma [99] and Bacillus [100]) are free-living and colonize the rhizosphere of plants, forming part of the plant microbiome [96], which consists of microorganisms present in leaves, rhizosphere, pollen, and other parts of the plant. The microbiome is reactive to biotic and abiotic plant interaction conditions, helping the plant to absorb nutrients and tolerate several types of stress [101,102].
There is evidence that root microorganisms can migrate to other parts of the plant, through the xylem apoplast, colonizing stems, and finally through leaves, establishing large populations [103]. These microorganisms play an essential role in antibiotic production, competition for nutrients, parasitism, and the production of suppressive metabolites (hydrogen cyanide, siderophores), as well as in the induction of systemic resistance or defense mechanisms in the plant against pathogen attack (Figure 3) [73].
The plant-developed defense system related to the interaction with microorganisms is complex and is mainly activated by molecular patterns (MAMPs) based on surface recognition on the cell membrane [104,105]. MAMPs are protein fragments essential for microorganisms [104], such as flagellins in bacteria or β-glucans in fungi, that are absent in plants. MAMPs control the plant’s colonization by beneficial microorganisms and the signaling pathways for salicylic acid (SA), jasmonic acid (JA), and ethylene (ET), which play a crucial role in controlling the functions of these microorganisms inside the plant [102,106,107].
MAMPs in plants are recognized by pattern recognition receptors (PRR), which are located on the plant surface [108,109], and when in direct contact, start the signaling cascade that leads to the activation of resistance mechanisms. Another type of MAMP recognition pattern is leucine-rich repeats (LRRs), which are part of the PRRs found intracellularly, rich in leucine as their name indicates; however, when the immune response is activated by LRR, it is named MAMP-triggered immunity (MTI). Many of these receptors LRRs contain nucleotide binding sites (NBS-LRRs); these receptors function by detecting pathogen virulence proteins (effectors), and in turn, this immunity is called effector-triggered immunity (ETI) [108,110]. This type of resistance can recognize cellular modifications in plants and proteins produced by pathogens and terminate pathogen infection [111]. Likewise, ETI is gene-for-gene-resistant and prevents infection by biotrophic pathogens by generating programmed cell death [60,86,112].
MAMP recognition in leaves triggers oxidative response, leading to ethylene and nitric oxide synthesis and the activation of mitogen-activated protein kinase (MAPK) signaling cascades, leading to defense-related gene transcription, as well as segregation of callose and β-glucans at the site of plant infection as a physical barrier, preventing plant infection (Figure 3) [113,114].
Generally, the rhizosphere is constantly exposed to microbial interaction, which generates induced systemic resistance [113,115]. This mechanism refers to the action of generalized or systemic resistance or defense capacity, as well as non-generalized local resistance to pathogens or diseases [116]. When we refer to ISR in plants, we refer to a generalized or systemic resistance action, whereas when we use the term induced resistance (IR), we refer to local resistance that does not generalize to pathogens or diseases [116]. Both types of resistance develop by contact with non-pathogenic microorganisms, unlike systemic acquired resistance, in which a series of protective mechanisms are generated in the plant due to direct contact with one or several pathogens This type of resistance is activated in plants by rhizosphere contact with plant growth-promoting fungi or bacteria and is mediated by ET and JA (Figure 3) [117,118].
On the other hand, this resistance can also be activated by priming caused by chemicals, environmental or abiotic factors, as well as other biological agents. Thus, once plants are in contact with the resistance elicitor or activator, a series of biochemical and molecular changes are triggered, but the plant’s defense mechanisms are not directly activated until the plant is exposed to the biotic stress [119]. The importance of this priming for plant defense lies in their adjustment to unpredictable environments, which has been studied in a wide variety of crops, ranging from herbaceous to woody plants [119,120]. This mechanism can be local and occur only in one part of the plant, or systematically in the whole plant [116,121].
In this sense, in plants, the NPR1 gene has been linked to both SAR and ISR, since its transcriptional role is necessary for ET synthesis and the cytosolic transcription of this gene in JA/ET signaling. NPR1, NPR3, and NPR4 genes are related to PGPM establishment in roots, leading to ISR activation [60,94,122]. On the other hand, the role of plant lipoxygenases (LOX) is associated with resistance in plants, playing crucial roles in biotic and abiotic stress [123] and defending plants from pathogen attack by synthesizing oxylipins, compounds that oxidize pathogen fatty acids [105]. The pathogen defense mechanism of these is based on the enzyme LOX, which is involved in the synthesis of JA and ET, along with lipid deoxygenation [121]. Oxylipin synthesis is also associated with green leafy vegetables in the form of volatile compounds, which release these compounds to activate the defense system against pathogens due to volatile compounds, which can induce crop tolerance to biotic stress [124,125].
The enhancement of the expression of genes that regulate the JA and ET pathways (Table 3) leads to increased defense sensitivity at the cellular level, better known as priming in ISR, and this in turn leads to gene overexpression of AP2/ERF, which is involved in regulating the jasmonate and ethylene pathways. One gene involved in this priming is MYB72, which is overexpressed when Arabidopsis thaliana are inoculated with PGPM, as well as regulation of iron homeostasis and the onset of ISR [126,127]. It has been found that after the onset of ISR, the accumulation of phytoalexins at the infection site is common, and phytoalexins are present in different types of stress. Long-distance signaling by jasmonate and ethylene causes stress in the plant, which leads to long-term adaptations to stresses at the whole-plant level [102,128].

6. Induced Systemic Resistance as a Biocontrol Mechanism

Worldwide use of pesticides has intensified from a consumption of ~1.3 kg/ha in 1992 to 2.57 kg/ha in 2016, an increase of 97.69% [138]. A wide range of bacterial pathogens have developed resistance against commonly used chemicals including antibiotics, such as streptomycin [139]. In addition, the exhaustive use of chemically formulated pest controllers has triggered biodiversity problems such as the loss of pollinators over the years, a phenomenon that threatens biodiversity conservation and sustainable food production worldwide [140]. The use of insecticides can have lethal effects on non-target organisms and reach higher trophic levels through food [141], as well as generating problems for human health related to the use of synthetic agricultural inputs to maintain crop yields.
PGPMs with controlling effects have been widely studied, of which the genus Pseudomonas and Bacillus are the most commercialized [142]. The adequate protection of plants by using biocontrol agents (BCA) is due mainly to the successful colonization of the host and the subsequent production of antibacterial properties, and the production of metabolites that induce systemic resistance and/or the subsequent induction of antibacterial properties [143]. The use of microorganisms capable of inducing ISR may be the key to crop pathogen protection sustainability [144]. For example, Peñafiel-Jaramillo et al. [82] analyzed a selection of genes (Lox2, Tlp1, Eir1, NPR1) in grapevine plants, which were activated for systemic resistance induced by rhizosphere microorganisms. The results showed that the Tlp1 gene was related to the ability of Pseudomonas to develop the ethylene (ET) pathway-induced symbiosis complex. The NPR1 gene was found in leaves and roots, with no significant differences between plants treated with phosphate-buffered saline (PBS) or Pseudomonas veronii R4. This proves how rhizosphere inoculation activates the induced systemic resistance state [82].
In another study, Jaimes-Suárez et al. [87] reported how induction of systemic resistance in tomatoes by Trichoderma koningiopsis Th003 delayed the Fusarium infection. They found that the incidence of root and collar rot in tomatoes was significantly reduced by 35% when compared to the control. The strain Th003 was positive to the induction of systemic resistance in plants since both microorganisms were inoculated spatially and not in direct contact; this effect can be attributed to the fact that Trichoderma stimulates the activity of PR proteins, such as β-1,3-endoglucanases and endochitinases in plants [87]. On the other hand, compared to traditional chemical control, resistance induced by biological agents does not have immediate and potent disease-reducing effects but has a long-lasting effect.
Resistance reactions begin soon after the use of the inducer and the effects generally persist for three weeks to two months, but resistance periods of up to six months have been reported [144]. In addition, plants become resistant to subsequent infections caused by different pathogens [5]. The major disadvantage of this type of resistance is that it can leave the plant unprotected from pathogen attack until the resistance is fully activated [118]. Thus, although ISR through PGPM has been vastly studied and described, research on it should continue, in order to seize its full potential as a method for the biocontrol of phytopathogens.
Exemplifying the above-mentioned ISR mechanisms, Li et al. [145] found that Bacillus amyloliquefaciens strain LJ02 can decrease powdery mildew fungal disease in cucurbits. Cucumber plants inoculated with this strain had a significant production of superoxide dismutase, peroxidase, polyphenol oxidase, and phenylalanine ammonia-lyase as compared to the control. In addition, the accumulation of free AS in cucumber leaves was observed, which was markedly increased in the 5-day treatment compared to the control, as well as overexpression of the PR-1 gene, which is commonly related to SA expression. These findings indicate that B. amyloliquefaciens strain LJ02 can induce systemic resistance in cucumber plants [145].
Mathys et al. [146] analyzed the transcriptomic profile of Arabidopsis thaliana after inoculation of Trichoderma hamatum T382 in roots, which conferred greater resistance to the attack of the necrotrophic pathogen Botrytis cinerea; the analysis of ISR gene expression was carried out by using microarrays. Gene expression analysis allowed the identification of genes related to MAMP-triggered defense and the subsequent production of reactive oxygen species, as well as the identification of genes related to the synthesis of JA and SA. This confirms the activation of induced systemic resistance in A. thaliana by T. hamatum T382 [146].
In addition to transcriptomic analysis of these types of interactions, the root metabolome can show us what metabolites are being produced when PGPM–root interactions are occurring; however, it is a broad niche to analyze as the soil ecosystem is very complex [147]. Bacterial metabolomic analysis of PGPM helps us to elucidate which metabolites are ISR elicitors [147]. Fatima and Anjum [68] analyzed the metabolomic profile of Pseudomonas aeruginosa PM12 in tomato plants, finding an increase in the synthesis of phytoalexins and phenolic compounds, as well as other RSI-related metabolites, concluding that the activation of this type of defense against Fusarium wilt was generated by PGPR [68].

7. Conclusions

Stresses generated by biotic factors in plants modify their normal physiological state, leading to the expression of defense genes, as well as the modification of growth and development. Interaction with pathogens can activate systemic acquired resistance (SAR) after a pathogen attack as a defense mechanism generating molecular cascades that lead to the synthesis of structural barriers and stomatal closure as a first line of defense. Subsequently, activation of programmed cell death can halt the progress of pathogens, and reactive oxygen species lead to the accumulation of oxidants, which attack the pathogen, inhibiting its proliferation. The synthesis of SA and MeSA is indispensable in this type of pathogen-generated biotic stress as they carry this pathogen alertness to other sites in the plant. Understanding this type of defense, as well as the genes expressed in it, is necessary for the control of pathogen-generated stress in crops. On the other hand, stress generated by beneficial microorganisms triggers a series of molecular processes in plants that lead to the activation of induced systemic resistance (ISR); this resistance is generally activated in the rhizosphere when microorganisms (PGPM) interact with it, generating benefits for the plant such as plant growth promotion—thanks to the increased absorption of nutrients in the root—and protection against pathogen attack due to molecular processes, where defense-related gene transcription and the segregation of callose and β-glucans at the site of plant infection act as a physical barrier, preventing plant infection. This type of defense (ISR) generates “beneficial stress” mediated by ET and JA. We can say that it is the equivalent of immunizing the plant against stress generated by pathogens, which results in a sustainable mechanism for pest control and crop adjustment to different stresses, in order to ensure global food security.
Comprehending the induction of these stresses in crops leads us to understand that the application of PGPMs helps reduce biotic and abiotic stresses, as PGPMs, thanks to signaling and establishment in the plant, generate a state of pre-alertness to pathogen attack, which is an alternative to the use of pesticides, being beneficial for the agroecosystem and economically speaking. The plant growth promotion provided by PGPMs helps to reduce the use of fertilizers by promoting plant growth and increasing agricultural production. However, there is still a long way to go in exploring metabolomic and transcriptomic approaches to plant–microorganism interactions and how they mitigate biotic stress.

Author Contributions

Conceptualization, S.d.l.S.-V. and F.I.P.-C.; writing—original draft preparation, A.M.G.-M., A.C.M.-M., P.H.M.-S., F.I.P.-C. and S.d.l.S.-V.; writing—review and editing, all authors; visualization, all authors; supervision, F.I.P.-C. and S.d.l.S.-V. All authors have read and agreed to the published version of the manuscript.

Funding

Ana María García-Montelongo was supported by a CONACYT master’s degree fellowship (number 1075003), and Amelia C. Montoya-Martínez was supported by CONACYT postdoctoral fellowship (number 2306476). This study was supported by the PROFAPI ITSON project 2023_001.

Data Availability Statement

Not applicable.

Acknowledgments

We thank all members of Laboratorio de Biotecnología del Recurso Microbiano www.itson.mx/LBRM, (accessed on 12 December 2022).

Conflicts of Interest

The authors declare that there are no conflict of interest.

References

  1. Tian, Z.; Wang, J.W.; Li, J.; Han, B. Designing future crops: Challenges and strategies for sustainable agriculture. Plant J. 2021, 105, 1165–1178. [Google Scholar] [CrossRef] [PubMed]
  2. Lassalle, G. Monitoring natural and anthropogenic plant stressors by hyperspectral remote sensing: Recommendations and guidelines based on a meta-review. Sci. Total Environ. 2021, 788, 147758. [Google Scholar] [CrossRef] [PubMed]
  3. Taiz, L.; Zeiger, E. Plant Physiology. 5th Edition. 2010. Available online: https://scholar.google.es/scholar?hl=es&as_sdt=0%2C5&q=Taiz%2C+L.+and+Zeiger%2C+E.+%282010%29+Plant+Physiology.+5th+Edition&btnG= (accessed on 8 November 2022).
  4. Verma, S.; Nizam, S.; Verma, P.K. Biotic and Abiotic Stress Signaling in Plants. Stress Signaling in Plants: Genomics and Proteomics Perspective. In Stress Signaling in Plants: Genomics and Proteomics Perspective; Sarwat, M., Ahmad, A., Abdin, M., Eds.; Springer: New York, NY, USA, 2013; Volume 1, pp. 25–50. Available online: https://link.springer.com/chapter/10.1007/978-1-4614-6372-6_2 (accessed on 8 November 2022).
  5. Agrios, G.N. Plant Pathology, 5th ed.; Academic Press: Cambridge, MA, USA, 2009; p. 952. ISBN 978-0120445653. [Google Scholar]
  6. Andersen, E.J.; Ali, S.; Byamukama, E.; Yen, Y.; Nepal, M.P. Disease Resistance Mechanisms in Plants. Genes 2018, 9, 339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Chaudhary, D.P.; Kumar, A.; Mandhania, S.S.; Srivastava, P.; Kumar, R.S. MAIZE AS FODDER? Available online: www.maizeindia.org (accessed on 5 December 2021).
  8. De los Santos-Villalobos, S.; Parra-Cota, F.I. Current trends in plant growth-promoting microorganisms research for sustainable food security. Curr. Res. Microb. Sci. 2021, 2, 100016. [Google Scholar] [CrossRef]
  9. Zhao, C.; Liu, B.; Piao, S.; Wang, X.; Lobell, D.B.; Huang, Y.; Huang, M.T.; Yao, Y.T.; Bassu, S.; Ciais, P.; et al. Temperature Increase Reduces Global Yields of Major Crops in Four Independent Estimates. Proc. Natl. Acad. Sci. USA 2017, 114, 9326–9331. Available online: https://www.pnas.org/doi/abs/10.1073/pnas.1701762114 (accessed on 8 November 2022). [CrossRef] [PubMed] [Green Version]
  10. Varshney, R.K.; Bansal, K.C.; Aggarwal, P.K.; Datta, S.K.; Craufurd, P.Q. Agricultural Biotechnology for Crop Improvement in a Variable Climate: Hope or Hype? Trends Plant Sci. 2011, 16, 363–371. Available online: http://www.cell.com/article/S1360138511000525/fulltext (accessed on 8 November 2022). [CrossRef] [Green Version]
  11. Ghini, R.; Bettiol, W.; Hamada, E. Diseases in tropical and plantation crops as affected by climate changes: Current knowledge and perspectives. Plant Pathol. 2011, 60, 122–132. [Google Scholar] [CrossRef]
  12. Sarker, A.; Ansary, M.W.R.; Hossain, M.N.; Islam, T. Prospect and Challenges for Sustainable Management of Climate Change-Associated Stresses to Soil and Plant Health by Beneficial Rhizobacteria. Stresses 2021, 1, 200–222. [Google Scholar] [CrossRef]
  13. Dean, R.; Van Kan, J.A.L.; Pretorius, Z.A.; Hammond-Kosack, K.E.; Di Pietro, A.; Spanu, P.D.; Rudd, J.J.; Dickman, M.; Kahmann, R.; Ellis, J.; et al. The Top 10 fungal pathogens in molecular plant pathology. Mol. Plant Pathol. 2012, 13, 414–430. [Google Scholar] [CrossRef] [Green Version]
  14. Nazarov, P.A.; Baleev, D.N.; Ivanova, M.I.; Sokolova, L.M.; Karakozova, M.V. Infectious plant diseases: Etiology, current status, problems and prospects in plant protection. Acta Nat. 2020, 12, 46–59. [Google Scholar] [CrossRef]
  15. Fernandez, J.; Orth, K. Rise of a Cereal Killer: The Biology of Magnaporthe oryzae Biotrophic Growth. Trends Microbiol. 2018, 26, 582–597. [Google Scholar] [CrossRef] [PubMed]
  16. Hua, L.; Yong, C.; Zhanquan, Z.; Boqiang, L.; Guozheng, Q.; Shiping, T. Pathogenic mechanisms and control strategies of Botrytis cinerea causing post-harvest decay in fruits and vegetables. Food Qual. Saf. 2018, 2, 111–119. [Google Scholar] [CrossRef] [Green Version]
  17. FAO. FAO Wheat Rust Disease Global Programme 2014–2017. Available online: https://www.fao.org/agriculture/crops/wheatrust (accessed on 27 December 2022).
  18. Figueroa, M.; Hammond-Kosack, K.E.; Solomon, P.S. A review of wheat diseases-A field perspective. Mol. Plant Pathol. 2018, 19, 1523–1536. [Google Scholar] [CrossRef]
  19. Montoya-Martínez, A.C.; Parra-Cota, F.I.; De los Santos-Villalobos, S. Beneficial Microorganisms in Sustainable Agriculture: Harnessing Microbes’ Potential to Help Feed the World. Plants 2022, 11, 372. [Google Scholar] [CrossRef] [PubMed]
  20. Santoyo, G.; Moreno-Hagelsieb, G.; Orozco-Mosqueda, M.C.; Glick, B.R. Plant growth-promoting bacterial endophytes. Microbiol. Res. 2016, 183, 92–99. [Google Scholar] [CrossRef] [PubMed]
  21. Schlechter, R.O.; Miebach, M.; Remus-Emsermann, M.N. Driving factors of epiphytic bacterial communities: A review. J. Adv. Res. 2019, 19, 57–65. [Google Scholar] [CrossRef] [PubMed]
  22. Knief, C.; Delmotte, N.; Chaffron, S.; Stark, M.; Innerebner, G.; Wassmann, R.; Von Mering, C.; Vorholt, J.A. Metaproteogenomic analysis of microbial communities in the phyllosphere and rhizosphere of rice. ISME J. 2011, 6, 1378–1390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Compant, S.; Samad, A.; Faist, H.; Sessitsch, A. A review on the plant microbiome: Ecology, functions, and emerging trends in microbial application. J. Adv. Res. 2019, 190, 29–37. [Google Scholar] [CrossRef]
  24. Hakim, S.; Naqqash, T.; Nawaz, M.S.; Laraib, I.; Siddique, M.J.; Zia, R.; Mirza, M.S.; Imran, A. Rhizosphere Engineering With Plant Growth-Promoting Microorganisms for Agriculture and Ecological Sustainability. Front. Sustain. Food Syst. 2021, 5, 617157. [Google Scholar] [CrossRef]
  25. Santoyo, G. How plants recruit their microbiome? New insights into beneficial interactions. J. Adv. Res. 2022, 40, 45–58. [Google Scholar] [CrossRef]
  26. Wirthmueller, L.; Maqbool, A.; Banfield, M. On the front line: Structural insights into plant-pathogen interactions. Nat. Rev. Genet. 2013, 11, 761–776. [Google Scholar] [CrossRef] [PubMed]
  27. Velásquez, A.C.; Castroverde, C.D.M.; He, S.Y. Plant–Pathogen Warfare under Changing Climate Conditions. Curr. Biol. 2018, 28, R619–R634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Berg, G.; Grube, M.; Schloter, M.; Smalla, K. Unraveling the plant microbiome: Looking back and future perspectives. Front. Microbiol. 2014, 5, 148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Nifakos, K.; Tsalgatidou, P.C.; Thomloudi, E.-E.; Skagia, A.; Kotopoulis, D.; Baira, E.; Delis, C.; Papadimitriou, K.; Markellou, E.; Venieraki, A.; et al. Genomic Analysis and Secondary Metabolites Production of the Endophytic Bacillus velezensis Bvel1: A Biocontrol Agent against Botrytis cinerea Causing Bunch Rot in Post-Harvest Table Grapes. Plants 2021, 10, 1716. [Google Scholar] [CrossRef] [PubMed]
  30. Tronsmo, A.M.; Collinge, D.B.; Djurle, A.; Munk, L.; Yuen, J.; Tronsmo, A. Plant Pathology and Plant Diseases; CAB International: Boston, MA, USA, 2020. [Google Scholar]
  31. Rangel Sánchez, G.; Castro Mercado, E.; Beltran Peña, E.; Reyes de la Cruz, H.; García Pineda, E. El acido salicílico y su participación en la resistencia a patógenos en plantas. Biológicas 2010, 12, 90–95. [Google Scholar]
  32. Lamb, C.J.; Lawton, M.A.; Dron, M.; Dixon, R.A. Signals and Transduction Mechanisms for Activation of Plant Defenses Against Microbial Attack. Cell 1989, 56, 215–224. Available online: https://pubmed.ncbi.nlm.nih.gov/2643475/ (accessed on 21 November 2022). [CrossRef]
  33. Matei, A.; Doehlemann, G. Cell biology of corn smut disease—Ustilago maydis as a model for biotrophic interactions. Curr. Opin. Microbiol. 2016, 34, 60–66. [Google Scholar] [CrossRef]
  34. Lamour, K.; Stam, R.; Jupe, J.; Huitema, E. The oomycete broad-host-range pathogen Phytophthora capsici. Mol. Plant Pathol. 2011, 13, 329–337. [Google Scholar] [CrossRef]
  35. Williamson, B.; Tudzynski, B.; Tudzynski, P.; Van Kan, J.A.L. Botrytis cinerea: The cause of grey mould disease. Mol. Plant Pathol. 2007, 8, 561–580. [Google Scholar] [CrossRef]
  36. Villa-Rodriguez, E.; Lugo-Enríquez, C.; Santos-Villalobos, S.D.L.; Parra-Cota, F.I.; Figueroa-López, P. First Report of Cochliobolus sativus Causing Spot Blotch on Durum Wheat (Triticum durum) in The Yaqui Valley, Mexico. Plant Dis. 2016, 100, 2329. [Google Scholar] [CrossRef]
  37. Bartholomew, E.S.; Xu, S.; Zhang, Y.; Yin, S.; Feng, Z.; Chen, S.; Sun, L.; Yang, S.; Wang, Y.; Liu, P.; et al. A Chitinase CsChi23 Promoter Polymorphism Underlies Cucumber Resistance against Fusarium oxysporum f. sp. cucumerinum. New Phytol. 2022, 236, 1471–1486. Available online: https://onlinelibrary.wiley.com/doi/full/10.1111/nph.18463 (accessed on 15 December 2022). [CrossRef] [PubMed]
  38. Burkhanova, G.F.; Veselova, S.V.; Sorokan, A.V.; Blagova, D.K.; Nuzhnaya, T.V.; Maksimov, I.V. Strains of Bacillus ssp. Regulate Wheat Resistance to Septoria nodorum Berk. Appl. Biochem. Microbiol. 2017, 53, 346–352. Available online: https://link.springer.com/article/10.1134/S0003683817030048 (accessed on 17 December 2022). [CrossRef]
  39. Lian, L.; Xie, L.; Zheng, L.; Lin, Q. Induction of Systemic Resistance in Tobacco against Tobacco Mosaic Virus by Bacillus spp. Biocontrol Sci. Technol. 2011, 21, 281–292. Available online: https://www.tandfonline.com/doi/abs/10.1080/09583157.2010.543667 (accessed on 17 December 2022). [CrossRef]
  40. Fürst, U.; Zeng, Y.; Albert, M.; Witte, A.K.; Fliegmann, J.; Felix, G. Perception of Agrobacterium Tumefaciens Flagellin by FLS2XL Confers Resistance to Crown Gall Disease. Nat. Plants 2020, 6, 22–27. Available online: https://www.nature.com/articles/s41477-019-0578-6 (accessed on 16 January 2022). [CrossRef] [PubMed]
  41. Lahoz, E.; Contillo, R.; Porrone, F. Induction of Systemic Resistance to Erysiphe orontii Cast in Tobacco by Application on Roots of an Isolate of Gliocladium roseum Bainier. J. Phytopathol. 2004, 152, 465–470. Available online: https://onlinelibrary.wiley.com/doi/full/10.1111/j.1439-0434.2004.00876.x (accessed on 18 December 2022). [CrossRef]
  42. Parra-Cota, F.I.; García-Pereyra, J.; Aviña-Martínez, G.N.; De los Santos-Villalobos, S. First Report of Fusarium Wilt on Citrus Sinensis Var. Valencia in the Yaqui Valley, Mexico. Mex. J. Phytopathol. 2018, 37, 193–201. Available online: https://www.smf.org.mx/rmf/ojs/index.php/RMF/article/view/139 (accessed on 18 December 2022).
  43. Garcia Pereyra, J.; Aviña Martínez, G.N.; De los Santos Villalobos, S.; Garcia Montelongo, A.M.; Alejandre-Iturbide, G.; Rubio Graciano, R.B. Biological Control of Erwinia Amylovora in Apple Trees Employing Antibacterial Agents. AshEse J. Agric. Sci. 2020, 3, 117–126. Available online: http://www.ashese.co.uk/ajas-issues/biological-control-of-erwinia-amylovora-in-apple-trees-employing-antibacterial-agents (accessed on 18 December 2022).
  44. Silva, M.S.S.; Arraes, F.B.M.; Campos, M.; Grossi-de-Sa, M.F.; Fernandez, D.; Cândido, E.D.; Cardoso, M.H.; Franco, O.L.; Grossi-de-Sa, M. Review: Potential biotechnological assets related to plant immunity modulation applicable in engineering disease-resistant crops. Plant Sci. 2018, 270, 72–84. [Google Scholar] [CrossRef]
  45. Castro-Moretti, F.R.; Gentzel, I.N.; Mackey, D.; Alonso, A.P. Metabolomics as an emerging tool for the study of plant–pathogen interactions. Metabolites 2020, 10, 52. [Google Scholar] [CrossRef] [Green Version]
  46. Glazebrook, J. Contrasting mechanisms of defense against biotrophic and necrotrophic pathogens. Annu. Rev. Phytopathol. 2005, 43, 205–227. [Google Scholar] [CrossRef]
  47. Chisholm, S.T.; Coaker, G.; Day, B.; Staskawicz, B.J. Host-microbe interactions: Shaping the evolution of the plant immune response. Cell 2006, 124, 803–814. [Google Scholar] [CrossRef] [Green Version]
  48. Keen, N.T. Gene-for-gene complementarity in plant-pathogen interactions. Annu. Rev. Genet. 1990, 24, 447–463. [Google Scholar] [CrossRef] [PubMed]
  49. Bigeard, J.; Colcombet, J.; Hirt, H. Signaling mechanisms in pattern-triggered immunity (PTI). Mol. Plant 2015, 8, 521–539. [Google Scholar] [CrossRef] [PubMed]
  50. Ma, W.; Smigel, A.; Tsai, Y.C.; Braam, J.; Berkowitz, G.A. Innate immunity signaling: Cytosolic Ca2+ elevation is linked to downstream nitric oxide generation through the action of calmodulin or a calmodulin-like protein. Plant Physiol. 2008, 148, 818–828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Kovtun, Y.; Chiu, W.L.; Tena, G.; Sheen, J. Functional analysis of oxidative stress-activated mitogen-activated protein kinase cascade in plants. Proc. Natl. Acad. Sci. USA 2000, 97, 2940–2945. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Noman, A.; Aqeel, M.; Qari, S.H.; Al Surhanee, A.A.; Yasin, G.; Alamri, S.; Hashem, M.; Al-Saadi, A.M. Plant hypersensitive response vs pathogen ingression: Death of few gives life to others. Microb. Pathog. 2020, 145, 104224. [Google Scholar] [CrossRef]
  53. Reichheld, J.; Vernoux, T.; Lardon, F.; Van Montagu, M.; Inzé, D. Specific checkpoints regulate plant cell cycle progression in response to oxidative stress. Plant J. 1999, 17, 647–656. [Google Scholar] [CrossRef]
  54. Khurana, S.M.P.; Pandey, S.K.; Sarkar, D.; Chanemougasoundharam, A. Apoptosis in plant disease response: A close encounter of the pathogen kind. Curr. Sci. 2005, 88, 740–752. [Google Scholar]
  55. Dangl, J.L.; Dietrich, R.A.; Richberg, M.H. Death Don’t Have No Mercy: Cell Death Programs in Plant-Microbe Interactions. Plant Cell 1996, 8, 1793. [Google Scholar] [CrossRef]
  56. Dempsey, D.A.; Klessig, D.F. SOS—Too Many Signals for Systemic Acquired Resistance? Trends Plant Sci. 2012, 17, 538–545. Available online: https://pubmed.ncbi.nlm.nih.gov/22749315/ (accessed on 21 November 2022). [CrossRef]
  57. Forouhar, F.; Yang, Y.; Kumar, D.; Chen, Y.; Fridman, E.; Park, S.W.; Chiang, Y.; Acton, T.B.; Montelione, G.T.; Pichersky, E.; et al. Structural and biochemical studies identify tobacco SABP2 as a methyl salicylate esterase and implicate it in plant innate immunity. Proc. Natl. Acad. Sci. USA 2005, 102, 1773–1778. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Diaz-Puentes, L.N. Systemic acquired resistance induced by salicylic acid. Biotecnol. En El Sect. Agropecu. Y Agroind. 2012, 10, 257–267. [Google Scholar]
  59. Chakraborty, S.; Moeder, W.; Yoshioka, K. Plant Immunity. Ref. Modul. Life Sci. 2017, 1–8. Available online: https://linkinghub.elsevier.com/retrieve/pii/B9780128096338121545 (accessed on 21 November 2022).
  60. 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. Available online: https://www.annualreviews.org/doi/abs/10.1146/annurev-phyto-082712-102340 (accessed on 10 November 2022). [CrossRef] [PubMed] [Green Version]
  61. Niderman, T.; Genetet, I.; Bruyere, T.; Gees, R.; Stintzi, A.; Legrand, M.; Fritig, B.; Mosinger, E. Pathogenesis-Related PR-1 Proteins Are Antifungal. Plant Physiol. 1995, 108, 17–27. [Google Scholar] [CrossRef] [Green Version]
  62. Bigeard, J.; Hirt, H. Nuclear Signaling of Plant MAPKs. Front. Plant Sci. 2018, 9, 469. [Google Scholar] [CrossRef]
  63. Moore, J.W.; Loake, G.; Spoel, S.H. Transcription Dynamics in Plant Immunity. Plant Cell 2011, 23, 2809–2820. [Google Scholar] [CrossRef] [Green Version]
  64. Schlaeppi, K.; Mauch, F. Indolic secondary metabolites protect Arabidopsis from the oomycete pathogen Phytophthora brassicae. Plant Signal. Behav. 2010, 5, 1099–1101. [Google Scholar] [CrossRef] [Green Version]
  65. Iakimova, E.T.; Sobiczewski, P.; Michalczuk, L.; Węgrzynowicz-Lesiak, E.; Mikiciński, A.; Woltering, E.J. Morphological and biochemical characterization of Erwinia amylovora-induced hypersensitive cell death in apple leaves. Plant Physiol. Biochem. 2013, 63, 292–305. [Google Scholar] [CrossRef]
  66. Moya-Elizondo, E.A.; Jacobsen, B.J. Integrated management of Fusarium crown rot of wheat using fungicide seed treatment, cultivar resistance, and induction of systemic acquired resistance (SAR). Biol. Control. 2016, 92, 153–163. [Google Scholar] [CrossRef]
  67. Köhl, J.; Kolnaar, R.; Ravensberg, W.J. Mode of Action of Microbial Biological Control Agents Against Plant Diseases: Relevance Beyond Efficacy. Front. Plant Sci. 2019, 10, 845. [Google Scholar] [CrossRef] [Green Version]
  68. Fatima, S.; Anjum, T. Identification of a Potential ISR Determinant from Pseudomonas aeruginosa PM12 against Fusarium Wilt in Tomato. Front. Plant Sci. 2017, 8, 848. [Google Scholar] [CrossRef] [Green Version]
  69. Poveda, J.; Hermosa, R.; Monte, E.; Nicolás, C. The Trichoderma harzianum Kelch Protein ThKEL1 Plays a Key Role in Root Colonization and the Induction of Systemic Defense in Brassicaceae Plants. Front. Plant Sci. 2019, 10, 1478. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Etesami, H.; Maheshwari, D.K. Use of Plant Growth Promoting Rhizobacteria (PGPRs) with Multiple Plant Growth Promoting Traits in Stress Agriculture: Action Mechanisms and Future Prospects. Ecotoxicol. Environ. Saf. 2018, 156, 225–246. Available online: https://pubmed.ncbi.nlm.nih.gov/29554608/ (accessed on 21 November 2022). [CrossRef] [PubMed]
  71. Mhlongo, M.I.; Piater, L.A.; Madala, N.E.; Labuschagne, N.; Dubery, I.A. The Chemistry of Plant–Microbe Interactions in the Rhizosphere and the Potential for Metabolomics to Reveal Signaling Related to Defense Priming and Induced Systemic Resistance. Front. Plant Sci. 2018, 9, 112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Mar Vázquez, M.; César, S.; Azcón, R.; Barea, J.M. Interactions between arbuscular mycorrhizal fungi and other microbial inoculants (Azospirillum, Pseudomonas, Trichoderma) and their effects on microbial population and enzyme activities in the rhizosphere of maize plants. Appl. Soil Ecol. 2000, 15, 261–272. [Google Scholar] [CrossRef]
  73. Goswami, D.; Thakker, J.N.; Dhandhukia, P.C. Portraying Mechanics of Plant Growth Promoting Rhizobacteria (PGPR): A Review. Cogent Food Agric. 2016, 2, 1127500. Available online: https://www.tandfonline.com/doi/abs/10.1080/23311932.2015.1127500 (accessed on 5 December 2021). [CrossRef]
  74. Fernandes, C.; Domingues, D.; Cecato, U.; Biserra, T.T.; Mamédio, D.; Galbeiro, S. Azospirillum spp. on grasses and forage crops. Review. Rev. Mex. De Cienc. Pecu. 2020, 11, 223–240. [Google Scholar] [CrossRef]
  75. Yang, J.; Lan, L.; Jin, Y.; Yu, N.; Wang, D.; Wang, E. Mechanisms underlying legume–rhizobium symbioses. J. Integr. Plant Biol. 2021, 64, 244–267. [Google Scholar] [CrossRef]
  76. Mommer, L.; Kirkegaard, J.; Van Ruijven, J. Root–Root Interactions: Towards A Rhizosphere Framework. Trends Plant Sci. 2016, 21, 209–217. [Google Scholar] [CrossRef] [PubMed]
  77. 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]
  78. 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. 2014, 22, 4907–4921. [Google Scholar] [CrossRef] [PubMed]
  79. Liu, W.-W.; Mu, W.; Zsu, B.-Y.; Du, Y.-C.; Liu, F. Antagonistic Activities of Volatiles from Four Strains of Bacillus spp. and Paenibacillus spp. Against Soil-Borne Plant Pathogens. Agric. Sci. China 2008, 7, 1104–1114. [Google Scholar] [CrossRef]
  80. 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] [PubMed] [Green Version]
  81. Rojas Padilla, J.; Encinas, L.A.C.; Montoya, R.I.R.; De Los Santos Villalobos, S. Growth promotion on wheat (Triticum turgidum L. subsp. durum) by co-inoculation of native Bacillus strains isolated from the Yaqui Valley, Mexico. Nova Sci. 2020, 12. Epub 2 July 2020. Available online: https://doi.org/10.21640/ns.v12i24.2136 (accessed on 16 December 2022).
  82. Peñafiel-Jaramillo, M.F.; Sánchez-Sepúlveda, E.; Cruz-Rosero, N.; Belezaca-Pinargote, C.; Prieto-Encalada, H.G.; Martínez, H.F.C. Activación de resistencia sistémica inducida en vid “Thompson Seedless”, en respuesta Pseudomonas veronii R4. Cienc. Y Tecnol. 2016, 9, 1–9. Available online: https://revistas.uteq.edu.ec/index.php/cyt/article/view/161 (accessed on 26 April 2022). [CrossRef]
  83. Toral, L.; Rodríguez, M.; Béjar, V.; Sampedro, I. Crop Protection against Botrytis cinerea by Rhizhosphere Biological Control Agent Bacillus velezensis XT1. Microorganisms 2020, 8, 992. Available online: https://doi.org/10.3390/microorganisms8070992 (accessed on 23 August 2022). [CrossRef] [PubMed]
  84. Villa-Rodríguez, E.; Parra-Cota, F.; Castro-Longoria, E.; López-Cervantes, J.; De Los Santos-Villalobos, S. Bacillus subtilis TE3: A promising biological control agent against Bipolaris sorokiniana, the causal agent of spot blotch in wheat (Triticum turgidum L. subsp. durum). Biol. Control 2019, 132, 135–143. [Google Scholar] [CrossRef]
  85. Lee, G.; Lee, S.-H.; Kim, K.M.; Ryu, C.-M. Foliar application of the leaf-colonizing yeast Pseudozyma churashimaensis elicits systemic defense of pepper against bacterial and viral pathogens. Sci. Rep. 2017, 7, 39432. Available online: www.nature.com/scientificreports (accessed on 17 December 2022). [CrossRef] [Green Version]
  86. Léon-Kloosterziel, K.M.; Verhagen, B.W.M.; Keurentjes, J.J.B.; Van Pelt, J.A.; Rep, M.; Van Loon, L.C.; Pieterse, C.M. Colonization of the Arabidopsis rhizosphere by fluorescent Pseudomonas spp. activates a root-specific, ethylene-responsive PR-5 gene in the vascular bundle. Plant Mol. Biol. 2005, 57, 731–748. Available online: http://www.bio.uu.nl/$fytopath (accessed on 17 December 2022). [CrossRef] [Green Version]
  87. Jaimes-Suárez, Y.Y.; Velandia, C.A.M.; Prado, A.M.C. Inducción de resistencia sistémica contra Fusarium oxysporum en tomate por Trichoderma koningiopsis Th003. Acta Biológica Colomb. 2009, 14, 111–120. Available online: https://revistas.unal.edu.co/index.php/actabiol/article/view/1344 (accessed on 1 March 2022).
  88. Félix-Pablos, C.M.; Parra-Cota, F.I.; Santoyo, G.; Orozco-Mosqueda, M.; Del, C.; De los Santos-Villalobos, S. Draft genome sequence of Bacillus sp. strain FSQ1, a biological control agent against white mold in common bean (Phaseolus vulgaris L.). Curr. Res. Microb. Sci. 2022, 3, 100138. [Google Scholar] [CrossRef]
  89. Ortega-Urquieta, M.E.; Valenzuela-Ruíz, V.; Mitra, D.; Hyder, S.; Elsheery, N.I.; Kumar Das Mohapatra, P.; Parra-Cota, F.I.; De los Santos-Villalobos, S. Draft Genome Sequence of Priestia sp. Strain TSO9, a Plant Growth-Promoting Bacterium Associated with Wheat (Triticum turgidum subsp. durum) in the Yaqui Valley, Mexico. Plants 2022, 11, 2231. Available online: https://www.mdpi.com/2223-7747/11/17/2231/htm (accessed on 18 December 2022). [CrossRef] [PubMed]
  90. Valenzuela-Ruiz, V.; Robles-Montoya, R.I.; Parra-Cota, F.I.; Santoyo, G.; Orozco-Mosqueda, M.D.C.; Rodríguez-Ramírez, R.; Santos-Villalobos, S.D.L. Draft genome sequence of Bacillus paralicheniformis TRQ65, a biological control agent and plant growth-promoting bacterium isolated from wheat (Triticum turgidum subsp. durum) rhizosphere in the Yaqui Valley, Mexico. 3 Biotech 2019, 9, 436. Available online: https://link.springer.com/article/10.1007/s13205-019-1972-5 (accessed on 18 December 2022). [CrossRef] [PubMed]
  91. Valenzuela-Ruiz, V.; Parra-Cota, F.I.; Santoyo, G.; De los Santos-Villalobos, S. Potential biocontrol mechanisms of Bacillus sp. TSO2 against Bipolaris sorokiniana, spot blotch in wheat. Mex. J. Phytopathol. 2022, 40, 230–239. Available online: https://doi.org/10.18781/R.MEX.FIT.2201-1 (accessed on 18 December 2022). [CrossRef]
  92. Villa-Rodriguez, E.; Lugo-Enríquez, C.; Ferguson, S.; Parra-Cota, F.I.; Cira-Chávez, L.A.; De los Santos-Villalobos, S. Trichoderma harzianum sensu lato TSM39: A wheat microbiome fungus that mitigates spot blotch disease of wheat (Triticum turgidum L. subsp. durum) caused by Bipolaris sorokiniana. Biol. Control. 2022, 175, 105055. [Google Scholar] [CrossRef]
  93. Morales-Cedeño, L.R.; Santos-Villalobos, S.D.L.; Santoyo, G. Functional and Genomic Analysis of Rouxiella badensis SER3 as a Novel Biocontrol Agent of Fungal Pathogens. Front. Microbiol. 2021, 12, 2184. [Google Scholar] [CrossRef]
  94. Khoshru, B.; Mitra, D.; Khoshmanzar, E.; Myo, E.M.; Uniyal, N.; Mahakur, B.; Das Mohapatra, P.K.; Panneerselvam, P.; Boutaj, H.; Alizadeh, M.; et al. Current scenario and future prospects of plant growth-promoting rhizobacteria: An economic valuable resource for the agriculture revival under stressful conditions. J. Plant Nutr. 2020, 43, 3062–3092. Available online: https://www.tandfonline.com/doi/abs/10.1080/01904167.2020.1799004 (accessed on 8 November 2022). [CrossRef]
  95. Cesari, A.; Paulucci, N.; López-Gómez, M.; Hidalgo-Castellanos, J.; Plá, C.L.; Dardanelli, M.S. Restrictive water condition modifies the root exudates composition during peanut-PGPR interaction and conditions early events, reversing the negative effects on plant growth. Plant Physiol. Biochem. 2019, 142, 519–527. [Google Scholar] [CrossRef]
  96. Dutta, S.; Podile, A.R. Plant Growth Promoting Rhizobacteria (PGPR): The bugs to debug the root zone. Crit. Rev. Microbiol. 2010, 36, 232–244. Available online: https://www.tandfonline.com/doi/abs/10.3109/10408411003766806 (accessed on 8 November 2022). [CrossRef]
  97. Basu, A.; Prasad, P.; Das, S.N.; Kalam, S.; Sayyed, R.Z.; Reddy, M.S.; El Enshasy, H. Plant Growth Promoting Rhizobacteria (PGPR) as Green Bioinoculants: Recent Developments, Constraints, and Prospects. Sustainability 2021, 13, 1140. Available online: https://www.mdpi.com/2071-1050/13/3/1140/htm (accessed on 8 November 2022).
  98. De Souza, R.; Ambrosini, A.; Passaglia, L.M.P. Plant growth-promoting bacteria as inoculants in agricultural soils. Genet Mol Biol. 2015, 38, 401–419. [Google Scholar] [CrossRef] [PubMed]
  99. Ketta, H.A.; Elkhateeb, N.M.; Saleh, M.M.; Kamel, S.M. Efficiency Assessment of Combinations Between Rhizobium leguminosarum and Trichoderma spp. for Controlling of Pea (Pisum sativum L.) Damping-off Disease. Egypt. J. Phytopathol. 2021, 49, 1–14. Available online: https://ejp.journals.ekb.eg/article_139637.html (accessed on 1 January 2023). [CrossRef]
  100. Sibponkrung, S.; Kondo, T.; Tanaka, K.; Tittabutr, P.; Boonkerd, N.; Yoshida, K.-I.; Teaumroong, N. Co-Inoculation of Bacillus velezensis Strain S141 and Bradyrhizobium Strains Promotes Nodule Growth and Nitrogen Fixation. Microorganisms 2020, 8, 678. Available online: https://www.mdpi.com/2076-2607/8/5/678/htm (accessed on 1 January 2023). [CrossRef] [PubMed]
  101. Timm, C.M.; Carter, K.R.; Carrell, A.A.; Jun, S.-R.; Jawdy, S.S.; Vélez, J.M.; Gunter, L.E.; Yang, Z.; Nookaew, I.; Engle, N.L.; et al. Abiotic Stresses Shift Belowground Populus-Associated Bacteria Toward a Core Stress Microbiome. Msystems 2018, 3, e00070-17. Available online: https://doi.org/10.1128/mSystems.00070-17 (accessed on 8 November 2022). [CrossRef] [PubMed] [Green Version]
  102. Liu, H.; Brettell, L.E.; Qiu, Z.; Singh, B.K. Microbiome-Mediated Stress Resistance in Plants. Trends Plant Sci. 2020, 25, 733–743. Available online: https://pubmed.ncbi.nlm.nih.gov/32345569/ (accessed on 8 November 2022). [CrossRef]
  103. Chi, F.; Shen, S.-H.; Cheng, H.-P.; Jing, Y.-X.; Yanni, Y.G.; Dazzo, F.B. Ascending Migration of Endophytic Rhizobia, from Roots to Leaves, inside Rice Plants and Assessment of Benefits to Rice Growth Physiology. Appl. Environ. Microbiol. 2005, 71, 7271–7278. Available online: https://journals.asm.org/doi/10.1128/AEM.71.11.7271-7278.2005 (accessed on 8 November 2022). [CrossRef] [Green Version]
  104. Buscaill, P.; Van der Hoorn, R.A.L. Defeated by the Nines: Nine Extracellular Strategies to Avoid Microbe-Associated Molecular Patterns Recognition in Plants. Plant Cell 2021, 33, 2116–2130. Available online: https://academic.oup.com/plcell/article/33/7/2116/6237920 (accessed on 8 November 2022). [CrossRef]
  105. Jones, J.D.G.; Dangl, J.L. The plant immune system. Nature 2006, 444, 323–329. Available online: https://www.nature.com/articles/nature05286 (accessed on 8 November 2022). [CrossRef]
  106. Lebeis, S.L.; Paredes, S.H.; Lundberg, D.S.; Breakfield, N.; Gehring, J.; McDonald, M.; Malfatti, S.; Glavina del rio, T.; Jones, C.D.; Tringe, S.G.; et al. Salicylic acid modulates colonization of the root microbiome by specific bacterial taxa. Science 2015, 349, 860–864. Available online: https://www.science.org/doi/10.1126/science.aaa8764 (accessed on 8 November 2022). [CrossRef] [Green Version]
  107. Nakagawa, T.; Kawaguchi, M. Shoot-applied MeJA Suppresses Root Nodulation in Lotus japonicus. Plant Cell Physiol. 2006, 47, 176–180. Available online: https://academic.oup.com/pcp/article/47/1/176/1867512 (accessed on 8 November 2022).
  108. Ronald, P.C.; Beutler, B. Plant and animal sensors of conserved microbial signatures. Science 2010, 330, 1061–1064. Available online: https://www.science.org/doi/10.1126/science.1189468 (accessed on 8 November 2022). [CrossRef] [Green Version]
  109. Hacquard, S.; Spaepen, S.; Garrido-Oter, R.; Schulze-Lefert, P. Interplay Between Innate Immunity and the Plant Microbiota. Annu. Rev. Phytopathol. 2017, 55, 565–589. Available online: https://www.annualreviews.org/doi/abs/10.1146/annurev-phyto-080516-035623 (accessed on 8 November 2022). [CrossRef] [PubMed]
  110. Macho, A.P.; Zipfel, C. Plant PRRs and the activation of innate immune signaling. Mol. Cell 2014, 54, 263–272. Available online: https://pubmed.ncbi.nlm.nih.gov/24766890/ (accessed on 8 November 2022). [CrossRef] [PubMed] [Green Version]
  111. Burbano-Figueroa, Ó. Resistencia de plantas a patógenos: Una revisión sobre los conceptos de resistencia vertical y horizontal. Rev. Argent. De Microbiol. 2020, 52, 245–255. Available online: https://www.elsevier.es/es-revista-revista-argentina-microbiologia-372-articulo-resistencia-plantas-patogenos-una-revision-S0325754120300328 (accessed on 8 November 2022). [CrossRef] [PubMed]
  112. Shah, J.; Zeier, J. Long-distance communication and signal amplification in systemic acquired resistance. Front. Plant Sci. 2013, 4, 30. Available online: https://pubmed.ncbi.nlm.nih.gov/23440336/ (accessed on 8 November 2022). [CrossRef] [PubMed] [Green Version]
  113. Millet, Y.A.; Danna, C.H.; Clay, N.K.; Songnuan, W.; Simon, M.D.; Werck-Reichhart, D.; Ausubel, F.M. Innate Immune Responses Activated in Arabidopsis Roots by Microbe-Associated Molecular Patterns. Plant Cell 2010, 22, 973–990. Available online: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2861455/ (accessed on 8 November 2022). [CrossRef] [PubMed] [Green Version]
  114. Lambais, M.R.; Barrera, S.E.; Santos, E.C.; Crowley, D.E.; Jumpponen, A. Phyllosphere Metaproteomes of Trees from the Brazilian Atlantic Forest Show High Levels of Functional Redundancy. Microb. Ecol. 2016, 73, 123–134. Available online: https://link.springer.com/article/10.1007/s00248-016-0878-6 (accessed on 8 November 2022). [CrossRef]
  115. Heil, M.; Baldwin, I.T. Fitness costs of induced resistance: Emerging experimental support for a slippery concept. Trends Plant Sci. 2002, 7, 61–67. [Google Scholar] [CrossRef]
  116. De Kesel, J.; Conrath, U.; Flors, V.; Luna, E.; Mageroy, M.H.; Mauch-Mani, B.; Pastor, V.; Pozo, M.J.; Pieterse, C.M.J.; Ton, J.; et al. The Induced Resistance Lexicon: Do’s and Don’ts. Trends Plant Sci. 2021, 26, 685–691. [Google Scholar] [CrossRef]
  117. Hammerschmidt, R.; M’Etraux, J.-P.; Van Loon, L. Inducing Resistance: A Summary of Papers Presented at the First International Symposium on Induced Resistance to Plant Diseases, Corfu, May 2000. Eur. J. Plant Pathol. 2001, 107, 1–6. [Google Scholar] [CrossRef]
  118. Heil, M.; Bostock, R.M. Induced Systemic Resistance (ISR) Against Pathogens in the Context of Induced Plant Defences. Ann. Bot. 2002, 89, 503. Available online: https://doi.org/10.1093/aob/mcf076 (accessed on 1 May 2022). [CrossRef] [Green Version]
  119. Martinez-Medina, A.; Flors, V.; Heil, M.; Mauch-Mani, B.; Pieterse, C.M.; 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] [PubMed] [Green Version]
  120. Hilker, M.; Schwachtje, J.; Baier, M.; Balazadeh, S.; Bäurle, I.; Geiselhardt, S.; Hincha, D.K.; Kunze, R.; Mueller-Roeber, B.; Rillig, M.C.; et al. Priming and memory of stress responses in organisms lacking a nervous system. Biol. Rev. 2015, 91, 1118–1133. Available online: https://onlinelibrary.wiley.com/doi/full/10.1111/brv.12215 (accessed on 27 October 2022). [CrossRef] [PubMed] [Green Version]
  121. Rasmann, S.; De Vos, M.; Jander, G. Plant Signaling & Behavior Ecological role of transgenerational resistance against biotic threats. Plant Signal. Behav. 2012, 7, 447–449. Available online: https://www.tandfonline.com/action/journalInformation?journalCode=kpsb20 (accessed on 8 November 2022). [PubMed] [Green Version]
  122. Sen, S. Role of Phytoalexins in Plant-Microbe Interactions and Human Health. Int. J. Sci. Res. Manag. 2017, 5, 18033–18056. [Google Scholar] [CrossRef]
  123. Song, H.; Wang, P.; Li, C.; Han, S.; Lopez-Baltazar, J.; Zhang, X.; Wang, X. Identification of lipoxygenase (LOX) genes from legumes and their responses in wild type and cultivated peanut upon Aspergillus flavus infection. Sci. Rep. 2016, 6, 35245. Available online: https://www.nature.com/articles/srep35245 (accessed on 28 April 2022). [CrossRef] [Green Version]
  124. Chen, Z.; Chen, X.; Yan, H.; Li, W.; Li, Y.; Cai, R.; Xiang, Y. The Lipoxygenase Gene Family in Poplar: Identification, Classification, and Expression in Response to MeJA Treatment. PLoS ONE 2015, 10, e0125526. Available online: https://pubmed.ncbi.nlm.nih.gov/25928711/ (accessed on 1 May 2022). [CrossRef] [Green Version]
  125. Liu, X.M.; Zhang, H. The effects of bacterial volatile emissions on plant abiotic stress tolerance. Front. Plant Sci. 2015, 6, 774. [Google Scholar] [CrossRef]
  126. Sofía, L.; Vanegas, C.; Barrera, C.C.; Bernal, A.J. De la raíz a la hoja: Capacidad de Bacilos Rizosféricos como Potenciales Inductores de Resistencia Sistémica (ISR) en Plantas de Arabidopsis Thaliana. Available online: https://repositorio.uniandes.edu.co/handle/1992/45022 (accessed on 8 November 2022).
  127. Yan, Z.; Reddy, M.S.; Ryu, C.-M.; McInroy, J.A.; Wilson, M.; Kloepper, J.W. Induced Systemic Protection Against Tomato Late Blight Elicited by Plant Growth-Promoting Rhizobacteria. Phytopathology 2002, 92, 1329–1333. Available online: https://apsjournals.apsnet.org/doi/10.1094/PHYTO.2002.92.12.1329 (accessed on 8 November 2022). [CrossRef]
  128. 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. Available online: https://www.nature.com/articles/nchembio.164 (accessed on 9 November 2022). [CrossRef] [Green Version]
  129. Gao, L.; Zhao, W.; Qu, H.; Wang, Q.; Lingxia, Z. The yellow-fruited tomato 1 (yft1) mutant has altered fruit carotenoid accumulation and reduced ethylene production as a result of a genetic lesion in ETHYLENE INSENSITIVE2. Theor. Appl. Genet. 2016, 129, 717–728. Available online: http://faostat.fao.org/ (accessed on 17 December 2022). [CrossRef]
  130. Huang, Y.; Wang, S.; Shi, L.; Xu, F. JASMONATE RESISTANT 1 negatively regulates root growth under boron deficiency in Arabidopsis. J. Exp. Bot. 2021, 72, 3108–3121. Available online: https://pubmed.ncbi.nlm.nih.gov/33530106/ (accessed on 17 December 2022). [CrossRef] [PubMed]
  131. Tungadi, T.; Watt, L.G.; Groen, S.C.; Murphy, A.; Du, Z.; E Pate, A.; Westwood, J.H.; Fennell, T.G.; Powell, G.; Carr, J. Infection of Arabidopsis by cucumber mosaic virus triggers jasmonate-dependent resistance to aphids that relies partly on the pattern-triggered immunity factor BAK1. Mol. Plant Pathol. 2021, 22, 1082–1091. [Google Scholar] [CrossRef] [PubMed]
  132. Lorenzo, O.; Piqueras, R.; Sánchez-Serrano, J.J.; Solano, R. ETHYLENE RESPONSE FACTOR1 Integrates Signals from Ethylene and Jasmonate Pathways in Plant Defense. Plant Cell 2003, 15, 165–178. Available online: www.plantcell.org/cgi/doi/10.1105/tpc.007468 (accessed on 17 December 2022). [CrossRef] [PubMed] [Green Version]
  133. Katsir, L.; Schilmiller, A.L.; Staswick, P.E.; He, S.Y.; Howe, G.A. COI1 is a critical component of a receptor for jasmonate and the bacterial virulence factor coronatine. Proc. Natl. Acad. Sci. USA 2008, 105, 7100–7105. Available online: https://www.pnas.org/doi/abs/10.1073/pnas.0802332105 (accessed on 17 December 2022). [CrossRef] [PubMed] [Green Version]
  134. Staswick, P.E.; Tiryaki, I.; Rowe, M.L. Jasmonate Response Locus JAR1 and Several Related Arabidopsis Genes Encode Enzymes of the Firefly Luciferase Superfamily That Show Activity on Jasmonic, Salicylic, and Indole-3-Acetic Acids in an Assay for Adenylation. Plant Cell 2002, 14, 1405–1415. Available online: https://academic.oup.com/plcell/article/14/6/1405/6009765 (accessed on 17 December 2022). [CrossRef] [Green Version]
  135. Liu, M.; Sun, W.; Ma, Z.; Zheng, T.; Huang, L.; Wu, Q.; Zhao, G.; Tang, Z.; Bu, T.; Li, C.; et al. Genome-wide investigation of the AP2/ERF gene family in tartary buckwheat (Fagopyum Tataricum). BMC Plant Biol. 2019, 19, 84. Available online: https://bmcplantbiol.biomedcentral.com/articles/10.1186/s12870-019-1681-6 (accessed on 17 December 2022). [CrossRef] [Green Version]
  136. Kim, D.S.; Hwang, B.K. An important role of the pepper phenylalanine ammonia-lyase gene (PAL1) in salicylic acid-dependent signalling of the defence response to microbial pathogens. J. Exp. Bot. 2014, 65, 2295–2306. Available online: https://doi.org/10.1093/jxb/eru109 (accessed on 18 December 2022). [CrossRef] [Green Version]
  137. Hayron, M.C.; Jaramillo, M.P.; Belezaca Pinargote, C.; Carranza Patiño, M.; Prieto Benavides, O.; Fernández, R.G. Respuesta de poblaciones microbianas que lideran el crecimiento en raíces y resistencia sistémica inducida. Cienc. Y Tecnol. 2015, 8, 1–11. Available online: https://revistas.uteq.edu.ec/index.php/cyt/article/view/150 (accessed on 18 December 2022).
  138. Valenzuela Ruiz, V.; Gálvez Gamboa, G.T.; Villa Rodríguez, E.D.; Parra Cota, F.I.; Santoyo, G.; De los Santos-Villalobos, S. Lipopéptidos producidos por agentes de control biológico del género Bacillus: Revisión de herramientas analíticas utilizadas para su estudio. Rev. Mex. De Cienc. Agrícolas 2020, 11, 419–432. Available online: http://www.scielo.org.mx/scielo.php?script=sci_arttext&pid=S2007-09342020000200419&lng=es&nrm=iso&tlng=es (accessed on 6 December 2022). [CrossRef]
  139. Holtappels, D.; Fortuna, K.; Lavigne, R.; Wagemans, J. The future of phage biocontrol in integrated plant protection for sustainable crop production. Curr. Opin. Biotechnol. 2021, 68, 60–71. [Google Scholar] [CrossRef]
  140. Botías, C.; Sánchez-Bayo, F. Papel de los plaguicidas en la pérdida de polinizadores. Ecosistemas 2018, 27, 34–41. Available online: https://www.revistaecosistemas.net/index.php/ecosistemas/article/view/1314 (accessed on 6 December 2022). [CrossRef] [Green Version]
  141. Mazari-Hiriart, M.; Pérez-Ortiz, G.; Orta-Ledesma, M.T.; Armas-Vargas, F.; Tapia, M.A.; Solano-Ortiz, R.; Silva, M.A.; Yañez-Noguez, I.; López-Vidal, Y.; Díaz-Ávalos, C. Final opportunity to rehabilitate an urban river as a water source for Mexico City. PLoS ONE 2014, 9, e102081. Available online: https://pubmed.ncbi.nlm.nih.gov/25054805/ (accessed on 31 January 2022). [CrossRef]
  142. De Los Santos-Villalobos, S.; María Díaz-Rodríguez, A.; Fernanda Ávila-Mascareño, M.; Denisse Martínez-Vidales, A.; Parra-Cota, F.I. COLMENA: A Culture Collection of Native Microorganisms for Harnessing the Agro-Biotechnological Potential in Soils and Contributing to Food Security. Diversity 2021, 13, 337. Available online: https://www.mdpi.com/1424-2818/13/8/337/htm (accessed on 31 January 2022). [CrossRef]
  143. Villa-Rodriguez, E.; Moreno-Ulloa, A.; Castro-Longoria, E.; Parra-Cota, F.I.; Santos-Villalobos, S.D.L. Integrated omics approaches for deciphering antifungal metabolites produced by a novel Bacillus species, B. cabrialesii TE3T, against the spot blotch disease of wheat (Triticum turgidum L. subsp. durum). Microbiol. Res. 2021, 251, 126826. Available online: https://pubmed.ncbi.nlm.nih.gov/34298216/ (accessed on 31 January 2022). [CrossRef]
  144. Peteira Delgado-Oramas, B. La resistencia inducida como alternativa para el manejo de plagas en las plantas de cultivo. Rev. Prot. Veg. 2020, 35, e07. Available online: http://scielo.sld.cu/scielo.php?script=sci_arttext&pid=S1010-27522020000100001&lng=es&tlng=es (accessed on 31 January 2022).
  145. Li, Y.; Gu, Y.; Li, J.; Xu, M.; Wei, Q.; Wang, Y. Biocontrol agent Bacillus amyloliquefaciens LJ02 induces systemic resistance against cucurbits powdery mildew. Front. Microbiol. 2015, 6, 883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Mathys, J.; De Cremer, K.; Timmermans, P.; Van Kerckhove, S.; Lievens, B.; Vanhaecke, M.; Cammue, B.P.A.; De Coninck, B. Genome-Wide Characterization of ISR Induced in Arabidopsis thaliana by Trichoderma hamatum T382 Against Botrytis cinerea Infection. Front. Plant Sci. 2012, 3, 108. Available online: www.frontiersin.org (accessed on 1 January 2023). [CrossRef] [PubMed] [Green Version]
  147. Mashabela, M.D.; Piater, L.A.; Dubery, I.A.; Tugizimana, F.; Mhlongo, M.I. Rhizosphere Tripartite Interactions and PGPR-Mediated Metabolic Reprogramming towards ISR and Plant Priming: A Metabolomics Review. Biology 2022, 11, 346. Available online: https://www.mdpi.com/2079-7737/11/3/346/htm (accessed on 22 February 2022). [CrossRef]
Figure 1. Disease triangle showing the interactions needed for the development of biotic stress in plants.
Figure 1. Disease triangle showing the interactions needed for the development of biotic stress in plants.
Stresses 03 00016 g001
Figure 2. Activation of the systemic acquired resistance, through the interaction of plants and pathogens.
Figure 2. Activation of the systemic acquired resistance, through the interaction of plants and pathogens.
Stresses 03 00016 g002
Figure 3. Priming generated by beneficial microorganisms in the rhizosphere, causing activation of resistance which may initially be local, but eventually becomes systemic by jasmonic acid and ethylene.
Figure 3. Priming generated by beneficial microorganisms in the rhizosphere, causing activation of resistance which may initially be local, but eventually becomes systemic by jasmonic acid and ethylene.
Stresses 03 00016 g003
Table 1. Interactions between pathogens and host plants.
Table 1. Interactions between pathogens and host plants.
PathogenHostDiseaseTypeReference
Ustilago maydisCornSmutBiotrophic[33]
Phytophthora capsiciTomato, pepper, other solanaceous, and cucurbit plantsRoot rot, stem necrosis, foliar blight, and fruit rotHemibiotrophic[34]
Botrytis cinereaStrawberry, grape, raspberry, blackberryGray mold, post-harvest rotsNecrotrophic[35]
Bipolaris sorokinianaWheatBlotch spotNecrotrophic[36]
Fusarium oxysporumCucumberFusarium wiltHemibiotrophic[37]
Septoria nodorum BerkWheatGlume blotchHemibiotrophic[38]
Tobacco mosaic virusTobaccoMosaic virus diseaseBiotrophic[39]
Agrobacterium tumefaciensWild grape species (Vitis riparia)Formation of tumors, rod-shaped, crown gall diseaseBiotrophic[40]
Erysiphe orontiiTobaccoPowdery mildew diseasesHemibiotrophic[41]
Fusarium solaniOrange trees (Citrus sinensis L. Osbeck)Darkening of the vascular systemHemibiotrophic[42]
Erwinia amylovoraApple treesFire blight diseaseSemi-necrotrophic or necrotrophic[43]
Table 2. Beneficial microorganisms for plants.
Table 2. Beneficial microorganisms for plants.
Beneficial MicroorganismHostFunctionReference
Bacillus megaterium, Bacillus paralicheniformis, and Bacillus cabrialesiiWheat (Triticum turgidum L. subsp. durum)Growth promotion[81]
Pseudomonas veronii R4Vid (Thompson Seedless)Induced systemic resistance (ISR) defense for leaves and roots[82]
Bacillus velezensis XT1Tomato and strawberry plantGrowth promotion and biocontrol[83]
Bacillus cabrialesiiWheat (Triticum turgidum L. subsp. durum)Biocontrol[84]
Bacillus subtilis Cohn, Bacillus thuringiensis BerlinerWheatPlant resistance[38]
Pseudozyma churashimaensisPepperElicits systemic defense against bacterial and viral pathogens[85]
Pseudomonas fluorescens WCS417rArabidopsis thalianaLocal resistance of roots[86]
Bacillus pumilus EN16, Bacillus subtilis SW1TobaccoSystemic resistance[39]
Trichoderma koningiopsis Th00TomatoInduced systemic resistance (ISR) controlling Fusarium sp.[87]
Gliocladium roseumTobaccoInduced systemic resistance (ISR) controlling Erysiphe orontii[41]
Bacillus sp. FSQ1Common bean (Phaseolus vulgaris)Biological control[88]
Priestia sp. TSO9Wheat (Triticum turgidum L. subsp. durum)Plant growth-promoting[89]
Bacillus paralicheniformis TRQ65Wheat (Triticum turgidum L. subsp. durum)Biological control agent and plant growth-promoting[90]
Bacillus sp. TSO2Wheat (Triticum turgidum L. subsp. durum)Biocontrol[91]
Trichoderma harzianum sensu lato TSM39Wheat (Triticum turgidum L. subsp. durum)Biocontrol[92]
Rouxiella badensis SER3StrawberryBiocontrol[93]
Table 3. Genes involved in plant defense.
Table 3. Genes involved in plant defense.
GeneFunctionSourceReference
CsChi23Antifungal activityCucumber[81]
Npr1AS-dependent regulatory factor to RSAVid (Thompson Seedless)[82]
Eir1Auxin efflux transport, being root-specific, RSI activationVid (Thompson Seedless)[82]
Lox2Lipoxygenase leading to JA biosynthesis, induced systemic resistance (ISR)Vid (Thompson Seedless)[82]
Tlp1Antifungal activityVid (Thompson Seedless)[82]
AtTLP1Encodes a thaumatin-like protein with antimicrobial propertiesArabidopsis thaliana[86]
EIN2Ethylene signaling, chromoplast developmentSolanum lycopersicum (tomato)[129]
JAR1JA signaling in rootArabidopsis thaliana[130]
BAK1Co-receptor enabling detection of microbe-associated molecular patterns and induction of PTIArabidopsis thaliana[131]
ERF1Ethylene and jasmonate pathways in plant defenseArabidopsis thaliana[132]
LOXResistance to biotic and abiotic stressArachis duranensis, Arachis ipaënsis, Cajanus cajan, Cicer arietinum, Glycine max, Lotus japonicus, Medicago truncatula[123]
COI1JA signalingTomato plants[133]
JAR1Encodes a JA–amido synthetase that catalyzes the formation of jasmonoyl-l-isoleucine (JA–Ile)Arabidopsis thaliana[134]
AP2/ERFBiotic and abiotic stresses responsesTartary buckwheat (Fagopyum Tataricum)[135]
PAL1SA-dependent signaling of the defense response to microbial pathogensPepper[136]
R2R3-MYBInitiate the ISR priming processArabidopsis thaliana[137]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

García-Montelongo, A.M.; Montoya-Martínez, A.C.; Morales-Sandoval, P.H.; Parra-Cota, F.I.; de los Santos-Villalobos, S. Beneficial Microorganisms as a Sustainable Alternative for Mitigating Biotic Stresses in Crops. Stresses 2023, 3, 210-228. https://doi.org/10.3390/stresses3010016

AMA Style

García-Montelongo AM, Montoya-Martínez AC, Morales-Sandoval PH, Parra-Cota FI, de los Santos-Villalobos S. Beneficial Microorganisms as a Sustainable Alternative for Mitigating Biotic Stresses in Crops. Stresses. 2023; 3(1):210-228. https://doi.org/10.3390/stresses3010016

Chicago/Turabian Style

García-Montelongo, Ana María, Amelia C. Montoya-Martínez, Pamela Helue Morales-Sandoval, Fannie Isela Parra-Cota, and Sergio de los Santos-Villalobos. 2023. "Beneficial Microorganisms as a Sustainable Alternative for Mitigating Biotic Stresses in Crops" Stresses 3, no. 1: 210-228. https://doi.org/10.3390/stresses3010016

Article Metrics

Back to TopTop