Next Article in Journal
Nitrogen Deficiency-Dependent Abiotic Stress Enhances Carotenoid Production in Indigenous Green Microalga Scenedesmus rubescens KNUA042, for Use as a Potential Resource of High Value Products
Previous Article in Journal
An Epitome of Building Floor Systems by Means of LCA Criteria
Previous Article in Special Issue
Attitude Gaps with Respect to GM Non-Food Crops and GM Food Crops and Confidence in the Government’s Management of Biotechnology: Evidence from Beijing Consumers, Chinese Farmers, Journalists, and Government Officials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Understanding Phytomicrobiome: A Potential Reservoir for Better Crop Management

1
State Key Laboratory for Conservation and Utilization of Subtropical Agro-bioresources, Guangdong Laboratory for Lingnan Modern Agriculture, Integrative Microbiology Research Centre, South China Agricultural University, Guangzhou 510642, China
2
Department of Biochemistry, College of Basic Science and Humanities, SD Agricultural University, Gujarat 385506, India
3
Department of Biotechnology, College of Basic Science and Humanities, SD Agricultural University, Gujarat 385506, India
4
Dum Dum Motijheel College, Microbiology Department, Kolkata, West Bengal 700074, India
5
Department of Chemistry, University of Petroleum and Energy Studies, Dehradun, Uttarakhand 248007, India
6
Department of Chemistry, Lovely Professional University, Phagwara, Punjab 144411, India
*
Authors to whom correspondence should be addressed.
Both authors contributed equally to this manuscript.
Sustainability 2020, 12(13), 5446; https://doi.org/10.3390/su12135446
Submission received: 20 May 2020 / Revised: 22 June 2020 / Accepted: 3 July 2020 / Published: 6 July 2020
(This article belongs to the Special Issue Biotech for Sustainable Agriculture)

Abstract

:
Recent crop production studies have aimed at an increase in the biotic and abiotic tolerance of plant communities, along with increased nutrient availability and crop yields. This can be achieved in various ways, but one of the emerging approaches is to understand the phytomicrobiome structure and associated chemical communications. The phytomicrobiome was characterized with the advent of high-throughput techniques. Its composition and chemical signaling phenomena have been revealed, leading the way for “rhizosphere engineering”. In addition to the above, phytomicrobiome studies have paved the way to best tackling soil contamination with various anthropogenic activities. Agricultural lands have been found to be unbalanced for crop production. Due to the intense application of agricultural chemicals such as herbicides, fungicides, insecticides, fertilizers, etc., which can only be rejuvenated efficiently through detailed studies on the phytomicrobiome component, the phytomicrobiome has recently emerged as a primary plant trait that affects crop production. The phytomicrobiome also acts as an essential modifying factor in plant root exudation and vice versa, resulting in better plant health and crop yield both in terms of quantity and quality. Not only supporting better plant growth, phytomicrobiome members are involved in the degradation of toxic materials, alleviating the stress conditions that adversely affect plant development. Thus, the present review compiles the progress in understanding phytomicrobiome relationships and their application in achieving the goal of sustainable agriculture.

1. Introduction

The phytomicrobiome (PM) can be described as the microbial community associated with a plant that constitutes the whole root as well as the shoot parts. We know that the advent of photosynthesis changed the overall evolutionary fate on earth [1]. On the earth, at present, plants are one of the main entry sources of energy [2]. Therefore, non-photosynthetic organisms including humans remain reliant on plants for their energy needs. Similarly, microorganisms form a complex association with plant parts, so this photo community system, which includes plants and microbes, is comprised of regulated molecular signals that modulate the activity of each other and satisfy mutual benefits [3]. The awareness of plant-associated microbial communities started with research on legume–microbial associations, where an intricate signaling system between the plant and microorganisms in the rhizosphere was found [4]. The rhizospheric region of the plant contains the pool of microbial strains acting as plant growth-promoting rhizobacteria (PGPR). Further research related to deciphering molecular signals and biochemical compounds, which take place in the rhizosphere, has revealed a sophisticated and complex signaling mechanism that helps plants to thrive in their surrounding environment (Table 1).
The microbial community interacts with the plant not only in the rhizospheric region but also in the phyllosphere, which contains a wide array of living organisms such as bacteria, fungi, yeasts, algae, nematodes, etc. It influences the phytophysiology and overall development [18]. The biotic and abiotic conditions regulate the microbiome composition of the phyllosphere [19]. This PM also varies with biotic factors, such as a pathogenic organism causing disease, where the host plant microbiome varies with disease progression [20]. Thus, understanding the phyllosphere microbial community gives an informative overview of plant health. PM analysis requires a holistic study involving the different techniques of genomics, transcriptomics, proteomics, and metabolomics to determine the relationship between the plant and its associated microbial community. An amplified rDNA restriction analysis (ARDRA) is usually utilized for characterizing the bacterial community present in any given sample, which in short involves enzymatic amplification followed by the restriction digestion of 16S rRNA amplified fragments [21]. Advanced techniques like Illumina-based microbiome analysis further helps to analyze the microbial community and their dynamic relationships in much less time [20,22]. Thus, recently, much research has been carried out concerning the PM structure and composition and their variations under different plant conditions.

2. Phytomicrobiome: Its Composition and Biomass

Plants are surrounded by various microbial communities which affect them either positively or negatively. These communities are present around all plant parts, including the roots and shoots, of which the most studied are the below-ground or root-associated microbial communities [23]. The rhizomicrobiome constitutes of root-associated microbes, the phyllomicrobiome constitutes of shoot-associated microbial communities, and the endomicrobiome consists of microbial communities present in the inner plant tissue system. Many studies have investigated the different components of the PM, revealing the wide diversity of microbes which varies from crop to crop [24]. It varies with the developmental stage of the same crop, as well as with the particular developmental stage in the presence of stress conditions [25]. A study with common bean, soybean, and canola related to the effect of biotic and abiotic stress on the phyllosphere composition at three locations in Ontario, Canada, revealed that the phyllosphere community structure changes under seasonal variations and that these community variations are very active [18]. Similarly, da Silva et al. (2014) carried out a study on the endomicrobiomes of different transgenic and nontransgenic maize genotypes. They reported about different factors which controlled the endomicrobiome microbial assemblage and identified these factors’ proteins which were expressed in the transgenic genotypes [26] (Figure 1).
Studies related to PM compositions have revealed that plants are in dynamic relationships with the surrounding microbial communities, which show variations as per the plant developmental stage and their environmental conditions. Either the rhizosphere or phyllosphere or even the endophytic microbes are seen to enable plant establishment in their surroundings. It is very interesting to note that the same plant grown in different soils shows the recruitment of the same microbial communities for their establishment, which strongly reflects the modifications of the soil microbiota by plants. This underlies an intricate pathway of signaling compounds for community assembly, which opens a new avenue of research. The root microbiome of the plant of different species shows variations in community structure, along with the presence of Actinobacteria, Proteobacteria, and Bacteriodetes as conserved phyla in the rhizomicrobiome of these plant species [27]. The plants Arabidopsis thaliana, A. lyrata, A. helleri, and Cardamine hirsute, which were grown in the same soil in the mentioned study, reflect the requirement for the exploration of PM signaling for sustainable cultivation.
Similarly, in phyllosphere studies it has been observed by many workers that there is little variation in the phyllomicrobiome among the similar plant species, even those planted in various geographical areas [28,29]. Recently, these PM-associated microorganisms were found to be crucial in imparting tolerance against biotic and abiotic stress. The role of the rhizosphere communities is well established, but in recent years workers have reported the involvement of various phyllospheric communities in plant stress alleviation. The phyllosphere bacteria of the Bacillus genus have been reported to mitigate drought stress in rice by previous researchers [30]. The phyllosphere isolates were found to be involved in the activation of the plant antioxidant system by an increase in proline-like osmolytes; phenolic content; and the activity of antioxidative enzymes, such as superoxide dismutase, catalase, peroxidase, etc. [31]. Such studies are now easy to conduct with the advent of what we call “next-generation sequencing”, which includes the methods of 454 pyrosequencing, Illumina sequencing, etc. As compared to the traditional techniques, these tools are advantageous and can be used for multidirectional analysis. Further, these techniques are supplemented by metaproteomic analysis, revealing the complex networks of these PM [32]. Thus, we can say that understanding the ecology of the PM is presently an important thrust area of research to achieve sustainability in agriculture in the coming years. Some of the recent PM studies are presented in Table 2.

3. Phytomicrobiome Signaling

Signaling molecules play a crucial role in communication among plants and the microbe [45]. The legume-rhizobium system is one of the widely studied systems for plant-microbe interactions, and many chemical signals are identified from this system [45,46]. The mycorrhizae group of fungi also has similar interactions [47]. The plant partner releases the chemical signals (flavonoid) in the form of root exudates, which makes a suitable condition for the synthesis of the signaling molecules lipo-chitooligosaccharides (LCOs) in microbial strains [48]. The LCOs consist of receptors that have kinase activity and are involved in root hormone profile variations and root nodule development [49,50]. The PM community composition varies as per the signaling by plant parts in response to the environmental conditions around plant [51,52,53,54,55]. The LCOs are found to be involved in plant growth stimulation [55,56,57], the elaboration of the root system [58], the acceleration of flowering and increased fruit yield [59], and stimulating early somatic embryo development [4,50,57,60]. The enhanced germination and growth of seedlings, along with the LCOs’ positive effect on mitotic cell division, suggest an activated meristem activity. Recently, various products based on LCOs have been applied to several million ha of cropland, mainly corn and soybean, each year during seed sowing. The LCO effects are much higher under the presence of stress (salt, drought, cold) as compared to optimum conditions [61] (Figure 2).
Plants mediate the immediate response using two types of receptors present on the outside and inside of the cell, which play an essential role in the recognition of a microbial reaction [62]. The outer receptor of the plant cell is governed by pattern recognition receptors (PRRs), a class of extracellular surface proteins. These receptors are evolutionary conserved and play a role in the recognition of the microbial cells. The activation of these receptors leads to intracellular signaling in plants and the biosynthesis of the molecules for that response. This response helps in the formation of a microbial biofilm in the rhizosphere as a result of the selective nutritional feeding of the beneficial microbial strains [63]. The composition of the root exudates helps the plant to be protected from pathogenic microbial strains [63,64]. The microbial colony created due to the plant PM is carefully maintained for a long time. These types of microbial interactions are particular to the plant root and shoot. These interactions form the channel of communication between the plant and microbes [65]. Many microbial species have been identified that secrete chemical compounds, extracellular enzymes, phytohormones, organic acids, and surface proteins based on the requirement of a particular instance. Examples of compounds recognized via the high-affinity cell surface PRRs of plants that activate an immune response are flagellins and lipopolysaccharides in Pseudomonas [66]. These signals control aspects of microbial as well as host plant activities and the overall community. The activation of the initial immune response of the plants towards the harmful pathogen also occurs by allowing access to beneficial endophytes [67].
Microbe-associated molecular patterns (MAMPs) also regulate immune responses, including antibiotic secretion. They have been shown to down-regulate during the secretion of plant root exudates in plant-associated Bacillus strains better to smooth the progress of root infection [68]. Plant-microbe beneficial interactions are very similar in the process as the pathogens infect the plant in different steps involving various chemical signals [69]. Plant signaling molecules mainly include the primary and secondary metabolites that belong to a variety of root exudates chemicals. Such signaling compounds are elevated in response to stress compounds, such as phytohormones, N-acyl-homoserine lactones (AHLs), phenols, and peptides, and are involved in microbe-to-microbe signaling as well as signaling between microbes and plants [68,69]. The Quorum sensing signaling molecules are produced by plant-associated bacteria and are utilized as signaling molecules for communication and for the regulation of gene expression [70]. AHLs have also been shown to affect the root development of Arabidopsis [71] and are involved in evoking induced systemic resistance (ISR), enabling plants to face various forms of biotic challenges which otherwise become lethal. Similarly, malic acid secretion was reported from Arabidopsis thaliana in response to foliage pathogenesis, which results in the promotion of biofilm formation in the rhizosphere of beneficial microbes [72]. Plants also face signaling materials produced by potential pathogens and responses by various chemical signals released by its activated defence and response systems [73]. Plants also exploit this microbial communication system in the modulation of gene expression in different microbial communities. The LuxR proteins of the bacteria activated by the plant signaling molecules [74]. This phenomenon is similar to the quorum sensing mechanism of the microbial strains. In the rhizospheric region, roots initiated the signaling between plant and microbes. The root produces the ethylene which acts as a signal with dual functions, perceiving biochemical cues and mediating the rhizospheric microbial assembly [75,76].
The root exudates of the plant are controlled by various factors such as time and space, physiology, nutritional status, the mechanical impendence of the plant, and nearby microbial activity in the rhizosphere [54,77,78]. The rhizosphere consists of three zones: the endorhizosphere, which includes the root tissue of the endodermis and cortical layers; the rhizoplane, which represents the root surface; and the ectorhizosphere, which is the soil near the root [79]. The microbial-rich nature of the rhizosphere in comparison to other soil was first proved by Hiltner in 1904 [80]. Similarly, the control of microbial load via root exudation was first predicted by Knudson in 1920 and Lyon and Wilson (1921) [81,82]. Root exudation utilizes most of the photosynthetically fixed carbon, and typically in the case of young seedlings it accounts for 30–40% of the total carbon fixation [83,84]. Root exudates contain majorly of carbon-based compounds along with ionic species, inorganic acids, oxygen, and water [84,85]. The organic component of exudate can be classified as low molecular weight compounds viz. amino acids, organic acids, sugars, phenolics; and an array of secondary metabolites and high high-molecular-weight compounds which comprise different proteins and mucilage compounds [86]. Some exudates are listed in Table 3.
Plant root exudates are involved in interactions of the microbes with the rhizospheric region. The beneficial interactions of the plant root exudates are associated with the mycorrhizae, rhizobia, and soil bacteria involved in positive activities [86,87,88]. There are several reports available on the root exudates and their effect on the phytomicrobiome [89,90]. Rhizospheric PM inter communications represent the above-ground plant structure and microbial associations [73,90]; similarly, pathogen or herbivore attacks above ground can affect rhizospheric microbes. Injuries to the upper parts of plants have been shown to cause the stimulation of signaling compound production in plant roots [69,91]. The microbial community composition of the rhizosphere was observed to be altered in response to the increased photosynthetic rates under elevated CO2 conditions [92,93]. The involvement of high-throughput proteomics and metabolomics in understanding plant-microbe interactions and the signaling thereof can help in the development of effective, economic agricultural practices alleviating the fossil fuel-based crop inputs [94,95,96]. These studies also enable PM engineering for sustainable crop cultivation under presently variable climatic conditions [97,98].

4. Rhizosphere Engineering

Rhizosphere engineering can boost plant health and alter the activity of root-associated microorganisms for the development of sustainable agriculture. The phytomicrobiome belongs to the genus Rhizobia, Pseudomonas, Bacillus, Burkholderia, Azospirillum, Klebsiella, and Gluconacetobacter applied in agricultural purposes [99]. To utilize the potential of the rhizosphere for the development of plants and the associated environment, it is essential to acknowledge the different root exudate molecules and their interactions with the rhizosphere microflora. The understanding of rhizosphere interactions is essential for the creation of sustainable agroecosystems [100]. The manipulation of the plant and its associated microorganisms changes the rhizosphere through the release of root exudation compounds, which positively influence microbial signaling compounds. Root exudates vary with the plant genotype and species. On the other side, microorganisms secrete signaling compounds categorized into phytohormones, extracellular enzymes, organic acids, antibiotics, volatile signals, and quorum-sensing molecules [101]. Several studies have investigated the modification of plant and rhizosphere microflora for the enrichment of the rhizospheric zone for sustainable agriculture (Figure 3).
Plant engineers have developed genetically modified plants to cope with several problems related to soil biotic and abiotic stress. Engineered plants change their root exudation secretion and the pH of the rhizosphere. This change stimulates a beneficial shift in microorganism activity. This change in the rhizosphere stimulates a beneficial shift in microorganism activity. Apart from plant engineering, microbial engineering plays an essential role in agriculture research, especially with PGPR, biological nitrogen fixation, the solubilization of phosphorous and iron, and the modulation of phytohormone and biocontrol activity [102]. Due to the multiple interactive biomes of the rhizosphere, a change in these interactions has the potential to alter plant productivity, health, and soil properties. Therefore, rhizosphere engineering is promising for soil improvement, crop quality, and productivity. Several studies have promoted rhizosphere engineering. In this series, extensive studies have been carried out on the plant and microbe interaction. For instance, specific Pseudomonas spp. of the maize rhizosphere influence the lateral root and root hair for plant growth [103], and Phyllobacterium brassicacearum stimulates the ethylene pathway for the development of root hairs [104]. Classical breeding approaches have developed resistance to significant pathogens, such as Pseudomonas syringae, and the Xanthomonas species. The engineering of PGPR microbes involves the introgression of the chromosome from resistant to susceptible lines, such as substituting the chromosome of the susceptible wheat line S-615 and rescuing from the resistant wheat line Apex chromosome 5B, which results in chromosome substituent lines SA5B that are as resistant as Apex against common rot [103,104,105]. It was noticed that plant density might be playing a role in the existence of the phytomicrobiome. Traditional breeding considers the plant densities as an essential part of crop improvement. Genetic potential for yield and other traits is fully expressed when plants are grown at very low plant densities, and it is significantly suppressed at high plant densities [106]. Studies on the engineering of the plant-microbe interaction for sustainable agriculture are listed in Table 4.

5. Phytomicrobiome and Rhizoremediation

Microbial strains have the potential to degrade organic pollutants from the soil and the rhizospheric region [130,131,132]. These organic pollutants are mainly pesticides, antibiotics, and steroids. Through their metabolism, they can able to degrade toxic man-made chemicals [132,133,134,135]. The rhizospheric region in plants consists of a pool of specific strains that facilitate the ability to degrade toxic pollutants, as well as promote plant growth [136,137,138]. The plant-associated mechanisms involved in the bioremediation using plants are known as phytotransformation, phytoextraction, phytostabilization, and phytovolatilization [139]. The technique is considered “Green Technology”, as it involves better management of soil contaminants along with the surrounding water and air. However, the success of this technology is limited by many constraints, especially the adaptation of the plant species to contaminated soils. The rhizospheric microbes can use organic contaminant as the sole source of nutrients [140]. These strains interacted with plants in a symbiotic manner and reduce a load of pollutants [141]. Rhizoremediation (RR) is related to this and involves the degradation of soil pollutants through the rhizospheric plant-microbe interactions. We know that plant development and its establishment in a particular habitat depends upon a physiologically robust root system. Recently, it had been found that plant roots not only establish the plant in the soil physically but also modify the soil conditions through their exudation profiles, recruiting various microbial species to help the plant accommodate to the habitat and develop better [142]. Thus, the rhizomicrobiome characteristics and composition can be helpful in the prediction of the rhizoremediation efficiency of plants. The rhizomicrobiome community forms a mega consortium under the guidance of the root exudation signaling of plants for sustaining plant growth and development, apart from the modification of physical soil conditions, as well as ecology. The plant microbiome has been shown to have a crucial role to play in nutrient acquisition, disease resistance, and stress tolerance [143]. Contaminated soil represents an ideal stress condition under which the plant undergoes multiple varieties of dynamic plant-microbe signaling interactions to combat unfavorable conditions and transform the condition to better suit the relationship between the plant metabolism and the PM (Figure 4).
Thus, potential investigations of the PM can support rhizoremediation technology for the better management of soil contaminants and the rejuvenation of various agricultural lands which are affected due to industrial activities. Martínez et al. [144] reported on the fungal community assembly of holm oak trees (Quercus ilex) growing in soil contaminated with trace elements after metalliferous tailing spills. It was concluded that the fungal community composition in ectomycorrhiza affects the plant performance in metal-contaminated soils, thus controlling the phytoremediation efficiency [46,141]. Such studies reflect the necessity of exploration of the PM to understand the critical role of the microbial community in managing the plant traits that can be utilized in purposes like rhizoremediation. Plants and associated microbial communities play an essential role in a constructed wetland [145,146,147]. The endophytic bacterial communities of Juncus acutus are able to degrade the pollutant from the soil of wetlands [44]. Phytoremediation capabilities can be modified by bioaugmentation methods, enhancing the accumulation of soil contaminants in plant tissues, thus sustaining the plant growth and development, along with the cleanup of soil pollutants [148,149,150]. PM engineering could improve plant survival in contaminated soil by planting willow plants. Petroleum-contaminated soils are treated with gamma irradiation and different inoculation strategies [42]. With a better understanding of the synergistic relationships of PM, rhizoremediation can be developed as a promising technology to tackle the problem of soils contaminated with a wide variety of contaminants [151,152,153]. It is now a well-established fact that the rhizoremediation of contaminated soil depends on PM functioning, and the more we understand the relationships of the PM, the better we can utilize the PM for rhizoremediation. This not only enables land reclamation but also the better utilization of various plant species for rhizoremediation, as well as the sustainable cultivation of multiple crops without harming the natural conditions.

6. Conclusions

The studies performed by researchers show the synergetic relationship between the microbiomes and environmental factors that are further associated with the developmental stages of the plant. Plants concomitant with microbes need a low energetic cost. Similarly, due to the growth and development of the plant, the microorganisms equally benefit, especially those associated with the roots. To enhance this synergy between plants and microorganisms, the study of PM signals (hormones or exo-hormones) can be beneficial. These advanced omics studies are a useful tool in the management of agro-ecosystems. The reasonable implication of PMs can assist humans in producing sustainable and profitable food materials to meet increasing global demands, along with nailing down the side-effects on the ecosystem and its stakeholders.

Author Contributions

Conceptualization: A.V. and P.B.; data analysis: A.V. and P.B.; writing—original draft preparation: A.V. and P.B.; writing—review and editing: S.V., M.S.A., P.P., H.M., and S.C.; supervision, funding acquisition and project administration: S.C. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Key-Area Research and Development Program of Guangdong Province, China (2018B020206001), China Postdoctoral Science Foundation (2020M672655), and the Guangdong Special Branch Plan for Young Talent with Scientific and Technological Innovation, China (2017TQ04N026).

Acknowledgments

Authors are thankful to the Integrative Microbiology Research Centre, South China Agricultural University, Guangzhou, China for providing the necessary facilities and staff time for completion of this work.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analysis, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Smith, D.L.; Subramanian, S.; Lamont, J.; Bywater-Ekegärd, M. Signaling in the phytomicrobiome: Breadth and potential. Front. Plant Sci. 2015, 6, 709. [Google Scholar] [CrossRef] [Green Version]
  2. Primo, E.D.; Cossovich, S.; Nievas, F.; Bogino, P.; Humm, E.A.; Hirsch, A.M.; Giordano, W. Exopolysaccharide production in Ensifer meliloti laboratory and native strains and their effects on alfalfa inoculation. Arch. Microbiol. 2019, 202, 391–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Bákonyi, N.; Kisvarga, S.; Barna, D.; Tóth, I.O.; El-Ramady, H.; Abdalla, N.; Kovács, S.; Rozbach, M.; Fehér, C.; Elhawat, N.; et al. Chemical traits of fermented alfalfa brown juice: Its implications on physiological, biochemical, anatomical, and growth parameters of Celosia. Agronomy 2020, 10, 247. [Google Scholar] [CrossRef] [Green Version]
  4. Khan, A.L.; Asaf, S.; Abed, R.M.; Ning Chai, Y.; Al-Rawahi, A.N.; Mohanta, T.K.; Schachtman, D.; Al-Harrasi, A.; Al-Rawahi, A.N.; Rawahi, A. Rhizosphere microbiome of arid land medicinal plants and extra cellular enzymes contribute to their abundance. Microorganisms 2020, 8, 213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Herman, M.; Davidson, J.K.; Smart, C.D. Induction of plant defense gene expression by plant activators and Pseudomonas syringae pv. tomato in greenhouse-grown tomatoes. Phytopathology 2008, 98, 1226–1232. [Google Scholar] [CrossRef] [Green Version]
  6. Siddiqui, M.F.; Sakinah, M.; Singh, L.; Zularisam, A. Targeting N-acyl-homoserine-lactones to mitigate membrane biofouling based on quorum sensing using a biofouling reducer. J. Biotechnol. 2012, 161, 190–197. [Google Scholar] [CrossRef] [Green Version]
  7. Audenaert, K.; Pattery, T.; Cornelis, P.; Höfte, M. Induction of systemic resistance to botrytis cinereain tomato by Pseudomonas aeruginosa 7NSK2: Role of salicylic acid, pyochelin, and pyocyanin. Mol. Plant-Microbe Interact. 2002, 15, 1147–1156. [Google Scholar] [CrossRef] [Green Version]
  8. Yasmin, S.; Hafeez, F.Y.; Mirza, M.S.; Rasul, M.; Arshad, H.M.I.; Zubair, M.; Iqbal, M. Biocontrol of bacterial leaf blight of rice and profiling of secondary metabolites produced by rhizospheric Pseudomonas aeruginosa BRp3. Front. Microbiol. 2017, 8, 1985. [Google Scholar] [CrossRef] [Green Version]
  9. De Meyer, G.; Höfte, M. Salicylic acid produced by the rhizobacterium Pseudomonas aeruginosa 7NSK2 induces resistance to leaf infection by Botrytis cinerea on bean. Phytopathology 1997, 87, 588–593. [Google Scholar] [CrossRef] [Green Version]
  10. Munhoz, L.D.; Fonteque, J.P.; Santos, I.M.O.; Navarro, M.O.P.; Simionato, A.; Goya, E.T.; Rezende, M.I.; Balbi-Peña, M.L.; De Oliveira, A.G.; Andrade, G. Control of bacterial stem rot on tomato by extracellular bioactive compounds produced by Pseudomonas aeruginosa LV strain. Cogent Food Agric. 2017, 3, 1282592. [Google Scholar] [CrossRef]
  11. Ma, Z.; Hua, G.K.H.; Ongena, M.; Höfte, M. Role of phenazines and cyclic lipopeptides produced by pseudomonas sp. CMR12a in induced systemic resistance on rice and bean. Environ. Microbiol. Rep. 2016, 8, 896–904. [Google Scholar] [CrossRef] [PubMed]
  12. Arseneault, T.; Goyer, C.; Filion, M. Phenazine production by Pseudomonas sp. LBUM223 contributes to the biological control of potato common scab. Phytopathology 2013, 103, 995–1000. [Google Scholar] [CrossRef] [Green Version]
  13. Weller, D.M.; Mavrodi, D.V.; Van Pelt, J.A.; Pieterse, C.M.; Van Loon, L.C.; Bakker, P.A. 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]
  14. Bán, R.; Baglyas, G.; Virányi, F.; Barna, B.; Posta, K.; Kiss, J.; Körösi, K. The chemical inducer, BTH (benzothiadiazole) and root colonization by mycorrhizal fungi (Glomus spp.) trigger resistance against white rot (Sclerotinia sclerotiorum) in sunflower. Acta Biol. Hung. 2017, 68, 50–59. [Google Scholar] [CrossRef] [Green Version]
  15. Martin, F.; Kohler, A.; Murat, C.; Veneault-Fourrey, C.; Hibbett, D.S. Unearthing the roots of ectomycorrhizal symbioses. Nat. Rev. Microbiol. 2016, 14, 760–773. [Google Scholar] [CrossRef]
  16. León-Martínez, D.G.; Vielle-Calzada, J.-P.; Olalde-Portugal, V. Expression of phenazine biosynthetic genes during the arbuscular mycorrhizal symbiosis of Glomus intraradices. Braz. J. Microbiol. 2012, 43, 716–738. [Google Scholar] [CrossRef] [Green Version]
  17. Akiyama, K.; Matsuzaki, K.-I.; Hayashi, H. Plant sesquiterpenes induce hyphal branching in arbuscular mycorrhizal fungi. Nature 2005, 435, 824–827. [Google Scholar] [CrossRef]
  18. Nuccio, E.E.; Starr, E.; Karaoz, U.; Brodie, E.L.; Zhou, J.; Tringe, S.G.; Malmstrom, R.R.; Woyke, T.; Banfield, J.F.; Firestone, M.K.; et al. Niche differentiation is spatially and temporally regulated in the rhizosphere. ISME J. 2020, 14, 999–1014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Copeland, J.K.; Yuan, L.; Layeghifard, M.; Wang, P.W.; Guttman, D.S. Seasonal community succession of the phyllosphere microbiome. Mol. Plant-Microbe Interact. 2015, 28, 274–285. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Blaustein, R.A.; Lorca, G.L.; Meyer, J.; Gonzalez, C.F.; Teplitski, M. Defining the core citrus leaf- and root-associated microbiota: Factors associated with community structure and implications for managing Huanglongbing (Citrus greening) disease. Appl. Environ. Microbiol. 2017, 83, e00210-17. [Google Scholar] [CrossRef] [Green Version]
  21. Khosravi, H.; Dolatabad, H.K. Identification and molecular characterization of Azotobacter chroococcum and Azotobacter salinestris using ARDRA, REP, ERIC, and BOX. Mol. Biol. Rep. 2019, 47, 307–316. [Google Scholar] [CrossRef] [PubMed]
  22. Tkacz, A.; Pini, F.; Turner, T.R.; Bestion, E.; Simmonds, J.; Howell, P.; Greenland, A.; Cheema, J.; Emms, D.M.; Uauy, C.; et al. Agricultural selection of wheat has been shaped by plant-microbe interactions. Front. Microbiol. 2020, 11, 132. [Google Scholar] [CrossRef] [PubMed]
  23. Turner, T.R.; Ramakrishnan, K.; Walshaw, J.; Heavens, D.; Alston, M.; Swarbreck, D.; Osbourn, A.; Grant, A.; Poole, P.S. Comparative metatranscriptomics reveals kingdom level changes in the rhizosphere microbiome of plants. ISME J. 2013, 7, 2248–2258. [Google Scholar] [CrossRef] [Green Version]
  24. Brisson, V.L.; Schmidt, J.E.; Northen, T.R.; Vogel, J.P.; Gaudin, A.C.M. Impacts of maize domestication and breeding on rhizosphere microbial community recruitment from a nutrient depleted agricultural soil. Sci. Rep. 2019, 9, 15611–15614. [Google Scholar] [CrossRef] [Green Version]
  25. Da Silva, D.A.F.; Cotta, S.R.; Vollú, R.E.; Jurelevicius, D.; Marques, J.M.; Marriel, I.E.; Seldin, L. Endophytic microbial community in two transgenic maize genotypes and in their near-isogenic non-transgenic maize genotype. BMC Microbiol. 2014, 14, 332. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Schlaeppi, K.; Dombrowski, N.; Oter, R.G.; Van Themaat, E.V.L.; Schulze-Lefert, P. Quantitative divergence of the bacterial root microbiota in Arabidopsis thaliana relatives. Proc. Natl. Acad. Sci. USA 2013, 111, 585–592. [Google Scholar] [CrossRef] [Green Version]
  27. Redford, A.J.; Bowers, R.M.; Knight, R.; Linhart, Y.; Fierer, N. The ecology of the phyllosphere: Geographic and phylogenetic variability in the distribution of bacteria on tree leaves. Environ. Microbiol. 2010, 12, 2885–2893. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Maignien, L.; Deforce, E.A.; Chafee, M.E.; Eren, A.M.; Simmons, S.L. Ecological succession and stochastic variation in the assembly of Arabidopsis thaliana phyllosphere communities. MBio 2014, 5, e00682-13. [Google Scholar] [CrossRef] [Green Version]
  29. Lebeis, S.L. The potential for give and take in plant–microbiome relationships. Front. Plant Sci. 2014, 5, 287. [Google Scholar] [CrossRef] [Green Version]
  30. Aswathy, S.K.; Sridar, R.; Sivakumar, U. Mitigation of drought in rice by a phyllosphere bacterium Bacillus altitudinis FD48. Afr. J. Microbiol. Res. 2017, 11, 1614–1625. [Google Scholar] [CrossRef] [Green Version]
  31. He, Z.; Piceno, Y.; Deng, Y.; Xu, M.; Lu, Z.; DeSantis, T.; Andersen, G.; Hobbie, S.E.; Reich, P.B.; Zhou, J. The phylogenic composition and structure of soil microbial communities shifts in response to elevated carbon dioxide. Int. Soc. Microb. Ecol. 2012, 6, 259–272. [Google Scholar]
  32. Compant, S.; Samad, A.; Faist, H.; Sessitsch, A. A review on the plant microbiome: Ecology, functions, and emerging trends in microbial application. J. Adv. Res. 2019, 19, 29–37. [Google Scholar] [CrossRef]
  33. Marasco, R.; Mosqueira, M.J.; Fusi, M.; Ramond, J.-B.; Merlino, G.; Booth, J.M.; Maggs-Kölling, G.; Cowan, D.A.; Daffonchio, D. Rhizosheath microbial community assembly of sympatric desert speargrasses is independent of the plant host. Microbiome 2018, 6, 215. [Google Scholar] [CrossRef] [PubMed]
  34. Yurgel, S.; Douglas, G.M.; DuSault, A.; Percival, D.; Langille, M.G.I. Dissecting community structure in wild blueberry root and soil microbiome. Front. Microbiol. 2018, 9, 1187. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Shakya, M.; Gottel, N.; Castro, H.F.; Yang, Z.K.; Gunter, L.; Labbé, J.; Muchero, W.; Bonito, G.; Vilgalys, R.; Tuskan, G.A.; et al. A multifactor analysis of fungal and bacterial community structure in the root microbiome of mature Populus deltoides trees. PLoS ONE 2013, 8, e76382. [Google Scholar] [CrossRef]
  36. Schreiter, S.; Ding, G.-C.; Heuer, H.; Neumann, G.; Sandmann, M.; Grosch, R.; Kropf, S.; Smalla, K. Effect of the soil type on the microbiome in the rhizosphere of field-grown lettuce. Front. Microbiol. 2014, 5, 144. [Google Scholar] [CrossRef]
  37. Dohrmann, A.B.; Küting, M.; Jünemann, S.; Jaenicke, S.; Schlüter, A.; Tebbe, C.C. Importance of rare taxa for bacterial diversity in the rhizosphere of Bt- and conventional maize varieties. ISME J. 2012, 7, 37–49. [Google Scholar] [CrossRef]
  38. Williams, T.R.; Marco, M.L. Phyllosphere microbiota composition and microbial community transplantation on lettuce plants grown indoors. MBio 2014, 5, 01564. [Google Scholar] [CrossRef] [Green Version]
  39. Dees, M.W.; Lysøe, E.; Nordskog, B.; Brurberg, M.B. Bacterial communities associated with surfaces of leafy greens: Shift in composition and decrease in richness over time. Appl. Environ. Microbiol. 2015, 81, 1530–1539. [Google Scholar] [CrossRef] [Green Version]
  40. Lopez-Velasco, G.; Welbaum, G.; Boyer, R.; Mane, S.; Ponder, M. Changes in spinach phylloepiphytic bacteria communities following minimal processing and refrigerated storage described using pyrosequencing of 16S rRNA amplicons. J. Appl. Microbiol. 2011, 110, 1203–1214. [Google Scholar] [CrossRef]
  41. Agler, M.; Ruhe, J.; Kroll, S.; Morhenn, C.; Kim, S.-T.; Weigel, D.; Kemen, E. Microbial hub taxa link host and abiotic factors to plant microbiome variation. PLoS Biol. 2016, 14, e1002352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Yergeau, E.; Bell, T.H.; Champagne, J.; Maynard, C.; Tardif, S.; Tremblay, J.; Greer, C.W. Transplanting soil microbiomes leads to lasting effects on willow growth, but not on the rhizosphere microbiome. Front. Microbiol. 2015, 6, 921. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Gil-Martínez, M.; López-García, A.; Domínguez, M.T.; Navarro-Fernández, C.M.; Kjøller, R.; Tibbett, M.; Marañón, T. Ectomycorrhizal fungal communities and their functional traits mediate plant–soil interactions in trace element contaminated soils. Front. Plant Sci. 2018, 9, 1682. [Google Scholar] [CrossRef] [PubMed]
  44. Syranidou, E.; Thijs, S.; Avramidou, M.; Weyens, N.; Venieri, D.; Pintelon, I.; Vangronsveld, J.; Kalogerakis, N. Responses of the endophytic bacterial communities of Juncus acutus to pollution with metals, emerging organic pollutants and to bioaugmentation with indigenous strains. Front. Plant Sci. 2018, 9, 1526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Narasimhan, K.; Basheer, C.; Bajic, V.B.; Swarup, S. Enhancement of plant-microbe interactions using a rhizosphere metabolomics-driven approach and its application in the removal of polychlorinated biphenyls1. Plant Physiol. 2003, 132, 146–153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Desbrosses, G.J.; Stougaard, J. Root nodulation: A paradigm for how plant-microbe symbiosis influences host developmental pathways. Cell Host Microbe 2011, 10, 348–358. [Google Scholar] [CrossRef] [Green Version]
  47. Bhatt, P.; Joshi, D.; Kumar, N.; Kumar, N. Recent Trends to Study the Functional Analysis of Mycorrhizosphere. In Mycorrhizosphere and Pedogenesis; Varma, A., Choudhary, D., Eds.; Springer: Singapore, 2019; pp. 181–190. [Google Scholar]
  48. 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]
  49. Gough, C.; Cullimore, J. Lipo-chitooligosaccharide signaling in endosymbiotic plant-microbe interactions. Mol. Plant-Microbe Interact. 2011, 24, 867–878. [Google Scholar] [CrossRef] [Green Version]
  50. Frantzeskasis, L.; Pietro, A.D.; Rep, M.; Schirawski, J.; Wu, C.H.; Panstruga, R. Rapid evolution in plant–microbe interactions—A molecular genomics perspective. New Phytol. 2020, 225, 1134–1142. [Google Scholar] [CrossRef] [Green Version]
  51. Antolín-Llovera, M.; Ried, M.K.; Binder, A.; Parniske, M. Receptor kinase signaling pathways in plant-microbe interactions. Annu. Rev. Phytopathol. 2012, 50, 451–473. [Google Scholar] [CrossRef]
  52. Zamioudis, C.; Mastranesti, P.; Dhonukshe, P.; Blilou, I.; Pieterse, C.M. Unraveling root developmental programs initiated by beneficial Pseudomonas spp. bacteria. Plant Physiol. 2013, 162, 304–318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Boriss, R. Phytostimulation and Biocontrol by the Plant-Associated Bacillus Amyloliquefaciens FZB42: An Update. In Phyto-Microbiome in Stress Regulation. Environmental and Microbial Biotechnology; Kumar, M., Kumar, V., Prasad, R., Eds.; Springer: Singapore, 2020. [Google Scholar]
  54. Chaparro, J.M.; Badri, D.V.; Vivanco, J.M. Rhizosphere microbiome assemblage is affected by plant development. ISME J. 2013, 8, 790–803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Evangelisti, E.; Rey, T.; Schornack, S. Cross-interference of plant development and plant–microbe interactions. Curr. Opin. Plant Biol. 2014, 20, 118–126. [Google Scholar] [CrossRef]
  56. Khan, W.; Prithiviraj, B.; Smith, D.L. Nod factor [Nod Bj V (C18:1, MeFuc)] and lumichrome enhance photosynthesis and growth of corn and soybean. J. Plant Physiol. 2008, 165, 1342–1351. [Google Scholar] [CrossRef] [PubMed]
  57. Wang, N.; Khan, W.; Smith, D.L. Changes in soybean global gene expression after application of lipo-chitooligosaccharide from Bradyrhizobium japonicum under sub-optimal temperature. PLoS ONE 2012, 7, e31571. [Google Scholar] [CrossRef] [Green Version]
  58. 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]
  59. Chen, M.; Wei, H.; Cao, J.; Liu, R.; Wang, Y.; Zheng, C. Expression of Bacillus subtilis proBA genes and reduction of feedback inhibition of proline synthesis increases proline production and confers osmotolerance in transgenic Arabidopsis. J. Biochem. Mol. Biol. 2007, 40, 396–403. [Google Scholar] [CrossRef] [Green Version]
  60. Dyachok, J.; Wiweger, M.; Kenne, L.; von Arnold, S. Endogenous nod factor-like signal molecules promote early somatic embryo development in Norway spruce. Plant Physiol. 2002, 128, 523–533. [Google Scholar] [CrossRef] [PubMed]
  61. Zeng, T.; Rodriguez-Moreno, L.; Mansurkhodzaev, A.; Wang, P.; Berg, W.V.D.; Gasciolli, V.; Cottaz, S.; Fort, S.; Thomma, B.P.H.J.; Bono, J.; et al. A lysin motif effector subverts chitin-triggered immunity to facilitate arbuscular mycorrhizal symbiosis. New Phytol. 2019, 225, 448–460. [Google Scholar] [CrossRef] [Green Version]
  62. Oldroyd, G.E.D.; 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]
  63. Dangl, J.L.; Horvath, D.M.; Staskawicz, B.J. Pivoting the plant immune system from dissection to deployment. Science 2013, 341, 746–751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. De-La-Peña, C.; Loyola-Vargas, V.M. Biotic interactions in the rhizosphere: A diverse cooperative enterprise for plant productivity. Plant Physiol. 2014, 166, 701–719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Engelmoer, D.J.P.; Behm, J.E.; Kiers, E.T. Intense competition between arbuscular mycorrhizal mutualists in anin vitroroot microbiome negatively affects total fungal abundance. Mol. Ecol. 2013, 23, 1584–1593. [Google Scholar] [CrossRef] [PubMed]
  66. Silva, R.J.S.; Micheli, F. RRGPredictor, a set-theory-based tool for predicting pathogen-associated molecular pattern receptors (PRRs) and resistance (R) proteins from plants. Genomics 2020, 112, 2666–2676. [Google Scholar] [CrossRef]
  67. 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] [PubMed] [Green Version]
  68. Lakshmanan, V.; Kitto, S.L.; Caplan, J.L.; Hsueh, Y.H.; Kearns, D.B.; Wu, Y.S.; Bais, H. Microbe-associated molecular patterns-triggered root responses mediate beneficial rhizobacterial recruitment in Arabidopsis. Plant Physiol. 2012, 160, 1642–1661. [Google Scholar] [CrossRef] [Green Version]
  69. Barea, J.M. Future challenges and perspectives for applying microbial biotechnology in sustainable agriculture based on a better understanding of plant-microbe interactions. J. Soil Sci. Plant Nutr. 2015, 15, 261–282. [Google Scholar]
  70. Berendsen, R.L.; Pieterse, C.M.; Bakker, P.A. The rhizosphere microbiome and plant health. Trends Plant Sci. 2012, 17, 478–486. [Google Scholar] [CrossRef]
  71. Ortiz-Castro, R.; Martínez-Trujillo, M.; López-Bucio, J. N-acyl-L-homoserine lactones: A class of bacterial quorum-sensing signals alter post-embryonic root development in Arabidopsis thaliana. Plant Cell Environ. 2008, 31, 1497–1509. [Google Scholar] [CrossRef]
  72. Rudrappa, T.; Czymmek, K.J.; Paré, P.W.; Bais, H. Root-secreted malic acid recruits beneficial soil bacteria. Plant Physiol. 2008, 148, 1547–1556. [Google Scholar] [CrossRef] [Green Version]
  73. 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] [PubMed] [Green Version]
  74. Mosquito, S.; Meng, X.; Devescovi, G.; Bertani, I.; Geller, A.M.; Levy, A.; Myers, M.P.; Bez, C.; Covaceuszach, S.; Venturi, V. LuxR solos in the plant endophyte Kosakonia sp. strain KO348. Appl. Environ. Microbiol. 2020, 86, e00622-20. [Google Scholar] [CrossRef] [PubMed]
  75. Tarkka, M.; Schrey, S.; Hampp, R. Plant associated soil microorganisms. In Molecular Mechanisms of Plant and Microbe Coexistence; Nautiyal, C., Dion, P., Eds.; Springer: Heidelberg, Germany, 2008; pp. 3–51. [Google Scholar]
  76. 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]
  77. Chaparro, J.M.; Badri, D.V.; Bakker, M.G.; Sugiyama, A.; Manter, D.K.; Vivanco, J.M. Root exudation of phytochemicals in Arabidopsis follows specific patterns that are developmentally programmed and correlate with soil microbial functions. PLoS ONE 2013, 8, e55731. [Google Scholar] [CrossRef]
  78. Bengough, A.G.; Mullins, C.E. Mechanical impedance to root growth: A review of experimental techniques and root growth responses. J. Soil Sci. 1990, 41, 341–358. [Google Scholar] [CrossRef]
  79. Fitter, A.H.; Lynch, J.M. The Rhizosphere. J. Appl. Ecol. 1991, 28, 748. [Google Scholar] [CrossRef]
  80. Hiltner, L. Uber neure erfahrungen und probleme auf dem Gebiet der Boden-bakteriologie und unter besondere Berucksichtigung der grundungung und Bracke. Arb. Dtsch. Landwirtsch. Ges. 1904, 98, 59–78. [Google Scholar]
  81. Knudson, L. The secretion of invertase by plant roots. Am. J. Bot. 1920, 7, 371–379. [Google Scholar] [CrossRef]
  82. Lyon, T.L.; Wilson, J.K. Liberation of Organic Matter by Roots of Growing Plants; New York Agricultural Experimental Station Bulletin: New York, NY, USA, 1921. [Google Scholar]
  83. Martin, M.H.; Marschner, H. The Mineral Nutrition of Higher Plants. J. Ecol. 1988, 76, 1250. [Google Scholar] [CrossRef] [Green Version]
  84. Whipps, J.M. Carbon economy. In The Rhizosphere; Lynch, J.M., Ed.; JohnWiley & Sons: Essex, UK, 1990. [Google Scholar]
  85. Uren, N. Types, Amounts, and Possible Functions of Compounds Released into the Rhizosphere by Soil-Grown Plants. In The Rhizosphere; CRC Press: Boca Raton, FL, USA, 2000; pp. 35–56. [Google Scholar]
  86. 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]
  87. Bais, H.P.; Broeckling, C.D.; Vivanco, J.M. Root Exudates Modulate Plant—Microbe Interactions in the Rhizosphere. In Secondary Metabolites in Soil Ecology. Soil Biology; Karlovsky, P., Ed.; Springer: Heidelberg, Germany, 2008; Volume 14, pp. 241–252. [Google Scholar]
  88. Bhatt, K.; Maheshwari, D.K. Decoding multifarious role of cow dung bacteria in mobilization of zinc fractions along with growth promotion of C. annuum L. Sci. Rep. 2019, 9, 14232. [Google Scholar] [CrossRef] [PubMed]
  89. Prithiviraj, B.; Paschke, M.W.; Vivanco, J.M.; Goodman, R.M. Root Communication: The Role of Root Exudates. In Encyclopedia of Plant and Crop Science; Routledge: New York, NY, USA, 2004; pp. 1–5. [Google Scholar]
  90. Segonzac, C.; Zipfel, C. Activation of plant pattern-recognition receptors by bacteria. Curr. Opin. Microbiol. 2011, 14, 54–61. [Google Scholar] [CrossRef] [PubMed]
  91. Nailwal, P.; Tapan, K.; Negi, L.; Panwar, A. Isolation and characterization of rhizobial isolates from the rhizospheric soil of an endangered plant Meizotropis pellita. Asian Microbiol. Biotech. Envion. Sci. 2014, 16, 301–306. [Google Scholar]
  92. Berlec, A. Novel techniques and findings in the study of plant microbiota: Search for plant probiotics. Plant Sci. 2012, 193, 96–102. [Google Scholar] [CrossRef] [PubMed]
  93. He, F.; Sheng, M.; Tang, M. Effects of Rhizophagus irregularis on Photosynthesis and Antioxidative Enzymatic System in Robinia pseudoacacia L. under Drought Stress. Front. Plant Sci. 2017, 8, 576. [Google Scholar] [CrossRef] [Green Version]
  94. Elmore, J.M.; Liu, J.; Smith, B.; Phinney, B.; Coaker, G. Quantitative proteomics reveals dynamic changes in the plasma membrane during arabidopsis immune signaling. Mol. Cell. Proteom. 2012, 11, 1796–1813. [Google Scholar] [CrossRef] [Green Version]
  95. Nguyen, T.H.N.; Brechenmacher, L.; Aldrich, J.T.; Clauss, T.R.; Gritsenko, M.A.; Hixton, K.K.; Libault, M.; Tanaka, K.; Yang, F.; Yao, Q.; et al. Quantitative phosphoproeomic analysis of soybean root hairs inoculated with Bradyrhizobium japonicum. Mol. Cell. Proteom. 2012, 11, 1140–1155. [Google Scholar] [CrossRef] [Green Version]
  96. Rose, C.M.; Venkateshwaran, M.; Volkening, J.D.; Grimsrud, P.A.; Maeda, J.; Bailey, D.J.; Park, K.; Howes-Podoll, M.; Os, D.; Yeun, L.H.; et al. Rapid phosphoproteomic and transcriptomic changes in the rhizobia-legume symbiosis. Mol. Cell. Proteom. 2012, 11, 724–744. [Google Scholar] [CrossRef] [Green Version]
  97. Watrous, J.; Roach, P.; Alexandrov, T.; Heath, B.S.; Yang, J.Y.; Kersten, R.D.; Van Der Voort, M.; Pogliano, K.; Gross, H.; Raaijmakers, J.M.; et al. Mass spectral molecular networking of living microbial colonies. Proc. Natl. Acad. Sci. USA 2012, 109, E1743–E1752. [Google Scholar] [CrossRef] [Green Version]
  98. Zhang, H.; Gao, Z.; Wang, W.J.; Liu, G.F.; Shtykova, E.V.; Xu, J.H.; Li, L.F.; Su, X.D.; Dong, Y. The crystal structure of the MPN domain from the COP9 signalosome subunit CSN6. FEBS Lett. 2012, 586, 1147–1153. [Google Scholar] [CrossRef]
  99. Rahi, P. Phytomicrobiome: A reservoir for sustainable agriculture. In Mining of Microbial Wealth and Metagenomics; Kalia, V., Shouche, Y., Purohit, H., Rahi, P., Eds.; Springer: Singapore, 2017. [Google Scholar]
  100. Kudjordjie, E.N.; Sapkota, R.; Steffensen, S.K.; Fomsgaard, I.S.; Nicolaisen, M. Maize synthesized benzoxazinoids affect the host associated microbiome. Microbiome 2019, 7, 59. [Google Scholar] [CrossRef]
  101. Li, T.; Zhang, J.; Shen, C.; Li, H.; Qiu, L. 1-Aminocyclopropane-1-Carboxylate: A novel and strong chemoattractant for the plant beneficial rhizobacterium Pseudomonas putida UW4. Mol. Plant-Microbe Interact. 2019, 32, 750–759. [Google Scholar] [CrossRef] [PubMed]
  102. Jochum, M.D.; McWilliams, K.L.; Pierson, E.A.; Jo, Y.K. Host-mediated microbiome engineering (HMME) of drought tolerance in the wheat rhizosphere. PLoS ONE 2019, 14, e0225933. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Chu, T.N.; Van Bui, L.; Hoang, M.T.T. Pseudomonas PS01 isolated from maize rhizosphere alters root system architecture and promotes plant growth. Microorganisms 2020, 8, 471. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Guichard, M.; Allain, J.M.; Bianchi, M.W.; Frachisse, J.M. Root Hair Sizer: An algorithm for high throughput recovery of different root hair and root developmental parameters. Plant Methods 2019, 15, 104–113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zhang, J.; Jiang, Y.; Wang, Y.; Guo, Y.; Long, H.; Deng, G.; Chen, Q.; Xuan, P. Molecular markers and cytogenetics to characterize a wheat-Dasypyrum villosum 3V (3D) substitution line conferring resistance to stripe rust. PLoS ONE 2018, 13, e0202033. [Google Scholar] [CrossRef] [Green Version]
  106. Fasoula, V.A. Prognostic breeding: A new paradigm for crop improvement. Plant Breed. Rev. 2013, 37, 297–347. [Google Scholar]
  107. López-Bucio, J.; De La Vega, O.M.; Guevara-García, A.; Herrera-Estrella, L. Enhanced phosphorus uptake in transgenic tobacco plants that overproduce citrate. Nat. Biotechnol. 2000, 18, 450–453. [Google Scholar] [CrossRef]
  108. Tesfaye, M.; Temple, S.J.; Allan, D.L.; Vance, C.P.; Samac, D.A. Overexpression of malate dehydrogenase in transgenic alfalfa enhances organic acid synthesis and confers tolerance to aluminum. Plant Physiol. 2001, 127, 1836–1844. [Google Scholar] [CrossRef]
  109. Yang, H.; Knapp, J.; Koirala, P.; Rajagopal, D.; Peer, W.A.; Silbart, L.K.; Murphy, A.; Gaxiola, R.A. Enhanced phosphorus nutrition in monocots and dicots over-expressing a phosphorus-responsive type IH+-pyrophosphatase. Plant Biotech. J. 2007, 5, 735–745. [Google Scholar] [CrossRef]
  110. Deng, W.; Luo, K.; Li, Z.; Yang, Y.; Hu, N.; Wu, Y. Overexpression of Citrus junos mitochondrial citrate synthase gene in Nicotiana benthamiana confers aluminum tolerance. Planta 2009, 230, 355–365. [Google Scholar] [CrossRef]
  111. Lorito, M.; Woo, S.L.; Fernández, I.G.; Colucci, G.; Harman, G.; Pintor-Toro, J.; Filippone, E.; Muccifora, S.; Lawrence, C.B.; Zoina, A.; et al. Genes from mycoparasitic fungi as a source for improving plant resistance to fungal pathogens. Proc. Natl. Acad. Sci. USA 1998, 95, 7860–7865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Volpi, C.; Janni, M.; Lionetti, V.; Bellincampi, D.; Favaron, F.; D’Ovidio, R. The ectopic expression of a pectin methyl esterase inhibitor increases pectin methyl esterification and limits fungal diseases in wheat. Mol. Plant-Microbe Interact. 2011, 24, 1012–1019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Pogorelko, G.; Lionetti, V.; Fursova, O.; Sundaram, R.M.; Qi, M.; Whitham, S.A.; Bogdanove, A.J.; Bellincampi, D.; Zabotina, O.A. Arabidopsis and brachypodium dis-tachyon transgenic plants expressing Aspergillus nidulans acetylesterases have decreased degree of polysaccharide acetylation and increased resistance to pathogens. Plant Physiol. 2013, 162, 9–23. [Google Scholar] [CrossRef] [Green Version]
  114. Azad, A.K.; Amin, L.; Sidik, N.M. Gene technology for papaya ringspot virus disease management. Sci. World J. 2014, 2014, 768038. [Google Scholar] [CrossRef] [PubMed]
  115. Wang, C.; Knill, E.; Glick, B.R.; Défago, G. Effect of transferring 1-aminocyclopropane-1-carboxylic acid (ACC) deaminase genes into Pseudomonas fluorescens strain CHA0 and its gacA derivative CHA96 on their growth-promoting and disease-suppressive capacities. Can. J. Microbiol. 2000, 46, 898–907. [Google Scholar] [CrossRef]
  116. Sundheim, L.; Poplawsky, A.R.; Ellingboe, A.H. Molecular cloning of two chitinase genes from Serratia marcescens and their expression in Pseudomonas species. Physiol. Mol. Plant Pathol. 1988, 33, 483–491. [Google Scholar] [CrossRef]
  117. Alsanius, B.W.; Hultberg, M.; Englund, J.E. Effect of lacZY marking of the 2, 4-diaceyl phloroglucinol producing Pseudomonas fluorescens strain 5–2/4 on its physiological performance and root colonization ability. Microbial. Res. 2002, 157, 39–45. [Google Scholar] [CrossRef]
  118. Dong, Y.H.; Xu, J.L.; Li, X.Z.; Zhang, L.H. AiiA, an enzyme that inactivates the acylhomoserine lactone quorum sensing signal and attenuates the virulence of Erwinia carotovora. Proc. Natl. Acad. Sci. USA 2000, 97, 3526–3531. [Google Scholar] [CrossRef]
  119. Savka, M.A.; Dessaux, Y.; Oger, P.M.; Rossbach, S. Engineering bacterial competitiveness and persistence in the phytosphere. Mol. Plant-Microbe Interact. 2002, 15, 866–874. [Google Scholar] [CrossRef]
  120. Jeong, J.S.; Kim, Y.S.; Baek, K.H.; Jung, H.; Ha, S.H.; Choi, Y.D.; Kim, M.; Reuzeau, C.; Kim, J.K. Root-specific expression of OsNAC10 improves drought tolerance and grain yield in rice under field drought conditions. Plant Physiol. 2010, 153, 185–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Hao, G.; Pitino, M.; Duan, Y.; Stover, E. Reduced susceptibility to Xanthomonas citriin transgenic citrus expressing the FLS2 receptor from Nicotiana benthamiana. Mol. Plant-Microbe Interact. 2016, 29, 132–142. [Google Scholar] [CrossRef]
  122. Raaijmakers, J.M.; van der Sluis, L.; Bakker, P.A.H.M.; Schippers, B.; Koster, M.; Weisbeek, P.J. Utilization of heterologous siderophores and rhizosphere competence of fluorescent Pseudomonas spp. Can. J. Microbiol. 1995, 41, 126–135. [Google Scholar] [CrossRef]
  123. Delgadillo, J.; Lafuente, A.; Doukkali, B.; Redondo-Gómez, S.; Mateos-Naranjo, E.; Caviedes, M.A.; Pajuelo, E.; Rodríguez-Llorente, I.D. Improving legume nodulation and Cu rhizostabilization using a genetically modified rhizobia. Environ. Technol. 2014, 36, 1237–1245. [Google Scholar] [CrossRef]
  124. Barac, T.; Taghavi, S.; Borremans, B.; Provoost, A.; Oeyen, L.; Colpaert, J.V.; Vangronsveld, J.; van der Lelie, N. Engineered endophytic bacteria improve phytoremediation of water-soluble, volatile, organic pollutants. Nat. Biotechnol. 2004, 22, 583–588. [Google Scholar] [CrossRef]
  125. Chen, J.M.C.; Yang, Y.B.Y.; Schultz, B.; McIver, A. Foliar application of lipo-chitooligosaccharides (Nod factors) to tomato (Lycopersicon esculentum) enhances flowering and fruit production. Can. J. Plant Sci. 2007, 87, 365–372. [Google Scholar]
  126. Nautiyal, C.S.; Srivastava, S.; Chauhan, P.S.; Seem, K.; Mishra, A.; Sopory, S.K. Plant growth-promoting bacteria Bacillus amyloliquefaciens NBRISN13 modulates gene expression profile of leaf and rhizosphere community in rice during salt stress. Plant Physiol. Biochem. 2013, 66, 1–9. [Google Scholar] [CrossRef]
  127. Kim, J.S.; Lee, J.; Seo, S.G.; Lee, C.; Woo, S.Y.; Kim, S.H. Gene expression profile affected by volatiles of new plant growth promoting rhizobacteria, Bacillus subtilis strain JS, in tobacco. Genes Genom. 2015, 37, 387–397. [Google Scholar] [CrossRef]
  128. Oliver, R.P.; Henricot, B.; Segers, G. Cladosporium fulvum, Cause of Leaf Mould of Tomato. In Fungal Pathology; Kronstad, J.W., Ed.; Springer: Dordrecht, The Netherlands, 2000; pp. 65–91. [Google Scholar]
  129. Yuttavanichakul, W.; Lawongsa, P.; Wongkaew, S.; Teaumroong, N.; Boonkerd, N.; Nomura, N.; Tittabutr, P. Improvement of peanut rhizobial inoculant by incorporation of plant growth promoting rhizobacteria (PGPR) as biocontrol against the seed borne fungus, Aspergillus niger. Biol. Control. 2012, 63, 87–97. [Google Scholar] [CrossRef]
  130. Bhatt, P.; Pal, K.; Bhandari, G.; Barh, A. Modelling of the methyl halide biodegradation in bacteria and its effect on environmental systems. Pestic. Biochem. Physiol. 2019, 158, 88–100. [Google Scholar] [CrossRef]
  131. Bhatt, P.; Gangola, S.; Chaudhary, P.; Khati, P.; Kumar, G.; Sharma, A.; Srivastava, A. Pesticide induced up-regulation of esterase and aldehyde dehydrogenase in indigenous Bacillus spp. Bioremediation J. 2019, 23, 42–52. [Google Scholar] [CrossRef]
  132. Bhatt, P.; Huang, Y.; Zhan, H.; Chen, S. Insight into microbial applications for the biodegradation of pyrethroid insecticides. Front. Microbiol. 2019, 10, 1778. [Google Scholar] [CrossRef] [PubMed]
  133. Bhatt, P.; Bhatt, K.; Huang, Y.; Lin, Z.; Chen, S. Esterase is a powerful tool for the biodegradation of pyrethroid insecticides. Chemosphere 2020, 244, 125507. [Google Scholar] [CrossRef] [PubMed]
  134. Bhatt, P.; Huang, Y.; Zhang, W.; Sharma, A.; Chen, S. Enhanced cypermethrin degradation kinetics and metabolic pathway in Bacillus thuringiensis strain SG4. Microorganisms 2020, 8, 223. [Google Scholar] [CrossRef] [Green Version]
  135. Bhatt, P.; Huang, Y.; Rene, E.R.; Kumar, A.J.; Chen, S. Mechanism of allethrin biodegradation by a newly isolated Sphingomonas trueperi strain CW3 from wastewater sludge. Bioresour. Technol. 2020, 305, 123074. [Google Scholar] [CrossRef] [PubMed]
  136. Lin, Z.; Zhang, W.; Pang, S.; Huang, Y.; Mishra, S.; Bhatt, P.; Chen, S. Current approaches to and future perspectives on methomyl degradation in contaminated soil/water environments. Molecules 2020, 25, 738. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Zhang, W.; Lin, Z.; Pang, S.; Bhatt, P.; Chen, S. Insights into the biodegradation of lindane (γ-hexachlorocyclohexane) using a microbial system. Front. Microbiol. 2020, 11, 522. [Google Scholar] [CrossRef] [Green Version]
  138. Mishra, S.; Zhang, W.; Lin, Z.; Pang, S.; Huang, Y.; Bhatt, P.; Chen, S. Carbofuran toxicity and its microbial degradation in contaminated environments. Chemosphere 2020, 259, 127419. [Google Scholar] [CrossRef]
  139. Zhan, H.; Huang, Y.; Lin, Z.; Bhatt, P.; Chen, S. New insights into the microbial degradation and catalytic mechanism of synthetic pyrethroids. Environ. Res. 2020, 182, 109138. [Google Scholar] [CrossRef]
  140. Negi, G.; Pankaj, S.A.; Sharma, A. In situ biodegradation of endosulfan, imidacloprid, and carbendazim using indigenous bacterial cultures of agriculture fields of Uttarakhand, India. Int. J. Biol. Food Vat. Agric. Eng. 2014, 8, 953–961. [Google Scholar]
  141. Bhatt, K.; Maheshwari, D.K. Zinc solubilizing bacteria (Bacillus megaterium) with multifarious plant growth promoting activities alleviates growth in Capsicum annuum L. 3 Biotech 2020, 10, 36. [Google Scholar] [CrossRef]
  142. Singh, H.; Verma, A.; Kumar, M.; Sharma, R.; Gupta, R.; Kaur, M. Phytoremediation: A green technology to clean up the sites with low and moderate level of heavy metals. Austin Biochem. 2017, 2, 1012. [Google Scholar]
  143. Jalil, S.U.; Ansari, M.I. Plant microbiome and its functional mechanism in response to environmental stress. Int. J. Green Pharm. 2018, 12, 81–92. [Google Scholar]
  144. Martínez, M.T.; San-José, M.D.C.; Arrillaga, I.; Cano, V.; Morcillo, M.; Cernadas, M.J.; Corredoira, E. Holm oak somatic embryogenesis: Current status and future perspectives. Front. Plant Sci. 2019, 10, 239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Huang, Y.; Lin, Z.; Zhang, W.; Pang, S.; Bhatt, P.; Rene, E.R.; Kumar, A.J.; Chen, S. New insights into the microbial degradation of D-cyphenothrin in contaminated water/soil environments. Microorganims 2020, 8, 473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Feng, Y.; Huang, Y.; Zhan, H.; Bhatt, P.; Chen, S. An overview of strobilurin fungicide degradation: Current status and future perspective. Front. Microbiol. 2020, 11, 389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Zhan, H.; Feng, Y.; Fan, X.; Chen, S. Recent advances in glyphosate biodegradation. Appl. Microbiol. Biotechnol. 2018, 10, 5033–5043. [Google Scholar] [CrossRef] [PubMed]
  148. Chen, S.; Chang, C.; Deng, Y.; An, S.; Dong, Y.H.; Zhou, J.; Hu, M.; Zhong, G.; Zhang, L.H. Fenpropathrin biodegradation pathway in Bacillus sp. DG-02 and its potential for bioremediation of pyrethroid-contaminated soils. J. Agric. Food Chem. 2014, 62, 2147–2157. [Google Scholar] [CrossRef]
  149. Pang, S.; Lin, Z.; Zhang, W.; Mishra, S.; Bhatt, P.; Chen, S. Insights into the microbial degradation and biochemical mechanisms of neonicotinoids. Front. Microbiol. 2020, 11, 868. [Google Scholar] [CrossRef]
  150. Feng, Y.; Zhang, W.; Pang, S.; Lin, Z.; Zhang, Y.; Huang, Y.; Bhatt, P.; Chen, S. Kinetics and new mechanism of azoxystrobin biodegradation by an Ochrobactrum anthropi strain SH14. Microorganisms 2020, 8, 625. [Google Scholar] [CrossRef]
  151. Huang, Y.; Zhan, H.; Bhatt, P.; Chen, S. Paraquat degradation from contaminated environments: Current achievements and perspectives. Front. Microbiol. 2019, 10, 1754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Chen, S.; Deng, Y.; Chang, C.; Lee, J.; Cheng, Y.; Cui, Z.; Zhou, J.; He, F.; Hu, M.; Zhang, L.H. Pathway and kinetics of cyhalothrin biodegradation by Bacillus thuringiensis strain ZS-19. Sci. Rep. 2015, 5, 8784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Bhatt, P.; Zhang, W.; Lin, Z.; Pang, S.; Huang, Y.; Chen, S. Biodegradation of allethrin by a novel fungus Fusarium proliferatum strain CF2, isolated from contaminated soils. Microorganisms 2020, 8, 593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Factors regulating the microbiome composition.
Figure 1. Factors regulating the microbiome composition.
Sustainability 12 05446 g001
Figure 2. Schematic representation of phytomicrobiome on environment.
Figure 2. Schematic representation of phytomicrobiome on environment.
Sustainability 12 05446 g002
Figure 3. Rhizosphere engineering for improving plant growth promoting activities and balancing soil environment.
Figure 3. Rhizosphere engineering for improving plant growth promoting activities and balancing soil environment.
Sustainability 12 05446 g003
Figure 4. Overview of the phytomicrobiome in the degradation of organic pollutants from soil.
Figure 4. Overview of the phytomicrobiome in the degradation of organic pollutants from soil.
Sustainability 12 05446 g004
Table 1. Signaling and chemical interactions of the rhizosphere.
Table 1. Signaling and chemical interactions of the rhizosphere.
Plant SystemMicroorganismInteraction SignalsTechniques UsedReferences
TomatoPseudomonas syringae pv. tomatoBenzothiadiazoleRT-PCR[5]
Mung bean (Vigna radiata)Agrobacterium TumefaciensN-Acyl-homoserine-lactones (AHLs)RT-PCR
Fluorescence Microscopy
[6]
TomatoPseudomonas aeruginosaPyochelin, its precursor salicylic acid & pyocyaninThin Layer Chromatography (TLC)[7]
RicePseudomonas aeruginosa1-Hydroxy-phenazine, pyocyanin, lahorenoic acid, pyochellin, rhamnolipidsMass spectrometric analysis[8]
BeanPseudomonas aeruginosaPyoverdin, pyochelin, and salicylic acidTLC and colorimetry[9]
TomatoPseudomonas aeruginosaPhenazineTLC and HPLC analysis[10]
RicePseudomonas sp. CMR12Orfamides and sessilins (cyclic lipopeptides); phenazineUPLC-MS[11]
BeanPseudomonas sp. CMR12Orfamides and sessilins (cyclic lipopeptides); phenazineUPLC-MS[11]
PotatoPseudomonas sp. LBUM223 PhenazineqPCR and RT-qPCR analysis[12]
Arabidopsis thalianaPseudomonas fluorescensPolyketide antibiotic 2,4-diacetylphloroglucinolDifferent assays[13]
Sunflower Glomus sp.Benzothiadiazole-[14]
Arabidopsis sp. Laccaria bicolorBenzothiadiazoleDifferent assays[15]
Sorghum sp. and Lolium sp.Glomus intraradicesPhenazineRT-PCR[16]
Lotus japonicusGigaspora margaritaStrigolactoneSpectroscopic analysis[17]
Orobanche weedAM fungisesquiterpene lactonesSpectroscopy[17]
Table 2. Phytomicrobiome ecology: a recent overview of the techniques involved and the communities identified.
Table 2. Phytomicrobiome ecology: a recent overview of the techniques involved and the communities identified.
PhytomicrobiomePlant SystemMethod of StudyCommunity StructureReferences
RhizomicrobiomeSpear grass Illumina sequencing and metaphylogenomic analysisBacterial (Actinobacteria and Alphaproteobacteria) and fungal (Curvularia, Aspergillus, and Thielavia) communities[33]
Root and soil microbiomeWild blueberryIllumina sequencingFungal (Glomeromycota, Mucoromycotina, and Chytridiomycota) and protist communities along with bacterial communities (Aprospirales, Actinomycetales, Rhizobiales, and Xanthomonadales)[34]
Root microbiomePopulus deltoides454 pyrosequencing35 bacterial and 4 fungal taxa in the rhizosphere and 1 bacterial and 1 fungal in the endosphere[35]
Rhizosphere Lettuce (Lactuca sativa) Pyrosequencing Bacterial communities (Sphingomonas, Rhizobium, Pseudomonas, and Variovorax)[36]
Rhizosphere MaizePyrosequencingBacterial diversity (Proteobacteria and Actinobacteria)[37]
PhyllosphereLettuce plants (Lactuca sativa) PyrosequencingBacterial communities (Enterobacteriaceae and Moraxellaceae families)[38]
PhyllosphereRocket salad (Diplotaxis tenuifolia) and lettuce (Lactuca sativa) Illumina sequencingBacterial colonization of leaves (Proteobacteria, Actinobacteria, Firmicutes, and Bacteroidetes)[39]
PhyllosphereSpinach PyrosequencingBacteria communities (Proteobacteria and Firmicutes)[40]
PhyllosphereArabidopsis thalianaIllumina sequencingBacteria (Caulobacter sp., Flavobacterium), fungi (Albugo sp., Dioszegia sp., Udeniomyces sp.) and oomycetes symbionts[41]
RhizosphereWillows (Salix purpurea “Fish Creek”)Illumina sequencingBacterial (Nitrososphaerales, Methanobacteriales, E2 group, Methanosarcinales, and Methanomicrobiales) and fungal (Sordariomycetes, Dothideomycetes, Chytridiomycetes, and Zygomycota) communities[42]
EndosphereHolm oak trees (Quercus ilex subsp. ballota)ITS regionFungal communities (Hebeloma cavipes and Thelephora terrestris)[43]
EndosphereTransgenic maizeDGGE analysis Endophytic communities (bacteria, archaea, and fungi)[44]
PhyllosphereBean, soybean, and canolaIllumina sequencingBacterial communities (Firmicutes, Bacteroidetes, Thermi, and Chloroflexi)[18]
PhyllosphereArabidopsis thaliana454 sequencingBacterial communities (Acinetobacter, Variovorax, Pseudomonas, unidentified Sphingobacteriaceae, Rhodococcus, Ochrobactrum, and Chryseobacterium)[28]
Table 3. Classes of compounds released in plant root exudates [45]. Class of compounds.
Table 3. Classes of compounds released in plant root exudates [45]. Class of compounds.
Different Chemical Components of Plant Root ExudateMicrobial Diversity
CarbohydratesArabinose, glucose, galactose, fructose, sucrose, pentose, rhamnose, raffinose, ribose, xylose and mannitolBacterial species like Actinobacteria, Proteobacteria and Firmicutes; Fungal species like Pythium
Amino acidsAll 20 proteinogenic amino acids, l-hydroxyproline, homoserine, mugineic acid, aminobutyric acidBacterial species like Proteobacteria, Acidobacteria, Actinobacteria; Fungal families like Gigasporaceae, Acaulosporaceae
Organic acidsAcetic acid, succinic acid, l-aspartic acid, malic acid, l-glutamic acid, salicylic acid, shikimic acid, isocitric acid, chorismic acid, sinapic acid, caffeic acid, p-hydroxybenzoic acid, gallic acid, tartaric acid, ferulic acid, protocatacheuic acid, p-coumaric acid, mugineic acid, oxalic acid, citric acid, piscidic acidBacterial species like Actinobacteria, Gemmatimonadetes, and Chloroflexi and fungal species like Rhizophagus intraradices, Funneliformis mosseae
FlavonolsNaringenin, kaempferol, quercitin, myricetin, naringin, rutin, genistein, strigolactone and their substitutes with sugarsBacterial species like Actinobacteria, Proteobacteria and fungal species like Funneliformis mosseae, and Rhizophagus irregularis
LigninsCatechol, benzoic acid, nicotinic acid, phloroglucinol, cinnamic acid, gallic acid, ferulic acid, syringic acid, sinapoyl aldehyde, chlorogenic acid, coumaric acid, vanillin, sinapyl alcohol, quinic acid, pyroglutamic acidBacterial species like Bacillus, Flavisolibacter, Actinobacteria, and fungal species like Rhizoctonia solani, and Scletorina sclerotium.
CoumarinsUmbelliferoneBacterial species like Lysobacter, Phormidium, Proteobacteria and fungal families like Ascomycetes, Scutellosporaceae
AuronesBenzyl aurones synapates, sinapoyl cholineBacterial species like Acidobacteria, Actinobacteria and fungal families like Basidiomycetes, Acaulosporaceae
GlucosinolatesCyclobrassinone, desuphoguconapin, desulphoprogoitrin, desulphonapoleiferin, desulphoglucoalyssinBacterial species like Actinobacteria, Proteobacteria and fungal species like Alternaria solani, Rhizophagus intraradices, Funneliformis mosseae.
AnthocyaninsCyanidin, delphinidin, pelargonidin and their substitutes with sugar moleculesBacterial species like Acidobacteria, Actinobacteria and fungal species like Fusarium equiseti, Rhizophagus intraradices, Rhizophagus irregularis
Indole compoundsIndole-3-acetic acid, brassitin, sinalexin, brassilexin, methyl indole carboxylate, camalexin glucosideBacterial species like Kaistobacter, Flavisolibacter, Actinobacteria and fungal species like Rhizoctonia solani
Fatty acidsLinoleic acid, oleic acid, palmitic acid, stearic acidBacterial species like Actinobacteria, Lysobacter, Balneimonas and fungal species like Funneliformis mosseae, Fusarium equiseti, Alternaria solani.
SterolsCampestrol, sitosterol, stigmasterolBacterial species like Flavisolibacter, Balneimonas and fungal species like Rhizoctonia solani, Rhizophagus irregularis
AllomonesJugulone, sorgoleone, 5,7,4′-trihydroxy-3′, 5′-dimethoxyflavone, DIMBOA, DIBOABacterial species like Gemmatimonadetes, Chloroflexi and fungal species like Alternaria solani, Verticillium sp.
Proteins and enzymesPR proteins, lectins, proteases, acid phosphatases, peroxidases, hydrolases, lipaseBacterial species like Balneimonas, Lysobacter, Actinobacteria and fungal families like Zygomycetes, Gigasporaceae
Table 4. Phytomicrobiome engineering: engineered plant/microbes for rhizosphere enrichment.
Table 4. Phytomicrobiome engineering: engineered plant/microbes for rhizosphere enrichment.
Engineered Plant/Microbes GeneHost EffectReferences
Tobacco rootsCitrate synthase genePseudomonas aeruginosaPhosphorous acquisition, Al tolerance [107]
Medicago sativa Malate dehydrogenaseMedicago sativaEnhanced the organic anion efflux [108]
Arabidopsis thalianaPyrophosphates gene Arabidopsis thalianaEnhanced the tolerant capacity of Al[109]
Nicotiana benthamiana Citrate synthase geneYuzu treeEnhanced the tolerant capacity of Al[110]
Transgenic tobacco, potato Trichoderma harzianum endochitinase -Enhanced the tolerate capacity from fungal pathogen (Alternaria alternate, Botrytis cinerea)[111]
Durum wheat transgenic linesPectin methyl esterase inhibitor geneGolden kiwi treeEnhanced the tolerate capacity from fungal pathogen (Fusarium graminearum, Bipolaris sorokiniana)[112]
Arabidopsis, purple false bromeAspergillus nidulans acetyl esterases -Enhanced the tolerate capacity from fungal pathogen (Botrytis cinere, Bipolaris sorokiniana)[113]
Papaya Papaya ring spot coat protein genePapayaVirus resistant plants[114]
Cucumber, Canola Pseudomonas fluorescens (CHA 0) transformed with ACC deaminase gene acdS from P. putida UW4-Improved root architect and plant protection[115]
Pseudomonas strain Chi A geneSerratia macrcescensEnhanced protection from fungal pathogen [116]
P. fluorescens 5–2/4DAPG biosynthesis operonPseudomonas fluorescens Q2–87Protection from plant pathogen P. ultimum[117]
Potato Bacterial lactonase gene aii ABacillus sp.Protection from plant pathogen Pectobacterium [118]
Lotus corniulatusOpines biosynthesis geneAgrobacterium tumefaciencePhytoremediation[119]
Rice OsNac10 geneRiceEnhanced drought tolerance and increased grain yield [120]
Citrus sweet orangePattern recognition receptor FLS 2Tobacco (Nicotiana benthamiana)Increased canker resistant and defence[121]
RadishHeterologous gene encoding siderophore responsible for iron uptakePseudomonas fluorescensEnhanced the competitiveness in soil [122]
Ensifer medicae straincopAB genes Pseudomonas fluorescensImproved the tolerance of plant in copper contaminated soil and enhanced nodulation[123]
Yellow lupin pTOM toluene-degradation plasmidBurkholderia cepacia G4Participated in phytoremidiation [124]
ArabidopsisproBA geneBacillus subtilisSalt tolerant [125]
Rice, maizeSN13, SQR9Bacillus amyloliquefaciensSalt tolerant[126]
Tobacco VOCs related geneBacillus subtilisEnhancement of plant growth [127]
Cultivar
Tomato
Cf-4
(Fungal gene)
Wild tomatoResistance to the fungal tomato pathogen Cladosporium fulvum[128]
Aspergillus nigerα-GalactosidaseAspergillus nigerIncreased the protein secretion 9 times[129]

Share and Cite

MDPI and ACS Style

Bhatt, P.; Verma, A.; Verma, S.; Anwar, M.S.; Prasher, P.; Mudila, H.; Chen, S. Understanding Phytomicrobiome: A Potential Reservoir for Better Crop Management. Sustainability 2020, 12, 5446. https://doi.org/10.3390/su12135446

AMA Style

Bhatt P, Verma A, Verma S, Anwar MS, Prasher P, Mudila H, Chen S. Understanding Phytomicrobiome: A Potential Reservoir for Better Crop Management. Sustainability. 2020; 12(13):5446. https://doi.org/10.3390/su12135446

Chicago/Turabian Style

Bhatt, Pankaj, Amit Verma, Shulbhi Verma, Md. Shahbaz Anwar, Parteek Prasher, Harish Mudila, and Shaohua Chen. 2020. "Understanding Phytomicrobiome: A Potential Reservoir for Better Crop Management" Sustainability 12, no. 13: 5446. https://doi.org/10.3390/su12135446

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