Next Article in Journal
Pepper Mild Mottle Virus as a Potential Indicator of Fecal Contamination in Influents of Wastewater Treatment Plants in Riyadh, Saudi Arabia
Next Article in Special Issue
The Exploitation of Microbial Antagonists against Postharvest Plant Pathogens
Previous Article in Journal
Gut Microbiome in Post-COVID-19 Patients Is Linked to Immune and Cardiovascular Health Status but Not COVID-19 Severity
Previous Article in Special Issue
Modes of Action of Biocontrol Agents and Elicitors for sustainable Protection against Bacterial Canker of Tomato
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Adaption of Pseudomonas ogarae F113 to the Rhizosphere Environment—The AmrZ-FleQ Hub

1
Departamento de Biología, Facultad de Ciencias, Universidad Autónoma de Madrid, Darwin 2, 28049 Madrid, Spain
2
Department of Fundamental Microbiology, University of Lausanne, 1015 Lausanne, Switzerland
*
Author to whom correspondence should be addressed.
Microorganisms 2023, 11(4), 1037; https://doi.org/10.3390/microorganisms11041037
Submission received: 20 March 2023 / Revised: 10 April 2023 / Accepted: 13 April 2023 / Published: 15 April 2023
(This article belongs to the Special Issue Latest Review Papers in Plant Microbe Interactions 2023)

Abstract

:
Motility and biofilm formation are two crucial traits in the process of rhizosphere colonization by pseudomonads. The regulation of both traits requires a complex signaling network that is coordinated by the AmrZ-FleQ hub. In this review, we describe the role of this hub in the adaption to the rhizosphere. The study of the direct regulon of AmrZ and the phenotypic analyses of an amrZ mutant in Pseudomonas ogarae F113 has shown that this protein plays a crucial role in the regulation of several cellular functions, including motility, biofilm formation, iron homeostasis, and bis-(3′-5′)-cyclic dimeric guanosine monophosphate (c-di-GMP) turnover, controlling the synthesis of extracellular matrix components. On the other hand, FleQ is the master regulator of flagellar synthesis in P. ogarae F113 and other pseudomonads, but its implication in the regulation of multiple traits related with environmental adaption has been shown. Genomic scale studies (ChIP-Seq and RNA-Seq) have shown that in P. ogarae F113, AmrZ and FleQ are general transcription factors that regulate multiple traits. It has also been shown that there is a common regulon shared by the two transcription factors. Moreover, these studies have shown that AmrZ and FleQ form a regulatory hub that inversely regulate traits such as motility, extracellular matrix component production, and iron homeostasis. The messenger molecule c-di-GMP plays an essential role in this hub since its production is regulated by AmrZ and it is sensed by FleQ and required for its regulatory role. This regulatory hub is functional both in culture and in the rhizosphere, indicating that the AmrZ-FleQ hub is a main player of P. ogarae F113 adaption to the rhizosphere environment.

1. The Rhizosphere and the Plant Growth Promoting Rhizobacteria (PGPR)

The term rhizosphere was first described by Lorenz Hiltner in 1904, who proposed that the area surrounding plant roots is a region with a high microbial activity which is shaped by the chemicals released from the roots [1]. Over the years, this first explanation has been revised to cover all the complexity of this environment, as the rhizosphere is not a sizeable entity and consists of biological, chemical, and physical components that can vary between root systems. Therefore, the rhizosphere is considered the part of the soil influenced by plant roots where soil, soil biota, and plant roots interact [2]. The rhizosphere biota includes fungi, bacteria, protists, nematodes, and invertebrates. The biomass and its associated activity are higher in the rhizosphere than in bulk soils [3]. Indeed, the rhizosphere is estimated to contain nearly 107 to 108 bacterial colony-forming units (CFUs) per gram of soil, being two orders of magnitude higher than in the surrounding soil [2,4]. Although the rhizosphere embraces higher biomass than bulk soil, several ecological and evolutionary processes result in a reduction of microbial diversity in the soil under the influence of the root [4,5]. These are the plant geographical distribution, variations in abiotic and biotic environmental factors, selection by the plant and environmental factors, and ecological drift that can both alter the abundance of the existent microorganisms, and also dispersal and evolutionary change that lead to the arise of new species [5]. The strong selection exerted by the plant on its root microbiome is mainly due to the secretion of a variety of metabolites known as exudates. Plant-root exudates are commonly known as rhizodeposition products and can include organic and inorganic compounds, being the organic component relevant for the rhizosphere processes as they can be used as energy sources by microorganisms [6]. Rhizodeposition products are of high energy cost for plants, but they can derive great benefits as they have the property to stimulate or inhibit microbial populations and their activities [7]. For instance, the function of these compounds can be related to nutrient acquisition, allelopathy, attraction of symbiotic partners such as rhizobia and legumes, or promotion of beneficial microorganisms that can, in turn, prevent the presence of pathogens or directly promote plant health such as Bacillus subtilis or Pseudomonas species [8,9]. In this sense, soil microorganisms have an important impact on plant health and productivity [10]. For example, it is known that plants can recruit beneficial microorganisms that promote plant growth when previous pathogens harm them, to increase the chances of survival [11]. The rhizosphere is inhabited by many microorganisms, among which bacteria are known as rhizobacteria [12]. Plant growth-promoting rhizobacteria (PGPR) are beneficial bacteria that can colonize the endorhizosphere, ectorhizosphere, and the rhizoplane [13,14], being Bacillus [15] and Pseudomonas [16] the most studied PGPR genera. A PGPR can boost plant growth in different ways that can be typically classified in biofertilization, phytostimulation, or biocontrol mechanisms depending on whether the activity has a positive direct impact on plant growth or indirectly by limiting the presence of potential pathogens [17,18]. The most common mechanisms are schematized in Figure 1.
PGPRs using direct mechanisms related to the nutrition of the plant are known as biofertilizers [18]. This process implies the facilitation of nutrient uptake from the environment. For instance, increasing iron availability for the plant due to siderophore production by certain bacteria or phosphate solubilization due to the secretion of organic acids or phosphatases [19]. Direct mechanisms used by PGPRs can also impact the hormonal balance of the plant, which can ultimately improve plant fitness. In this case, PGPRs are known as phytostimulants [18]. Phytostimulation is manifested by the synthesis of compounds, mainly phytohormones (e.g., ethylene, gibberellins, auxins, and cytokinins) [2], exopolysaccharides (EPSs) [20], or compatible solutes (e.g., glycine, proline, betaine, or trehalose) that can help the plant cope with abiotic stress and enhance plant growth [21]. A noteworthy example is the PGPR production of 1-aminocyclopropane-1-carboxylate (ACC) deaminase. The enzyme ACC deaminase can regulate plant ethylene levels in the plant in response to stress as it hydrolyzes the ACC (precursor of the phytohormone ethylene) into ammonia and α-ketobutyrate [22]. Biocontrol takes place when PGPRs diminish or avoid the deleterious effects of phytopathogens [17,23]. It involves the synthesis of antagonistic substances such as antibiotics (e.g., hydrogen cyanide, or 2,4-diacetylphloroglucinol (DAPG)) [16], bacteriocins [24], hydrolytic enzymes that can lyse pathogenic fungal cells [25], siderophores that can reduce the presence of pathogens by decreasing iron availability [16,26], direct competition for nutrients and niches with pathogens [27], or by triggering the plant-induced systemic resistance (ISR) to pathogens that make uninfected parts of the plant more resistant to pathogens as it is primed for accelerated activation of defense [28,29].

2. Pseudomonas as PGPR

The genus Pseudomonas comprises Gram-negative, γ-proteobacteria and, as previously mentioned, it is among the most used PGPR in agriculture. Furthermore, it is one of the most diverse bacterial genera, comprising more than 200 recognized species [30,31,32,33]. Members of this genus are known for having a versatile metabolism [30,34] and produce a variety of secondary metabolites [35]. All of the above make Pseudomonas a ubiquitous genus found inhabiting extremely different environments, including aquatic environments, soils, the rhizosphere, and associated with several eukaryotic hosts, including plants but also animals [36]. For example, the phytopathogen Pseudomonas syringae has been linked with the water cycle as it has been found in clouds, rain, snow, lakes, and plants [37]. Pseudomonas species include the known human pathogen P. aeruginosa [38], and plant pathogens such as P. syringae [39]. Other species are also relevant as PGPR, including P. fluorescens, P. brassicacearum, P. protegens, and P. chlororaphis [40], and certain strains also have applications in bioremediation [41] and industry, for the production of relevant compounds [42,43]. The Pseudomonas genus has been divided into different groups based on multilocus sequence analysis (MLSA), genome-to-genome blast distance phylogeny (GBDP), and other phylogenomic and comparative genomic approaches [40,44,45]. Figure 2 shows a phylogenetic tree of the genus Pseudomonas, indicating the major groups. In this genus, there are two major lineages, P. aeruginosa and P. fluorescens. Within the P. fluorescens lineage, three large groups, also referred to as complex of species, have been identified: P. fluorescens, P. syringae, and P. putida. In turn, the P. fluorescens complex of species has been subdivided into several subgroups: P. fluorescens, P. gesardii, P. fragi, P. mandelii, P. jessenii, P. koreensis, P. chlororaphis, P. protegens, and P. corrugata [40,46]. Within the Pseudomonas genus, the P. fluorescens complex of species [40] includes species of chemoheterotrophs and motile bacteria by means of polar flagella. They are aerobes, but some can utilize nitrate as the final electron acceptor and most of the strains can use it as a nitrogen source [47]. Certain subgroups in the P. fluorescens complex such as P. fluorescens, P. koreensis, P. brassicacearum, P. protegens, and P. chlororaphis have been typically related with PGPR traits mainly due to their ability to suppress plant diseases [16,40,47,48,49,50,51] but also for directly stimulating plant growth. For instance, they can boost plant nutrition through inorganic phosphate solubilization [52]; increase iron availability due to siderophore production when this element is scarce [53], being the pyoverdine the most common, a yellow-green fluorescent pigment that gave these bacteria the name [54]; enhance root development through the synthesis of phytohormone-like compounds such as auxins [55]; modification of the plant hormonal balance by the production of ACC deaminase [56], and production of antibiotics and antifungals such as DAPG [48,57] or hydrogen cyanide [58].
Pseudomonas ogarae F113 [59] (henceforth F113), was isolated from the sugar beet (Beta vulgaris) rhizosphere in Ireland [60]. It is able to colonize a wide variety of staple plants such as tomato (Lycopersicum esculentum), potato (Solanum tuberosum), pea (Pisum sativum), alfalfa (Medicago sativa), wheat (Triticum aestivum), strawberry (Fragaria vesca), maize (Zea mays), the model plant Arabidopsis thaliana, and willow trees [57,61,62,63,64,65,66,67,68,69,70,71,72,73], and it is considered a model for rhizosphere colonization [65,70,74]. F113 genome consists of a single circular chromosome and expands over 6.8 Mbp, with an average GC content of 60.8%, and 5862 protein-coding genes, and belongs to the P. corrugata subgroup, together with its closest relatives P. brassicacearum and P. kilonensis [75]. The genome of F113 encodes several features related to its PGP ability, including ACC deaminase, the siderophore pyoverdine that increases iron solubility and limits the proliferation of other microorganisms [16,26], secondary metabolites such as hydrogen cyanide or DAPG, and a large number of secretory systems [75,76], important for inter-bacterial competition [77]. Furthermore, the genome of this bacterium contains the gene gcd encoding a glucose dehydrogenase (Gdc), and the pqqE and pqqB genes which are part of the cluster encoding the pyrroloquinoline carrier and biosynthetic proteins, respectively. Gdc and Pqq were shown to be involved in the solubilization of Ca3(PO4)2 that can increase phosphate bioavailability for the plant [52]. Moreover, the F113 genome encodes around 50 proteins involved in denitrification, being able to grow anaerobically using nitrate and nitrite as electron acceptors [76,78], a process that has been associated with rhizosphere competence in P. fluorescens rhizosphere isolates [79]. F113 biocontrol properties are mostly facilitated by DAPG production [60] that have been shown to protect against several phytopathogens, such as Pythium ultimum-mediated damping-off of sugar beet [80,81], Pectobacterium caratovorum-mediated soft rot of potato [62], the potato cyst nematode Globodera rostochiensis [62], Fusarium oxysporum, the cause of Fusarium wilt in sugar beet [80] and tomato [71], and Phytophthora cactorum, the cause of root rot in strawberry [71].

3. Bacterial Lifestyles and Rhizosphere Colonization: Traits Involved in Rhizosphere Colonization

Bacteria can live in a free-swimming state, known as planktonic, or they can be sessile and adhere to surfaces. Planktonic cells can be found in water films in different environments remaining in suspension like colloidal particles. They can also be transported over considerable distances due to water currents induced by fluid dynamic forces, or they can actively move using bacterial appendages such as flagella. In response to certain stimuli, bacteria can switch from a planktonic state to a sessile one. There are different physicochemical mechanisms known to mediate this lifestyle switch: deposition of bacterial cells when they are transported near a solid surface due to lift and frictional forces, sedimentation of cell–cell or cell–particle aggregates, chemotactic responses towards the surfaces due to a nutrient gradient, Brownian motion of the bacteria close to a surface, long-range forces such as attraction or repulsion to the solid–liquid interface or charged substratum surfaces, short-range forces by the linking of extracellular polymers in the cell surface to a substratum surface, and thermodynamics processes [82]. As a consequence of this transition, cells experience physiological and phenotypic changes [83], such as the slowdown of certain metabolic activities [84] and the production of an extracellular matrix [85]. The transition is not strictly homogeneous, and there are phenotypic variation events that lead to subpopulations with different life states [83].
In the rhizosphere environment, PGPRs must compete with the rest of the microorganisms for nutrients, rhizodeposits, and space [86]. Most of the PGPR activities, such as antibiosis, ISR, or niche exclusion, depend on the bacterial ability to colonize and persist in the root system [27,87,88]. Therefore, increasing the knowledge about the mechanisms underlying a successful colonization, competitiveness with other indigenous populations, and persistence in the rhizosphere is essential for PGPR application. Likewise, bacterial competence to colonize and survive in the rhizosphere relies mainly on motility, chemotaxis, attachment, growth, and stress resistance [89]. Specifically, in Pseudomonas, there are multiple factors linked with rhizosphere colonization. Many of them are common to all pseudomonads, such as motility, secretion systems, metabolic adaption, nutrient uptake, EPS synthesis, and biofilm formation [50,86]. However, different bacteria can use very distinct strategies for rhizosphere colonization.
Regarding F113, the fact that non-motile or reduced-motility mutants are impaired in competitive colonization of the rhizosphere [61,90] and that hypermotile derivatives can be isolated from the rhizosphere [74], highlights the importance of motility in the process. This bacterium does not form typical mature biofilms in the rhizoplane but rather microcolonies [65,70]. Furthermore, certain mutants impaired in biofilm formation on abiotic surfaces do not display a deficiency in rhizosphere competitive colonization [70]. Another essential determinant of rhizosphere colonization in Pseudomonas is the phase variation process, which implies a genetic diversification of subpopulations that allows drastic phenotypic variations via small genetic changes that enhance bacterial adaption [91]. Phase variation has been shown as an important trait for rhizophere colonization by F113 [92,93].
In the last decade, the use of omics has greatly increased the knowledge about the important determinants during rhizosphere colonization [94,95,96]. Attempts to set the gene map or functional categories associated with root colonization revealed the importance of motility, defense, fimbrial low-molecular weight protein (Flp) pilus assembly, iron homeostasis, Type Three Secretion Systems and Type Six Secretion Systems (T6SSs) as major determinants in addition to the categories mentioned above [97,98,99]. Additionally, Arruda et al. (2019) created a synthetic microbial community with the ability to promote growth in maize plants to elucidate traits associated with successful colonization [100,101]. This work demonstrated that the typical PGPR traits did not appear as determinants of robust colonization, whereas the ability to acquire and transport nutrients such as sugars, organic acids, and amino acids, and produce EPSs were shown as essential for the successful colonization of plants.

4. Motility

Motility provides an advantage to bacteria as they can seek out favorable environments in terms of nutrients or avoid the presence of toxins [102]. Bacterial motility can occur on surfaces or in liquids. Swimming motility is the movement of individual bacteria in liquid or low-viscosity environments (below 0.3% agar concentration, in which movement is appreciated as concentric haloes), powered by flagella rotation. On the other hand, bacterial surface motility allows colonization of surfaces and can be distinguished into swarming, twitching, gliding, and sliding [103]. Surface motility is controlled at a physical level via appendages and motors, but it is also influenced by environmental factors [104]. Swarming is a rapid multicellular movement that is powered by flagellar rotation, increase in the number of flagella, cell–cell interactions, and morphological differentiation produced on liquid layers or semisolid surfaces (between 0.5–2% agar concentration) and generally, the movement is observed as a dendritic pattern [103,104]. In this type of motility, surfactants such as rhamnolipids in P. aeruginosa are often produced [105]. Twitching is a type IV pili-mediated surface motility along solid or semisolid surfaces under humid conditions. Type IVa, IVb, and IVb tight adherence (Tad) pili are involved in the assembly of the pilus, extension, attachment to surfaces, and retraction of cells allowing movement [106]. On the other hand, gliding is mediated by surface proteins on top of semisolid agar [107]. Lastly, sliding requires no appendages or cellular components; instead, it is a translocation produced by the expansive forces during growth and special properties in the bacterial surfaces or self-produced substances that reduce the friction with the substrate surface [103]; for instance, P. syringae pv. tomato can slide by producing syringafactin [108]. All the types of motility except gliding have been observed in the Pseudomonas genus [104]. Pseudomonads can respond to chemical gradients in the environment either using polar flagella or type IV pili coupled to a chemosensory system known as chemotaxis [109]. For a suitable plant–bacteria interaction, the free-living bacteria must actively reach the plant roots in a movement mediated by the flagellum and chemotaxis, which has been demonstrated both in vitro and in microcosm experiments in the soil [110,111]. Non-motile mutants or mutants affected in chemotaxis are amongst the most severely impaired Pseudomonas rhizosphere competitive colonization mutants [87,89,90,112], and are also affected in the plant growth promotion and biocontrol abilities [71,87]. Finally, the fact that variants isolated from the rhizosphere are hypermotile and display an enhanced colonization ability emphasizes the importance of motility during the F113 rhizosphere colonization process [74].
Flagella are rotating, rigid helical proteinaceous filaments protruding from the cell surface that drive cells through liquids or surfaces. As previously described, flagellum-dependent movements allow a single bacterium to swim in liquids or bacterial clusters to swarm on surfaces [113,114]. When coupled with a chemotaxis system, swimming motility enables the bacterium to actively evade non-favorable environments and seek out more advantageous conditions resulting in a survival benefit [115]. Flagella are also required for other functions such as community aggregation as they are relevant for measuring environmental conditions (e.g., viscosity, wetness), turning them into a key factor in propagation [116] and virulence (Feldmann et al., 1998). Bacterial flagella transform the movement of ions (H+ and Na+) across the cell membrane into a mechanical torque to move the bacterial cell through its environment [117]. Flagella synthesis is tightly regulated and entails several regulatory pathways in response to environmental signals, guided by the master regulator FleQ in pseudomonads [90,118,119]. In addition to FleQ, the alternative sigma factors FliA (σ28) and RpoN (σ54) are needed for the assembly and flagellin expression [119].
Pseudomonads can produce single or multiple flagella in one pole of the cells and sometimes in a subpolar position. In the genome of F113, two regions of 61 kb and 7.7 kb in length, were identified for their participation in the synthesis of the main flagellar apparatus, which consists of one or two polar flagella with a length ranging from 2 to 4 µm [92]. Flagellar synthesis was studied by Capdevila et al. (2004) in this bacterium and has shown that mutants affected in fleQ and other flagellar structural genes render in no flagellin production giving rise to non-flagellated bacteria. These mutants are non-motile and were displaced by the wild-type strain in competitive root colonization experiments [90]. In addition to the polar flagella, a 41 kb cluster, forming a genetic island, was found in the genome of F113 containing 45 genes encoding proteins involved in the production of a second flagellar apparatus [76,120]. This second flagellum is highly peculiar as it is only found in a few other Pseudomonas strains, mostly belonging to the P. fluorescens complex of species. The genes are homologous and exhibit synteny to the ones encoding flagella in Azotobacter vinelandii and Enterobacteria. The encoding proteins form a tuft of polar flagella that are not produced under laboratory conditions but can be observed in bacteria recovered from the rhizosphere. This second flagellum increases the motility ability and is important for rhizosphere competitive colonization [120].
In P. ogarae F113, several independent pathways regulate motility and the second messenger bis-(3′-5′)-cyclic dimeric guanosine monophosphate (c-di-GMP) plays an important role in motility repression [121,122]. This messenger was already described as responsible for the switching between motility and biofilm formation, among other functions [123]. Low levels of this molecule are associated with a motile lifestyle and high levels with biofilm formation [124]. In F113, the environmental regulation of motility is conducted at three main levels: the synthesis of the first and second flagellar apparatus and flagella rotation. The main regulatory pathways for motility in F113 are summarized in Figure 3 and will be further described.
The synthesis of the primary polar flagellum in F113 is controlled by two membrane proteins, GacS and AdrA [122]. In response to unidentified signals, GacS phosphorylates GacA, which is a positive transcriptional regulator of the small RNAs, rsmX, rsmY, and rsmZ, which in turn can titrate RsmA and RsmE proteins that usually are blocking the translation of specific messenger RNAs, such as the sigma factor AlgU [125]. When the GacA/S two-component system is active, the RsmA and RsmE proteins are recruited by the sRNAs allowing AlgU translation [125,126]. Consequently, when AlgU is translated, it activates the transcription of amrZ, which in turn, repress fleQ, the gene encoding the master flagellar regulator FleQ [125,127,128]. AdrA is a membrane-bound diguanylate cyclase (DGC) that produces the messenger molecule c-di-GMP, which is sensed by the SadB protein [122]. SadB then activates the transcription of the gene encoding the AlgU sigma factor [125]. Therefore, the GacS/GacA pathway and the AdrA/SadB pathway converge in AlgU to regulate the synthesis of flagellar components. As stated before, the expression of fleQ in F113 is negatively regulated by AmrZ [125], which is a protein identified as a regulator of c-di-GMP levels by controlling the transcription of genes encoding most of the DGCs in F113 [128]. In this same work, several AmrZ-regulated c-di-GMP-related proteins with a putative role in swimming motility were identified, namely, DipA, GcbA, and the previously mentioned AdrA. The second flagellar apparatus encoded by F113 genome is cryptic under laboratory conditions and is differentially regulated from the main polar flagellum. The master regulator of this flagellum is FlhDC. The regulatory cascade displayed by the second flagellar apparatus in this bacterium is very similar to the one observed in Enterobacteria and A. vinelandii. The kinase KinB, the adenylate cyclase CyaA, and the cyclic 3′-5′ adenosine phosphate (c-AMP) binding protein Vfr control the master regulator flhDC [120], but their mechanisms remain elusive. Although in other pseudomonads, Vfr was shown to regulate the expression of fleQ [129], in the case of F113, Vfr is only implicated in the regulation of the second flagellar apparatus [120,125]. In this bacterium, AlgU is a negative regulator for the synthesis of the second flagellar apparatus acting over KinB [120]. On the other hand, AmrZ downregulates the expression of flhDC [127]. Interestingly, the production of the primary flagellum is necessary to produce the second one [120].
In F113, c-di-GMP is not only implicated in flagella filament synthesis, as described before. This secondary messenger also controls the flagellar function [71,121]. In F113, there is a PilZ domain-containing protein named FlgZ which is involved in c-di-GMP sensing. FlgZ subcellular localization depends on the intracellular levels of this second messenger, which modulates flagellar rotation likely acting as a clutch [130]. Moreover, Wsp is a chemotaxis-like system whose output protein, WspR, is a DGC involved in the synthesis of c-di-GMP that negatively regulates motility, and positively regulates biofilm formation on abiotic surfaces independently of FleQ in this bacterium [70,121]. The pool of c-di-GMP, produced by the activity of the DGCs, WspR, and SadC, or the phosphodiesterase (PDE) BifA, is sensed by the PilZ domain of FlgZ.

5. Chemotaxis

Chemotaxis allows bacteria to move towards or avoid different environmental signals to ensure beneficial growth conditions [131]. It is a behavior present in movements driven by flagella: swimming and swarming. Chemotaxis is also a highly energetically expensive process due to the requirement of ATP hydrolysis [132] and the flagellar export and assembly apparatus. Chemotaxis is a crucial mechanism for rhizosphere colonization and plant–bacteria interactions allowing bacteria to move towards the plant [133]. Furthermore, chemotaxis pathways are especially abundant in plant-interacting bacteria compared to bacteria of other niches [131]. It has been demonstrated that the motile but non-chemotactic P. fluorescens WCS365 is severely impaired in rhizosphere colonization [112]. Similarly, mutants of P. fluorescens Pf0-1 in genes encoding chemoreceptor proteins involved in recognizing amino acids or organic acids [134,135] are affected in rhizosphere colonization. The chemotaxis cascade resembles a peculiar form of a two-component signaling process. It starts with the recognition of environmental stimuli in the form of chemical gradients (chemo effectors) [131] or changes in the internal energetic conditions such as redox potential [136]. These stimuli are recognized by chemoreceptors, known as methyl-accepting chemotaxis proteins (MCPs) found in the membrane, and that transduce the signal [131]. The MCPs form a ternary complex with the histidine kinase CheA and the adaptor CheW. When the MCP recognizes the external signal, CheA suffers an autophosphorylation process and transphosphorylates the response regulator CheY, which controls flagella rotation [131]. Unlike other pseudomonads, P. ogarae F113 encodes three complete and functional chemotactic systems that are not interchangeable: Che1, Che2, and Che3. Che1 is required for chemotactic motility, being necessary for swimming motility in aerobic and anaerobic conditions. Che3 is required for chemotaxis under anaerobic conditions. Che2 plays a secondary role in chemotaxis. However, the three systems are required for competitive rhizosphere motility, being the mutant affecting Che1 the most impaired [78]. Aside from the Che system, F113 presents two additional chemotaxis-related systems: the Wsp system mentioned earlier [70,121] and the Chp system, located close to the pil genes and thus, with a putative role in twitching motility [76].

6. Microcolony and Biofilm Formation: The Extracellular Matrix (ECM)

The first formal definition of biofilm was established in the late 1990s as “a structured community of bacterial cells enclosed in a self-produced polymeric matrix and adherent to an inert or living surface” [137,138]. Bacteria can adhere to natural or artificial surfaces or themselves, forming biofilms structured by single- or multi-species [139]. Biofilm formation is a protected mode of growth that provides multiple advantages, offering a niche for individuals to establish social interactions such as competition or act as a group with cooperative behavior, commonly using water channels present inside the biofilm structure to exchange nutrients and genetic material. Cells in biofilms also can increase the chances of survival in hostile environments and colonize new niches by dispersal [139,140].
In plant–bacteria interactions, root exudates act as a chemoattractant for bacteria that can attach to the root surface and form microcolonies. Microcolonies can, eventually and in certain strains, grow into larger mature biofilms that contain several layers of cells encased in an extracellular matrix (ECM) or a polymer layer produced as a defensive mechanism by the colonized host [140]. Although F113 is able to form biofilms on inorganic materials, during alfalfa rhizosphere colonization, F113 forms microcolonies encased in a polymeric sheath likely produced by the plant [70]. Generally, bacteria inhabiting the rhizoplane are considered to form biofilms [141] regardless if they form microcolonies or mature biofilms. In the case of F113, it has been shown than the loss of biofilm forming ability does not impair rhizosphere competitive colonization. Both microcolonies and biofilms could be embedded in an ECM, composed of a complex mix of extracellular polymeric substances [142] including: polysaccharides, nucleic acids (extracellular DNA (eDNA) and extracellular RNA), proteins, lipids, and lipoproteins, although the structure and composition of the biofilms can strongly differ between species and environmental conditions [139,142,143]. The ECM composition of model biofilm-forming bacteria is represented in Figure 4. Over the years, several emerging properties and functions, like its role in protecting free-living cells, persistence, collective behavior, stimulation or prevention of biofilm formation, signaling, cell migration, genetic exchange, ion reservoirs, virulence and microbial tolerance have been attributed to it, and are extensively reviewed in Dragoš and Kovács (2017) [144].
The ECM promotes bacterial adhesion to surfaces, often through adhesin–receptor interaction and mechanosensory appendages such as flagella. Once attached, further ECM production surrounds the cells, keeping them in proximity and allowing intercellular interactions (cell–cell cohesion), and being constantly remodeled with different components in each stage. A diverse array of biomolecules secreted to the ECM has been identified and are classified according to their localization (cell surface-associated or extracellularly secreted) [142,145]. Furthermore, the secreted polymeric substances mostly include EPSs such as alginate, proteins, nucleic acids, and lipopolysaccharides (LPSs) with a role in scaffolding and other specialized functions [142,144]. Different species of the genus Pseudomonas produce different ECM components. Regarding exopolysaccharides, some like alginate, are present in almost every species of pseudomonads, while others, such as Psl, poly-N-acetyl-glucosamine (PNAG) or cellulose are present only in a subset of species [146]. P. ogarae F113 encodes in its genome the genes required to produce alginate, PNAG, the Pseudomonas acidic polysaccharide (Pap), and levan. Pap is produced by a limited number of strains, mostly within the Pseudomonas fluorescens complex of species. Many of this species are plant-associated, suggesting a role for this exopolysaccharide in rhizosphere adaption [146]. Regarding extracellular proteins, F113 harbor genes to produce the adhesins LapA and MapA, the large extracellular protein PsmE, the tight adhesion pili Tad, and the functional amyloid proteins Fap. Interestingly, both Tad and Fap proteins are different from their P. aeruginosa counterparts, and seem to have co-evolved with the exopolysaccharide Pap, being therefore also associated with a plant environment [146].
Biofilm formation, the opposed lifestyle to motility, is also tightly regulated. Indeed, the mechanisms that regulate both processes are highly interlinked [147]. There are several factors that influence biofilm formation and dispersion, and different mechanisms by which cells respond to environmental signals. Typical biofilm regulation in Pseudomonas implies quorum sensing, c-di-GMP signaling, and sRNAs through the Gac/Rsm pathway [148]. However, P. ogarae F113 does not have a known quorum sensing signaling system. The Gac/Rsm pathway is also relevant for biofilm formation in several pseudomonads, including F113 [70,71,74,125,149]. As described for the regulation of motility, GacA/GacS post-transcriptionally regulate AlgU through small RNAs. AlgU is a positive regulator of AmrZ that ultimately represses the expression of fleQ. AmrZ is involved in the transcriptional activation of DGCs and repression of PDEs [128] that could ultimately influence flagellar synthesis or function and the ECM composition in this bacterium. Although flagellar-driven motility is the opposed lifestyle to biofilm formation, the role of flagella is indispensable for biofilm formation, especially during the initial step of attachment [150]. When intracellular levels of c-di-GMP are high, FleQ is also involved in the expression of biofilm-related genes and attachment [151,152,153]. The regulation by c-di-GMP is already present in the first step of biofilm formation, during the initial attachment, due to its interaction with FleQ and FlgZ. FleQ bound to c-di-GMP changes its conformation with important consequences in adherence and finally, biofilm formation [151], as it has been shown to repress or activate biofilm formation from the same promoter regions depending on the levels of this second messenger in P. aeruginosa [151,152,153,154]. Likewise, the FlgZ-c-di-GMP complex, which modulates flagellar function, can alter the initial bacterial attachment [130]. Moreover, the c-di-GMP produced by AdrA/SadB, SadC, the Wsp system, and BifA, also play a key role in the regulation of biofilm formation in this bacterium and other pseudomonads [70,121,122,130,149]. The role of c-di-GMP in biofilm formation in pseudomonads has been repeatedly demonstrated. To date, many components of the ECM in several pseudomonads are known to be transcriptionally or post-transcriptionally regulated by c-di-GMP. For example, the LapD protein can sense c-di-GMP and mediates the stability of the important surface adhesin LapA in P. putida and P. fluorescens [155,156]. Similarly, biofilm dispersal is mediated by LapG, also subjected to c-di-GMP control via the PDEs BifA in P. putida and RapA in P. fluorescens [157,158]. A connection between the c-di-GMP and Gac/Rsm signaling pathways has been demonstrated to control biofilm formation through SadC in P. aeruginosa [159] and with quorum sensing via RsmA and RsmE in P. fluorescens 2P24 [160]. Therefore, the interlink between the different pathways is also essential for proper biofilm development.

7. Regulation of Rhizosphere Adaption: The AmrZ-FleQ Hub

Efficient rhizosphere colonization depends on bacterial regulation in response to environmental changes. Aside from the role of the nucleotide messenger c-di-GMP as an integrator of external and internal signals with drastic consequences in bacterial behavior and adaption, transcription factors (TFs) also play a major contribution in the regulation of lifestyle transition in pseudomonads. TFs have been found mainly associated with repression of genes (29.4%), but a considerable number of them can activate and/or repress gene expression (23.9%) and are less common than the TFs that only act as activators (18.1%) [161]. TFs are more abundant in free-living organisms in contrast with pathogenic, extremophilic, or intracellular organisms. The reason is that more complex environments require larger genomes and a tight regulation that allow bacteria to rapidly sense and respond to environmental changes [162,163]. These proteins have been grouped into more than 30 families in prokaryotes, but the majority belong to five major families: LysR, TetR/AcrR, GntR, OmpR, and AraC/XylS [161]. However, the roles of most TFs in prokaryotic organisms remain largely unknown. Within the Pseudomonas genus, a recent work identifying DNA-binding motifs for 100 predicted TFs in P. syringae revealed the existence of a group of master TFs that are involved in multiple important pathways, and a few master TFs involved in other processes such as c-di-GMP turnover, bacterial motility, biofilm formation, siderophore production, and reactive oxygen species [164]. Although there are numerous TFs with key implications in adaption to the rhizosphere environment, research in the last years outlined the role of the global regulator AmrZ and the flagellar master regulator FleQ as a central hub for environmental adaption in pseudomonads.

8. AmrZ

AmrZ is a DNA-binding protein from the ribbon–helix–helix (RHH) protein superfamily that belongs to the AraC family of TFs and is conserved across pseudomonads [165]. AmrZ is composed of a flexible N-terminus, a DNA-binding RHH domain, and a C-terminal tetramerization domain. RHH proteins bind to DNA via recognition of nucleotide sequence by an antiparallel β-sheet and the insertion of an α-helix into the DNA major groove that allows β-sheet binding [166]. It has been demonstrated that AmrZ is a global and bi-functional regulator of gene expression in P. aeruginosa, regulating several important processes such as virulence, motility, EPS synthesis, and c-di-GMP metabolism [167]. AmrZ is controlled by the extra-cytoplasmatic function sigma factor AlgU [168] and is known for being responsible for alginate biosynthesis led by algD [169,170,171,172]. However, a role of this TF in the regulation of other extracellular matrix components aside from alginate has also been demonstrated over the years, such as Psl (negative regulation) and Pel (positive regulation) polysaccharides [167,172,173,174,175]. AmrZ is also involved in the repression of flagellum biosynthesis, swimming, and swarming motilities in P. aeruginosa as AmrZ is a transcriptional repressor of the master flagellar regulator FleQ [175,176,177]. Moreover, AmrZ has been shown to repress the type IV pili-driven twitching motility by inhibiting the expression of pilin (PilA) in P. aeruginosa mucoid and non-mucoid strains [178], and influencing its proper surface localization [179]. In addition, AmrZ can be a repressor or activator of several virulence-related genes in this bacterium [167,180,181]. For instance, it is a direct transcriptional regulator of T6SSs, positively influencing H1- and H3-T6SSs and negatively controlling the H2-T6SS [167,182]. Additionally, AmrZ plays a role in the metabolism of the second messenger c-di-GMP. In P. aeruginosa, mutants in amrZ display elevated levels of c-di-GMP compared with the wild-type strain [167,175], mainly due to the DGC GcbA [167,174,175]. Interestingly, AmrZ has opposite effects in P. syringae pv. tomato DC3000, in which it functions as a positive regulator of flagellar synthesis, alginate [183], swarming motility and virulence in tomato plants [184], whereas it is a negative regulator of cellulose production [183,184]. It is also a positive regulator of motility in P. stutzeri [185]. Similarly, in P. putida KT2440, AmrZ is also a negative regulator of the Pea polysaccharide synthesis [186]. Taken together this information, it has become clear that AmrZ has different behaviors in different Pseudomonas species. This plasticity was the object of study by Baltrus et al. (2018) and led to the observation of a functional switch that took place at least twice independently across the Pseudomonas genus as far as motility regulation [185].
In P. ogarae F113, the model of regulation for AmrZ is more similar to the one observed in P. aeruginosa concerning motility, but has differences in other cellular processes. The first studies described AmrZ as a negative regulator of fleQ and swimming motility [125]. Later, Martínez-Granero et al. (2014b) carried out a chromatin immunoprecipitation sequencing (ChIP-Seq) analysis of AmrZ in this bacterium showing that this protein is a global regulator able to bind to DNA in multiple promoter regions. These regions affect hundreds of genes related to motility and chemotaxis, regulation, and signal transduction, including a great amount of c-di-GMP metabolic enzymes and iron homeostasis, among others. Furthermore, in this study, gene expression analyses showed that AmrZ is mostly a negative regulator of motility and iron homeostasis. These findings suggest an important role of AmrZ in rhizosphere environmental sensing and adaption [127]. Another study of the F113 AmrZ regulon by using RNA-Seq approach and phenotypic analysis of an amrZ mutant [128], showed that this mutant presents a hypermotile phenotype caused by overproduction of flagellar components [125,128], reduced production of c-di-GMP, different colony morphology and dye-binding ability, reduced biofilm formation, and a dramatically impaired rhizosphere competitive colonization ability [128], albeit hypermotility phenotypes were previously shown as advantageous during rhizosphere colonization [70].
One of the main differences between the amrZ mutants in F113 and P. aeruginosa PAO1 is the c-di-GMP levels. Whereas amrZ mutants in P. aeruginosa PAO1 accumulate high levels of this molecule, the mutants in F113 have drastically reduced levels of the second messenger [128,167]. RNA-Seq analysis of the amrZ mutant in F113 also showed that AmrZ is a transcriptional regulator of most DGCs and PDEs encoded in this bacterium [128]. Many of these genes had been previously identified as direct regulatory targets in the ChIP-Seq assay [127]. Further analysis of these AmrZ-regulated genes has shown that most of the phenotypes of the amrZ mutant on motility, ECM components, and biofilm formation, were due to the low levels of c-di-GMP and could be reverted by ectopic production of c-di-GMP [187]. Interestingly, c-di-GMP did not complement the phenotype of impaired rhizosphere colonization, since the amrZ mutant with ectopic production of c-di-GMP was even more impaired in competitive colonization [187].
Jones et al. (2014) proposed that the function of AmrZ as an activator or repressor mostly depends on the location of the DNA-binding site. According to their findings, AmrZ acts as a repressor when bound near the transcriptional start site of genes [167]. A few years later, Xu et al. (2016) also found that AmrZ requires multiple binding sites to function mainly due to the creation of higher-order DNA-AmrZ complexes, as occurs with the algD promoter in which four binding sites are necessary [172]. More recently, Xu et al. (2020) further found multiple binding sites in the pilA promoter region. AmrZ also works as a transcriptional repressor by binding to two sites upstream of its own promoter [171]. This finding is consistent with the fact that AmrZ binds DNA as a dimer of dimers, suggesting it interacts with its targets as oligomers and most likely as tetramers [166,172]. A conceivable scenario is that when AmrZ tetramers are bound to four binding sites, the oligomers can interact through the bent DNA, resulting in a proximity between two DNA ends [172]. Another finding in this sense showed that the C-terminal of AmrZ is necessary for tetramerization, binding, and function [188]. In the last decade, the consensus binding sequence for AmrZ has been described and appears to be conserved among pseudomonads [127,167,180,181].

9. FleQ

FleQ has long been known as a TF belonging to the NtrC family of σN-dependent promoter activators, and is the master regulator of flagellar synthesis in pseudomonads [118,119,189] as described earlier. The FleQ protein contains an N-terminal REC domain, a central AAA+ domain with ATPase activity, and the ability to bind the RpoN factor (σ54). Its C-terminal is a helix–turn–helix DNA-binding domain [118,153,190]. The mode of action for this TF is by activation of the RNA polymerase in concert with the alternate sigma factor RpoN due to its ATPase activity [118]. Its function also relies on direct interactions with FleN, another ATPase that acts as an antagonist [191]. FleQ in solution is found as a dimer, trimer, tetramer, and hexamer [153,190]. As demonstrated in P. aeruginosa, FleQ is the first TF known for its ability to bind c-di-GMP. FleQ does not possess a PilZ domain, but c-di-GMP can interact with the central AAA+ ATP-binding site domain, acting as a competitive inhibitor with much higher affinity compared with ATP, and thus inhibiting its ATPase activity [151,190,192] and making it a c-di-GMP effector [151,154,190,193]. The current accepted model of action is that c-di-GMP binding to FleQ results in an obstruction of its active site, hexameric ring destabilization, quaternary structure reorganization, and allosteric ATPase inhibition [153].
In the last decades, the known roles of FleQ have expanded. In P. fluorescens Pf0-1, the fleQ homolog (adnA) encodes a transcriptional factor that controls persistence and spread in soil, bacterial adhesion, and motility [194,195,196]. In P. aeruginosa, FleQ also was identified as a regulator of Pel and Psl polysaccharides, and CdrA adhesin expression and virulence [151,153,154,197]. Indeed, FleQ can regulate gene expression independently of RpoN, as occurs with biofilm-related genes, such as pel and cdrA in P. aeruginosa [154,197]. Moreover, FleQ has a double function in the regulation of Pel, as an activator or repressor independently of its ATPase activity [151,153,154]. These observations were also evident in P. putida strains KT2440 and KT2442, in which a mutant in fleQ is impaired in flagellar synthesis and biofilm formation, and FleQ has been associated with regulation of synthesis of ECM components such as LapA and EPSs, T6SS, and some c-di-GMP-related genes [193,198,199,200,201]. The interplay with c-di-GMP has also been observed in P. putida and P. fluorescens SBW25 in which FleQ is a transcriptional activator under high c-di-GMP conditions of the bcs and wss operons necessary for the cellulose synthesis [198,199,201,202,203]. Additionally, in P. syringae pv. tomato DC3000, a fleQ mutant is non-motile, displays altered surface spreading on semisolid agar, overproduces the biosurfactant syringafactin, has increased cellulose expression under low c-di-GMP conditions and decreased when the levels of this molecule are high, and it is also altered in virulence [108,184]. In the case of P. ogarae F113, ChIP and RNA-Seq assays have demonstrated the role of FleQ as a global regulator [204,205]. This role is observed both under laboratory cultivation and during rhizosphere colonization [204]. Many of the genes and traits regulated by FleQ in other pseudomonads, such as flagellar synthesis, lapA, alginate production, and the T6SS, are also regulated by FleQ in F113. In the case of motility, fleQ genes from P. putida and P. ogarae are functionally equivalent [205]. In F113 and KT2440, FleQ has been shown to regulate genes implicated in iron homeostasis, ECM component production, biofilm production, and c-di-GMP turnover, among others.

10. The AmrZ-FleQ Hub

There is an interplay between AmrZ and FleQ. These two TFs share a large part of their regulons. In F113, at least 45 genes are regulated by both TFs, generally in opposite ways [205]. This contrasting function is also observed during rhizosphere colonization [204]. Furthermore, AmrZ is a negative regulator of fleQ [127] and FleQ is a negative regulator of amrZ [205]. Based on these results, a model has been proposed (Figure 5). According to this model, AmrZ and FleQ form a regulatory hub which controls environmental adaption genes in opposites ways: AmrZ positively controls iron homeostasis and exopolysaccharide production genes, and negatively controls motility-related genes. FleQ positively controls motility genes and negatively controls exopolysaccharide-related genes and iron homeostasis. The messenger molecule c-di-GMP plays a crucial role in this hub since its production is activated by AmrZ and its sensing by FleQ is necessary for its regulatory activity.
The regulatory role of AmrZ and FleQ in F113 during rhizosphere colonization has been studied by RNA-Seq [204]. It is interesting to note that rhizosphere colonization is the main driver of the F113 transcriptome, showing more influence than the effect of the growth stage (exponential vs. stationary). Under rhizosphere colonization conditions, AmrZ and FleQ regulate traits related to environmental adaption, such as biofilm formation, motility, chemotaxis, denitrification, and c-di-GMP turnover. Many of these genes are regulated in an opposite way by both transcriptional regulators. In this sense, AmrZ appears as an activator of genes implicated in biofilm formation while FleQ act as a repressor of these genes. Similar regulation occurs with denitrification genes, c-di-GMP turnover genes, and chemotaxis genes, which are positively regulated by FleQ and negatively by AmrZ. AmrZ and FleQ also regulate the T6SS in the rhizosphere. P. ogarae F113 harbors three T6SSs, and at least two of them, F1 and F3, are functional for bacterial killing [206]. Transcriptomic data has shown that in the rhizosphere, AmrZ and FleQ act as negative and positive regulators, respectively, for F1-T6SS and F3-T6SS. It is interesting to note that a mutant affecting these two T6SSs is impaired in adaption and persistence in the rhizosphere microbiome, highlighting its relevance for the adaption to the rhizosphere niche.

11. Concluding Remarks

Rhizosphere colonization is the main lifestyle of Pseudomonas ogarae F113 and other plant-associated pseudomonads. Research has shown that many traits, such as motility, biofilm and microcolony formation, iron scavenging and protein secretion systems, among others, are important for adaption to the rhizosphere environment. It has also been shown that the messenger molecule c-di-GMP and the transcriptional factors AmrZ and FleQ regulate many of these traits in a coordinated manner, forming a regulatory hub that works like an oscillator, which regulates rhizosphere adaption traits in an opposite mode. Future research in this field will identify novel genes implicated in the environmental regulation of motility. It will also clarify the role of individual components of extracellular matrix components in biofilm formation and their role in rhizosphere colonization. The role of specific proteins and/or polysaccharides in attachment to biotic surfaces for rhizoplane colonization will also contribute to our knowledge of bacterial adaption to the rhizosphere environment.

Author Contributions

Conceptualization, performed a review of the literature, and writing—original manuscript preparation, E.B.-R., M.M. and R.R.; reviewing and editing, E.B.-R., M.M., R.R., M.R.-N., D.G.-S. and D.D. All authors have read and agreed to the published version of the manuscript.

Funding

Work related to environmental adaption of pseudomonads in the authors’ laboratory is currently funded by a Spanish Ministerio de Ciencia e Innovación FEDER/EU Grant PID2021-125070OB-I00.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hiltner, L. Uber nevere Erfahrungen und Probleme auf dem Gebiet der Boden Bakteriologie und unter besonderer Beurchsichtigung der Grundungung und Broche. Arbeit. Deut. Landw. Ges. Berl. 1904, 98, 59–78. [Google Scholar]
  2. Lynch, J.; de Leij, F. Rhizosphere; eLS, John Wiley & Sons, Ltd.: Chichester, UK, 2012. [Google Scholar]
  3. Bazin, M.; Markham, P.; Scott, E.; Lynch, J. Population dynamics and rhizosphere interactions. In The Rhizosphere; John Wiley & Sons, Ltd.: Chichester, UK, 1990; pp. 99–127. [Google Scholar]
  4. Wang, N.R.; Haney, C.H. Harnessing the genetic potential of the plant microbiome. Biochemist 2020, 42, 20–25. [Google Scholar] [CrossRef]
  5. Fitzpatrick, C.R.; Salas-González, I.; Conway, J.M.; Finkel, O.M.; Gilbert, S.; Russ, D.; Teixeira, P.J.P.L.; Dangl, J.L. The Plant Microbiome: From Ecology to Reductionism and Beyond. Annu. Rev. Microbiol. 2020, 74, 81–100. [Google Scholar] [CrossRef]
  6. Jones, D.L.; Nguyen, C.; Finlay, R.D. Carbon flow in the rhizosphere: Carbon trading at the soil–root interface. Plant Soil 2009, 321, 5–33. [Google Scholar] [CrossRef]
  7. Walker, T.S.; Bais, H.P.; Grotewold, E.; Vivanco, J.M. Root exudation and rhizosphere biology. Plant Physiol. 2003, 132, 44–51. [Google Scholar] [CrossRef] [Green Version]
  8. 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] [PubMed]
  9. Venturi, V.; Keel, C. Signaling in the rhizosphere. Trends Plant Sci. 2016, 21, 187–198. [Google Scholar] [CrossRef]
  10. Bloemberg, G.V.; Lugtenberg, B.J. Molecular basis of plant growth promotion and biocontrol by rhizobacteria. Curr. Opin. Plant Biol. 2001, 4, 343–350. [Google Scholar] [CrossRef]
  11. Berendsen, R.L.; Vismans, G.; Yu, K.; Song, Y.; de Jonge, R.; Burgman, W.P.; Burmølle, M.; Herschend, J.; Bakker, P.A.; Pieterse, C.M. Disease-induced assemblage of a plant-beneficial bacterial consortium. ISME J. 2018, 12, 1496–1507. [Google Scholar] [CrossRef] [Green Version]
  12. Schroth, M.N.; Hancock, J.G. Disease-suppressive soil and root-colonizing bacteria. Science 1982, 216, 1376–1381. [Google Scholar] [CrossRef]
  13. Gray, E.; Smith, D. Intracellular and extracellular PGPR: Commonalities and distinctions in the plant–bacterium signaling processes. Soil Biol. Biochem. 2005, 37, 395–412. [Google Scholar] [CrossRef]
  14. Kloepper, J.W. Plant growth-promoting rhizobacteria on radishes. In Proceedings of the 4th International Conference on Plant Pathogenic Bacter, Station de Pathologie Vegetale et Phytobacteriologie, INRA, Angers, France, 27 August–2 September 1978; pp. 879–882. [Google Scholar]
  15. Jacobsen, B.; Zidack, N.; Larson, B. The role of Bacillus-based biological control agents in integrated pest management systems: Plant diseases. Phytopathology 2004, 94, 1272–1275. [Google Scholar] [CrossRef] [Green Version]
  16. Haas, D.; Défago, G. Biological control of soil-borne pathogens by fluorescent pseudomonads. Nat. Rev. Microbiol. 2005, 3, 307–319. [Google Scholar] [CrossRef]
  17. Beattie, G.A. Plant-associated bacteria: Survey, molecular phylogeny, genomics and recent advances. In Plant-Associated Bacteria; Springer: Dordrecht, The Netherlands, 2007; pp. 1–56. [Google Scholar]
  18. Pastor-Bueis, R.; Mulas, R.; Gómez, X.; González-Andrés, F. Innovative liquid formulation of digestates for producing a biofertilizer based on Bacillus siamensis: Field testing on sweet pepper. J. Plant Nutr. Soil Sci. 2017, 180, 748–758. [Google Scholar] [CrossRef]
  19. Shirley, M.; Avoscan, L.; Bernaud, E.; Vansuyt, G.; Lemanceau, P. Comparison of iron acquisition from Fe–pyoverdine by strategy I and strategy II plants. Botany 2011, 89, 731–735. [Google Scholar] [CrossRef]
  20. Khan, N.; Bano, A.; Rahman, M.A.; Guo, J.; Kang, Z.; Babar, M.A. Comparative physiological and metabolic analysis reveals a complex mechanism involved in drought tolerance in chickpea (Cicer arietinum L.) induced by PGPR and PGRs. Sci. Rep. 2019, 9, 2097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Ortiz, N.; Armada, E.; Duque, E.; Roldán, A.; Azcón, R. Contribution of arbuscular mycorrhizal fungi and/or bacteria to enhancing plant drought tolerance under natural soil conditions: Effectiveness of autochthonous or allochthonous strains. J. Plant Physiol. 2015, 174, 87–96. [Google Scholar] [CrossRef] [PubMed]
  22. Glick, B.R. Bacteria with ACC deaminase can promote plant growth and help to feed the world. Microbiol. Res. 2014, 169, 30–39. [Google Scholar] [CrossRef]
  23. Glick, B.R. The enhancement of plant growth by free-living bacteria. Can. J. Microbiol. 1995, 41, 109–117. [Google Scholar] [CrossRef]
  24. Riley, M.A.; Wertz, J.E. Bacteriocins: Evolution, ecology, and application. Annu. Rev. Microbiol. 2002, 56, 117–137. [Google Scholar] [CrossRef] [Green Version]
  25. Neeraja, C.; Anil, K.; Purushotham, P.; Suma, K.; Sarma, P.; Moerschbacher, B.M.; Podile, A.R. Biotechnological approaches to develop bacterial chitinases as a bioshield against fungal diseases of plants. Crit. Rev. Biotechnol. 2010, 30, 231–241. [Google Scholar] [CrossRef]
  26. Złoch, M.; Thiem, D.; Gadzała-Kopciuch, R.; Hrynkiewicz, K. Synthesis of siderophores by plant-associated metallotolerant bacteria under exposure to Cd2+. Chemosphere 2016, 156, 312–325. [Google Scholar] [CrossRef]
  27. Kamilova, F.; Validov, S.; Azarova, T.; Mulders, I.; Lugtenberg, B. Enrichment for enhanced competitive plant root tip colonizers selects for a new class of biocontrol bacteria. Environ. Microbiol. 2005, 7, 1809–1817. [Google Scholar] [CrossRef] [PubMed]
  28. Van Loon, L.; Bakker, P.; Pieterse, C. Induction and expression of PGPR-mediated induced resistance against pathogens. IOBC/Wprs Bull. 1998, 21, 103–110. [Google Scholar]
  29. Beneduzi, A.; Ambrosini, A.; Passaglia, L.M. Plant growth-promoting rhizobacteria (PGPR): Their potential as antagonists and biocontrol agents. Genet. Mol. Biol. 2012, 35, 1044–1051. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Stanier, R.Y.; Palleroni, N.J.; Doudoroff, M. The aerobic pseudomonads a taxonomic study. Microbiology 1966, 43, 159–271. [Google Scholar] [CrossRef] [Green Version]
  31. Palleroni, N.J. Prokaryote taxonomy of the 20th century and the impact of studies on the genus Pseudomonas: A personal view. Microbiology 2003, 149, 1–7. [Google Scholar] [CrossRef] [Green Version]
  32. Palleroni, N.J. The Pseudomonas story. Environ. Microbiol. 2010, 12, 1377–1383. [Google Scholar] [CrossRef]
  33. Parte, A.C. LPSN—List of prokaryotic names with standing in nomenclature (bacterio.net), 20 years on. Int. J. Syst. Evol. Microbiol. 2018, 68, 1825–1829. [Google Scholar] [CrossRef]
  34. Palleroni, N.J.; Genus, I. Pseudomonas. In Bergey’s Manual of Systematic Bacteriology, 1st ed.; Bergey, D.H., Krieg, N.R., Holt, J.G., Eds.; The Williams and Wilkins Co.: Baltimore, MD, USA, 1984; Volume 1, pp. 141–199. [Google Scholar]
  35. Raaijmakers, J.M.; Mazzola, M. Diversity and natural functions of antibiotics produced by beneficial and plant pathogenic bacteria. Annu. Rev. Phytopathol. 2012, 50, 403–424. [Google Scholar] [CrossRef]
  36. Ramos, J.L. Pseudomonas Volume 1: Genomics, Life Style and Molecular Architecture; Springer: New York, NY, USA, 2004. [Google Scholar]
  37. Morris, C.E.; Sands, D.C.; Vinatzer, B.A.; Glaux, C.; Guilbaud, C.; Buffiere, A.; Yan, S.; Dominguez, H.; Thompson, B.M. The life history of the plant pathogen Pseudomonas syringae is linked to the water cycle. ISME J. 2008, 2, 321–334. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Waters, C.M.; Goldberg, J.B. Pseudomonas aeruginosa in cystic fibrosis: A chronic cheater. Proc. Natl. Acad. Sci. USA 2019, 116, 6525–6527. [Google Scholar] [CrossRef] [Green Version]
  39. Xin, X.-F.; Kvitko, B.; He, S.Y. Pseudomonas syringae: What it takes to be a pathogen. Nat. Rev. Microbiol. 2018, 16, 316. [Google Scholar] [CrossRef]
  40. Garrido-Sanz, D.; Meier-Kolthoff, J.P.; Göker, M.; Martín, M.; Rivilla, R.; Redondo-Nieto, M. Genomic and Genetic Diversity within the Pseudomonas fluorescens Complex. PLoS ONE 2016, 11, e0150183. [Google Scholar] [CrossRef] [Green Version]
  41. Wasi, S.; Tabrez, S.; Ahmad, M. Use of Pseudomonas spp. for the bioremediation of environmental pollutants: A review. Environ. Monit. Assess. 2013, 185, 8147–8155. [Google Scholar] [CrossRef] [PubMed]
  42. Olivera, E.R.; Carnicero, D.; Jodra, R.; Miñambres, B.; García, B.; Abraham, G.A.; Gallardo, A.; Román, J.S.; García, J.L.; Naharro, G. Genetically engineered Pseudomonas: A factory of new bioplastics with broad applications. Environ. Microbiol. 2001, 3, 612–618. [Google Scholar] [CrossRef]
  43. Nikel, P.I.; De Lorenzo, V. Pseudomonas putida as a functional chassis for industrial biocatalysis: From native biochemistry to trans-metabolism. Metab. Eng. 2018, 50, 142–155. [Google Scholar] [CrossRef] [PubMed]
  44. Mulet, M.; Lalucat, J.; García-Valdés, E. DNA sequence-based analysis of the Pseudomonas species. Environ. Microbiol. 2010, 12, 1513–1530. [Google Scholar]
  45. Gomila, M.; Peña, A.; Mulet, M.; Lalucat, J.; García-Valdés, E. Phylogenomics and systematics in Pseudomonas. Front. Microbiol. 2015, 6, 214. [Google Scholar] [CrossRef] [Green Version]
  46. Lalucat, J.; Mulet, M.; Gomila, M.; García-Valdés, E. Genomics in Bacterial Taxonomy: Impact on the Genus Pseudomonas. Genes 2020, 11, 139. [Google Scholar] [CrossRef] [Green Version]
  47. Seaton, S.C.; Silby, M.W. Genetics and functional genomics of the Pseudomonas fluorescens Group. In Genomics of Plant-Associated Bacteria; Springer: Berlin/Heidelberg, Germany, 2014; pp. 99–125. [Google Scholar]
  48. Weller, D.M.; Landa, B.; Mavrodi, O.; Schroeder, K.; De La Fuente, L.; Blouin Bankhead, S.; Allende Molar, R.; Bonsall, R.; Mavrodi, D.; Thomashow, L. Role of 2, 4-diacetylphloroglucinol-producing fluorescent Pseudomonas spp. in the defense of plant roots. Plant Biol. 2007, 9, 4–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Couillerot, O.; Prigent-Combaret, C.; Caballero-Mellado, J.; Moënne-Loccoz, Y. Pseudomonas fluorescens and closely-related fluorescent pseudomonads as biocontrol agents of soil-borne phytopathogens. Lett. Appl. Microbiol. 2009, 48, 505–512. [Google Scholar] [CrossRef] [PubMed]
  50. Loper, J.E.; Hassan, K.A.; Mavrodi, D.V.; Davis II, E.W.; Lim, C.K.; Shaffer, B.T.; Elbourne, L.D.; Stockwell, V.O.; Hartney, S.L.; Breakwell, K. Comparative genomics of plant-associated Pseudomonas spp.: Insights into diversity and inheritance of traits involved in multitrophic interactions. PLoS Genet. 2012, 8, e1002784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Garrido-Sanz, D.; Arrebola, E.; Martínez-Granero, F.; García-Méndez, S.; Muriel, C.; Blanco-Romero, E.; Martín, M.; Rivilla, R.; Redondo-Nieto, M. Classification of isolates from the Pseudomonas fluorescens complex into phylogenomic groups based in group-specific markers. Front. Microbiol. 2017, 8, 413. [Google Scholar] [CrossRef] [Green Version]
  52. Miller, S.H.; Browne, P.; Prigent-Combaret, C.; Combes-Meynet, E.; Morrissey, J.P.; O’Gara, F. Biochemical and genomic comparison of inorganic phosphate solubilization in Pseudomonas species. Environ. Microbiol. Rep. 2010, 2, 403–411. [Google Scholar] [CrossRef] [PubMed]
  53. Raaijmakers, J.M.; Leeman, M.; Van Oorschot, M.M.; Van der Sluis, I.; Schippers, B.; Bakker, P. Dose-response relationships in biological control of Fusarium wilt of radish by Pseudomonas spp. Phytopathology 1995, 85, 1075–1080. [Google Scholar] [CrossRef]
  54. Visca, P.; Imperi, F.; Lamont, I.L. Pyoverdine siderophores: From biogenesis to biosignificance. Trends Microbiol. 2007, 15, 22–30. [Google Scholar] [CrossRef]
  55. Picard, C.; Bosco, M. Maize heterosis affects the structure and dynamics of indigenous rhizospheric auxins-producing Pseudomonas populations. FEMS Microbiol. Ecol. 2005, 53, 349–357. [Google Scholar] [CrossRef] [Green Version]
  56. Blaha, D.; Prigent-Combaret, C.; Mirza, M.S.; Moënne-Loccoz, Y. Phylogeny of the 1-aminocyclopropane-1-carboxylic acid deaminase-encoding gene acdS in phytobeneficial and pathogenic Proteobacteria and relation with strain biogeography. FEMS Microbiol. Ecol. 2006, 56, 455–470. [Google Scholar] [CrossRef] [Green Version]
  57. Couillerot, O.; Combes-Meynet, E.; Pothier, J.F.; Bellvert, F.; Challita, E.; Poirier, M.-A.; Rohr, R.; Comte, G.; Moënne-Loccoz, Y.; Prigent-Combaret, C. The role of the antimicrobial compound 2,4-diacetylphloroglucinol in the impact of biocontrol Pseudomonas fluorescens F113 on Azospirillum brasilense phytostimulators. Microbiology 2011, 157, 1694–1705. [Google Scholar] [CrossRef] [Green Version]
  58. Nandi, M.; Selin, C.; Brawerman, G.; Fernando, W.D.; de Kievit, T. Hydrogen cyanide, which contributes to Pseudomonas chlororaphis strain PA23 biocontrol, is upregulated in the presence of glycine. Biol. Control 2017, 108, 47–54. [Google Scholar] [CrossRef]
  59. Garrido-Sanz, D.; Redondo-Nieto, M.; Martin, M.; Rivilla, R. Comparative genomics of the Pseudomonas corrugata subgroup reveals high species diversity and allows the description of Pseudomonas ogarae sp. nov. Microb. Genom. 2021, 7, 000593. [Google Scholar] [CrossRef]
  60. Shanahan, P.; O’Sullivan, D.J.; Simpson, P.; Glennon, J.D.; O’Gara, F. Isolation of 2, 4-diacetylphloroglucinol from a fluorescent pseudomonad and investigation of physiological parameters influencing its production. Appl. Environ. Microbiol. 1992, 58, 353–358. [Google Scholar] [CrossRef] [Green Version]
  61. Simons, M.; Van Der Bij, A.; Brand, I.; De Weger, L.; Wijffelman, C.; Lugtenberg, B. Gnotobiotic system for studying rhizosphere colonization by plant growth-promoting Pseudomonas bacteria. Mol. Plant-Microbe Interact. MPMI 1996, 9, 600–607. [Google Scholar] [CrossRef] [PubMed]
  62. Cronin, D.; Moënne-Loccoz, Y.; Fenton, A.; Dunne, C.; Dowling, D.N.; O’Gara, F. Ecological interaction of a biocontrol Pseudomonas fluorescens strain producing 2, 4-diacetylphloroglucinol with the soft rot potato pathogen Erwinia carotovora subsp. atroseptica. FEMS Microbiol. Ecol. 1997, 23, 95–106. [Google Scholar] [CrossRef]
  63. Naseby, D.; Lynch, J. Effects of Pseudomonas fluorescens F113 on ecological functions in the pea rhizosphere are dependent on pH. Microb. Ecol. 1999, 37, 248–256. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Dekkers, L.C.; Mulders, I.H.; Phoelich, C.C.; Chin-A-Woeng, T.F.; Wijfjes, A.H.; Lugtenberg, B.J. The sss colonization gene of the tomato-Fusarium oxysporum f. sp. radicis-lycopersici biocontrol strain Pseudomonas fluorescens WCS365 can improve root colonization of other wild-type Pseudomonas spp. bacteria. Mol. Plant-Microbe Interact. 2000, 13, 1177–1183. [Google Scholar] [PubMed] [Green Version]
  65. Villacieros, M.; Power, B.; Sánchez-Contreras, M.; Lloret, J.; Oruezabal, R.I.; Martín, M.; Fernández-Piñas, F.; Bonilla, I.; Whelan, C.; Dowling, D.N.; et al. Colonization behaviour of Pseudomonas fluorescens and Sinorhizobium meliloti in the alfalfa (Medicago sativa) rhizosphere. Plant Soil 2003, 251, 47–54. [Google Scholar] [CrossRef]
  66. Villacieros, M.; Whelan, C.; Mackova, M.; Molgaard, J.; Sánchez-Contreras, M.; Lloret, J.; de Cárcer, D.A.; Oruezábal, R.I.; Bolaños, L.; Macek, T.; et al. Polychlorinated biphenyl rhizoremediation by Pseudomonas fluorescens F113 derivatives, using a Sinorhizobium meliloti nod system to drive bph gene expression. Appl. Environ. Microbiol. 2005, 71, 2687–2694. [Google Scholar] [CrossRef] [Green Version]
  67. De La Fuente, L.; Landa, B.B.; Weller, D.M. Host crop affects rhizosphere colonization and competitiveness of 2,4-diacetylphloroglucinol-producing Pseudomonas fluorescens. Phytopathology 2006, 96, 751–762. [Google Scholar] [CrossRef] [Green Version]
  68. De Cárcer, D.A.; Martín, M.; Karlson, U.; Rivilla, R. Changes in bacterial populations and in biphenyl dioxygenase gene diversity in a polychlorinated biphenyl-polluted soil after introduction of willow trees for rhizoremediation. Appl. Environ. Microbiol. 2007, 73, 6224–6232. [Google Scholar] [CrossRef] [Green Version]
  69. Rein, A.; Fernqvist, M.M.; Mayer, P.; Trapp, S.; Bittens, M.; Karlson, U.G. Degradation of PCB congeners by bacterial strains. Appl. Microbiol. Biotechnol. 2007, 77, 469–481. [Google Scholar] [CrossRef] [Green Version]
  70. Barahona, E.; Navazo, A.; Yousef-Coronado, F.; Aguirre de Cárcer, D.; Martínez-Granero, F.; Espinosa-Urgel, M.; Martín, M.; Rivilla, R. Efficient rhizosphere colonization by Pseudomonas fluorescens F113 mutants unable to form biofilms on abiotic surfaces. Environ. Microbiol. 2010, 12, 3185–3195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Barahona, E.; Navazo, A.; Martínez-Granero, F.; Zea-Bonilla, T.; Pérez-Jiménez, R.M.; Martín, M.; Rivilla, R. Pseudomonas fluorescens F113 mutant with enhanced competitive colonization ability and improved biocontrol activity against fungal root pathogens. Appl. Environ. Microbiol. 2011, 77, 5412–5419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Vacheron, J.; Combes-Meynet, E.; Walker, V.; Gouesnard, B.; Muller, D.; Moënne-Loccoz, Y.; Prigent-Combaret, C. Expression on roots and contribution to maize phytostimulation of 1-aminocyclopropane-1-decarboxylate deaminase gene acdS in Pseudomonas fluorescens F113. Plant Soil 2016, 407, 187–202. [Google Scholar] [CrossRef]
  73. Vacheron, J.; Desbrosses, G.; Renoud, S.; Padilla, R.; Walker, V.; Muller, D.; Prigent-Combaret, C. Differential contribution of plant-beneficial functions from Pseudomonas kilonensis F113 to root system architecture alterations in Arabidopsis thaliana and Zea mays. Mol. Plant-Microbe Interact. 2018, 31, 212–223. [Google Scholar] [CrossRef] [Green Version]
  74. Martínez-Granero, F.; Rivilla, R.; Martín, M. Rhizosphere selection of highly motile phenotypic variants of Pseudomonas fluorescens with enhanced competitive colonization ability. Appl. Environ. Microbiol. 2006, 72, 3429–3434. [Google Scholar] [CrossRef] [Green Version]
  75. Redondo-Nieto, M.; Barret, M.; Morrisey, J.P.; Germaine, K.; Martínez-Granero, F.; Barahona, E.; Navazo, A.; Sánchez-Contreras, M.; Moynihan, J.A.; Giddens, S.R.; et al. Genome sequence of the biocontrol strain Pseudomonas fluorescens F113. J. Bacteriol. 2012, 194, 1273–1274. [Google Scholar] [CrossRef] [Green Version]
  76. Redondo-Nieto, M.; Barret, M.; Morrissey, J.; Germaine, K.; Martínez-Granero, F.; Barahona, E.; Navazo, A.; Sánchez-Contreras, M.; Moynihan, J.A.; Muriel, C.; et al. Genome sequence reveals that Pseudomonas fluorescens F113 possesses a large and diverse array of systems for rhizosphere function and host interaction. BMC Genom. 2013, 14, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Rezzonico, F.; Zala, M.; Keel, C.; Duffy, B.; Moënne-Loccoz, Y.; Défago, G. Is the ability of biocontrol fluorescent pseudomonads to produce the antifungal metabolite 2, 4-diacetylphloroglucinol really synonymous with higher plant protection? New Phytol. 2007, 173, 861–872. [Google Scholar] [CrossRef]
  78. Muriel, C.; Jalvo, B.; Redondo-Nieto, M.; Rivilla, R.; Martín, M. Chemotactic motility of Pseudomonas fluorescens F113 under aerobic and denitrification conditions. PLoS ONE 2015, 10, e0132242. [Google Scholar] [CrossRef] [Green Version]
  79. Ghirardi, S.; Dessaint, F.; Mazurier, S.; Corberand, T.; Raaijmakers, J.M.; Meyer, J.-M.; Dessaux, Y.; Lemanceau, P. Identification of traits shared by rhizosphere-competent strains of fluorescent pseudomonads. Microb. Ecol. 2012, 64, 725–737. [Google Scholar] [CrossRef]
  80. Fenton, A.; Stephens, P.; Crowley, J.; O’Callaghan, M.; O’Gara, F. Exploitation of gene (s) involved in 2, 4-diacetylphloroglucinol biosynthesis to confer a new biocontrol capability to a Pseudomonas strain. Appl. Environ. Microbiol. 1992, 58, 3873–3878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Moënne-Loccoz, Y.; Tichy, H.-V.; O’Donnell, A.; Simon, R.; O’Gara, F. Impact of 2, 4-Diacetylphloroglucinol-Producing Biocontrol Strain Pseudomonas fluorescens F113 on Intraspecific Diversity of Resident Culturable Fluorescent Pseudomonads Associated with the Roots of Field-Grown Sugar Beet Seedlings. Appl. Environ. Microbiol. 2001, 67, 3418–3425. [Google Scholar] [CrossRef] [Green Version]
  82. Marshall, K.C. Planktonic versus sessile life of prokaryotes. Prokaryotes 2006, 2, 3–15. [Google Scholar]
  83. Sauer, K.; Camper, A.K.; Ehrlich, G.D.; Costerton, J.W.; Davies, D.G. Pseudomonas aeruginosa displays multiple phenotypes during development as a biofilm. J. Bacteriol. 2002, 184, 1140–1154. [Google Scholar] [CrossRef] [Green Version]
  84. Wan, N.; Wang, H.; Ng, C.K.; Mukherjee, M.; Ren, D.; Cao, B.; Tang, Y.J. Bacterial metabolism during biofilm growth investigated by 13C tracing. Front. Microbiol. 2018, 9, 2657. [Google Scholar] [CrossRef] [Green Version]
  85. O’Toole, G.; Kaplan, H.B.; Kolter, R. Biofilm formation as microbial development. Annu. Rev. Microbiol. 2000, 54, 49–79. [Google Scholar] [CrossRef] [PubMed]
  86. Lugtenberg, B.J.; Dekkers, L.; Bloemberg, G.V. Molecular determinants of rhizosphere colonization by Pseudomonas. Annu. Rev. Phytopathol. 2001, 39, 461–490. [Google Scholar] [CrossRef]
  87. Chin-A-Woeng, T.F.; Bloemberg, G.V.; Mulders, I.H.; Dekkers, L.C.; Lugtenberg, B.J. Root colonization by phenazine-1-carboxamide-producing bacterium Pseudomonas chlororaphis PCL1391 is essential for biocontrol of tomato foot and root rot. Mol. Plant-Microbe Interact. 2000, 13, 1340–1345. [Google Scholar] [CrossRef] [Green Version]
  88. Pliego, C.; De Weert, S.; Lamers, G.; De Vicente, A.; Bloemberg, G.; Cazorla, F.M.; Ramos, C. Two similar enhanced root-colonizing Pseudomonas strains differ largely in their colonization strategies of avocado roots and Rosellinia necatrix hyphae. Environ. Microbiol. 2008, 10, 3295–3304. [Google Scholar] [CrossRef] [PubMed]
  89. Bulgarelli, D.; Schlaeppi, K.; Spaepen, S.; Van Themaat, E.V.L.; Schulze-Lefert, P. Structure and functions of the bacterial microbiota of plants. Annu. Rev. Plant Biol. 2013, 64, 807–838. [Google Scholar] [CrossRef] [Green Version]
  90. Capdevila, S.; Martínez-Granero, F.M.; Sánchez-Contreras, M.; Rivilla, R.; Martín, M. Analysis of Pseudomonas fluorescens F113 genes implicated in flagellar filament synthesis and their role in competitive root colonization. Microbiology 2004, 150, 3889–3897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  91. Saunders, N.J.; Moxon, E.R.; Gravenor, M.B. Mutation rates: Estimating phase variation rates when fitness differences are present and their impact on population structure. Microbiology 2003, 149, 485–495. [Google Scholar] [CrossRef] [Green Version]
  92. Sánchez-Contreras, M.; Martín, M.; Villacieros, M.; O’Gara, F.; Bonilla, I.; Rivilla, R. Phenotypic selection and phase variation occur during alfalfa root colonization by Pseudomonas fluorescens F113. J. Bacteriol. 2002, 184, 1587–1596. [Google Scholar] [CrossRef] [Green Version]
  93. Martinez-Granero, F.; Capdevila, S.; Sanchez-Contreras, M.; Martin, M.; Rivilla, R. Two site-specific recombinases are implicated in phenotypic variation and competitive rhizosphere colonization in Pseudomonas fluorescens. Microbiology 2005, 151, 975–983. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Ramos-González, M.I.; Matilla, M.A.; Quesada, J.M.; Ramos, J.L.; Espinosa-Urgel, M. Using genomics to unveil bacterial determinants of rhizosphere life style. Mol. Microb. Ecol. Rhizosphere 2013, 1, 5–16. [Google Scholar]
  95. Levy, A.; Conway, J.M.; Dangl, J.L.; Woyke, T. Elucidating bacterial gene functions in the plant microbiome. Cell Host Microbe 2018, 24, 475–485. [Google Scholar] [CrossRef] [Green Version]
  96. Levy, A.; González, I.S.; Mittelviefhaus, M.; Clingenpeel, S.; Paredes, S.H.; Miao, J.; Wang, K.; Devescovi, G.; Stillman, K.; Monteiro, F. Genomic features of bacterial adaptation to plants. Nat. Genet. 2018, 50, 138–150. [Google Scholar] [CrossRef] [Green Version]
  97. Ofek-Lalzar, M.; Sela, N.; Goldman-Voronov, M.; Green, S.J.; Hadar, Y.; Minz, D. Niche and host-associated functional signatures of the root surface microbiome. Nat. Commun. 2014, 5, 4950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Cole, B.J.; Feltcher, M.E.; Waters, R.J.; Wetmore, K.M.; Mucyn, T.S.; Ryan, E.M.; Wang, G.; Ul-Hasan, S.; McDonald, M.; Yoshikuni, Y. Genome-wide identification of bacterial plant colonization genes. PLoS Biol. 2017, 15, e2002860. [Google Scholar] [CrossRef] [PubMed]
  99. Helmann, T.C.; Deutschbauer, A.M.; Lindow, S.E. Genome-wide identification of Pseudomonas syringae genes required for fitness during colonization of the leaf surface and apoplast. Proc. Natl. Acad. Sci. USA 2019, 116, 18900–18910. [Google Scholar] [CrossRef] [Green Version]
  100. Armanhi, J.S.L.; de Souza, R.S.C.; Damasceno, N.d.B.; de Araújo, L.M.; Imperial, J.; Arruda, P. A community-based culture collection for targeting novel plant growth-promoting bacteria from the sugarcane microbiome. Front. Plant Sci. 2018, 8, 2191. [Google Scholar] [CrossRef] [Green Version]
  101. Arruda, B.; Rodrigues, M.; Gumiere, T.; Richardson, A.E.; Andreote, F.D.; Soltangheisi, A.; Gatiboni, L.C.; Pavinato, P.S. The impact of sugarcane filter cake on the availability of P in the rhizosphere and associated microbial community structure. Soil Use Manag. 2019, 35, 334–345. [Google Scholar] [CrossRef]
  102. Wei, Y.; Wang, X.; Liu, J.; Nememan, I.; Singh, A.H.; Weiss, H.; Levin, B.R. The population dynamics of bacteria in physically structured habitats and the adaptive virtue of random motility. Proc. Natl. Acad. Sci. USA 2011, 108, 4047–4052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Henrichsen, J. Bacterial surface translocation: A survey and a classification. Bacteriol. Rev. 1972, 36, 478. [Google Scholar] [CrossRef]
  104. Mattingly, A.E.; Weaver, A.A.; Dimkovikj, A.; Shrout, J.D. Assessing travel conditions: Environmental and host influences on bacterial surface motility. J. Bacteriol. 2018, 200, e00014-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Shrout, J.D.; Chopp, D.L.; Just, C.L.; Hentzer, M.; Givskov, M.; Parsek, M.R. The impact of quorum sensing and swarming motility on Pseudomonas aeruginosa biofilm formation is nutritionally conditional. Mol. Microbiol. 2006, 62, 1264–1277. [Google Scholar] [CrossRef] [PubMed]
  106. Burrows, L.L. Pseudomonas aeruginosa twitching motility: Type IV pili in action. Annu. Rev. Microbiol. 2012, 66, 493–520. [Google Scholar] [CrossRef] [Green Version]
  107. Nan, B. Bacterial gliding motility: Rolling out a consensus model. Curr. Biol. 2017, 27, R154–R156. [Google Scholar] [CrossRef] [Green Version]
  108. Nogales, J.; Vargas, P.; Farias, G.A.; Olmedilla, A.; Sanjuán, J.; Gallegos, M.-T. FleQ coordinates flagellum-dependent and-independent motilities in Pseudomonas syringae pv. tomato DC3000. Appl. Environ. Microbiol. 2015, 81, 7533–7545. [Google Scholar] [CrossRef] [Green Version]
  109. Sampedro, I.; Parales, R.E.; Krell, T.; Hill, J.E. Pseudomonas chemotaxis. FEMS Microbiol. Rev. 2015, 39, 17–46. [Google Scholar]
  110. Bashan, Y. Migration of the rhizosphere bacteria Azospirillum brasilense and Pseudomonas fluorescens towards wheat roots in the soil. Microbiology 1986, 132, 3407–3414. [Google Scholar] [CrossRef] [Green Version]
  111. Espinosa-Urgel, M.; Kolter, R.; Ramos, J.L. Root colonization by Pseudomonas putida: Love at first sight. Microbiology 2002, 148, 341–343. [Google Scholar] [CrossRef] [Green Version]
  112. De Weert, S.; Vermeiren, H.; Mulders, I.H.; Kuiper, I.; Hendrickx, N.; Bloemberg, G.V.; Vanderleyden, J.; De Mot, R.; Lugtenberg, B.J. Flagella-driven chemotaxis towards exudate components is an important trait for tomato root colonization by Pseudomonas fluorescens. Mol. Plant-Microbe Interact. 2002, 15, 1173–1180. [Google Scholar] [CrossRef] [Green Version]
  113. Harshey, R.M. Bacterial motility on a surface: Many ways to a common goal. Annu. Rev. Microbiol. 2003, 57, 249–273. [Google Scholar] [CrossRef]
  114. Kearns, D.B. A field guide to bacterial swarming motility. Nat. Rev. Microbiol. 2010, 8, 634–644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Armitage, J.P. Bacterial tactic responses. In Advances in Microbial Physiology; Elsevier: Amsterdam, The Netherlands, 1999; Volume 41, pp. 229–289. [Google Scholar]
  116. Thormann, K.M.; Paulick, A. Tuning the flagellar motor. Microbiology 2010, 156, 1275–1283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Sowa, Y.; Berry, R.M. Bacterial flagellar motor. Q. Rev. Biophys. 2008, 41, 103–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Arora, S.K.; Ritchings, B.W.; Almira, E.C.; Lory, S.; Ramphal, R. A transcriptional activator, FleQ, regulates mucin adhesion and flagellar gene expression in Pseudomonas aeruginosa in a cascade manner. J. Bacteriol. 1997, 179, 5574–5581. [Google Scholar] [CrossRef] [Green Version]
  119. Dasgupta, N.; Wolfgang, M.C.; Goodman, A.L.; Arora, S.K.; Jyot, J.; Lory, S.; Ramphal, R. A four-tiered transcriptional regulatory circuit controls flagellar biogenesis in Pseudomonas aeruginosa. Mol. Microbiol. 2003, 50, 809–824. [Google Scholar] [CrossRef] [Green Version]
  120. Barahona, E.; Navazo, A.; Garrido-Sanz, D.; Muriel, C.; Martínez-Granero, F.; Redondo-Nieto, M.; Martín, M.; Rivilla, R. Pseudomonas fluorescens F113 can produce a second flagellar apparatus, which is important for plant root colonization. Front. Microbiol. 2016, 7, 1471. [Google Scholar] [CrossRef] [Green Version]
  121. Navazo, A.; Barahona, E.; Redondo-Nieto, M.; Martínez-Granero, F.; Rivilla, R.; Martín, M. Three independent signalling pathways repress motility in Pseudomonas fluorescens F113. Microb. Biotechnol. 2009, 2, 489–498. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Muriel, C.; Blanco-Romero, E.; Trampari, E.; Arrebola, E.; Durán, D.; Redondo-Nieto, M.; Malone, J.G.; Martín, M.; Rivilla, R. The diguanylate cyclase AdrA regulates flagellar biosynthesis in Pseudomonas fluorescens F113 through SadB. Sci. Rep. 2019, 9, 8096. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Römling, U.; Galperin, M.Y.; Gomelsky, M. Cyclic di-GMP: The first 25 years of a universal bacterial second messenger. Microbiol. Mol. Biol. Rev. 2013, 77, 1–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Povolotsky, T.L.; Hengge, R. ‘Life-style’control networks in Escherichia coli: Signaling by the second messenger c-di-GMP. J. Biotechnol. 2012, 160, 10–16. [Google Scholar] [CrossRef] [PubMed]
  125. Martínez-Granero, F.; Navazo, A.; Barahona, E.; Redondo-Nieto, M.; Rivilla, R.; Martín, M. The Gac-Rsm and SadB signal transduction pathways converge on AlgU to downregulate motility in Pseudomonas fluorescens. PLoS ONE 2012, 7, e31765. [Google Scholar] [CrossRef] [Green Version]
  126. Schubert, M.; Lapouge, K.; Duss, O.; Oberstrass, F.C.; Jelesarov, I.; Haas, D.; Allain, F.H. Molecular basis of messenger RNA recognition by the specific bacterial repressing clamp RsmA/CsrA. Nat. Struct. Mol. Biol. 2007, 14, 807–813. [Google Scholar] [CrossRef]
  127. Martínez-Granero, F.; Redondo-Nieto, M.; Vesga, P.; Martín, M.; Rivilla, R. AmrZ is a global transcriptional regulator implicated in iron uptake and environmental adaption in P. fluorescens F113. BMC Genom. 2014, 15, 237. [Google Scholar] [CrossRef] [Green Version]
  128. Muriel, C.; Arrebola, E.; Redondo-Nieto, M.; Martínez-Granero, F.; Jalvo, B.; Pfeilmeier, S.; Blanco-Romero, E.; Baena, I.; Malone, J.G.; Rivilla, R.; et al. AmrZ is a major determinant of c-di-GMP levels in Pseudomonas fluorescens F113. Sci. Rep. 2018, 8, 1979. [Google Scholar] [CrossRef] [Green Version]
  129. Dasgupta, N.; Ferrell, E.P.; Kanack, K.J.; West, S.E.; Ramphal, R. fleQ, the gene encoding the major flagellar regulator of Pseudomonas aeruginosa, is σ70 dependent and is downregulated by Vfr, a homolog of Escherichia coli cyclic AMP receptor protein. J. Bacteriol. 2002, 184, 5240–5250. [Google Scholar] [CrossRef] [Green Version]
  130. Martínez-Granero, F.; Navazo, A.; Barahona, E.; Redondo-Nieto, M.; De Heredia, E.G.; Baena, I.; Martín-Martín, I.; Rivilla, R.; Martín, M. Identification of flgZ as a flagellar gene encoding a PilZ domain protein that regulates swimming motility and biofilm formation in Pseudomonas. PLoS ONE 2014, 9, e87608. [Google Scholar] [CrossRef] [Green Version]
  131. Matilla, M.A.; Krell, T. The effect of bacterial chemotaxis on host infection and pathogenicity. FEMS Microbiol. Rev. 2018, 42, fux052. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Hess, J.F.; Oosawa, K.; Matsumura, P.; Simon, M.I. Protein phosphorylation is involved in bacterial chemotaxis. Proc. Natl. Acad. Sci. USA 1987, 84, 7609–7613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Scharf, B.E.; Hynes, M.F.; Alexandre, G.M. Chemotaxis signaling systems in model beneficial plant–bacteria associations. Plant Mol. Biol. 2016, 90, 549–559. [Google Scholar] [CrossRef] [PubMed]
  134. Oku, S.; Komatsu, A.; Tajima, T.; Nakashimada, Y.; Kato, J. Identification of chemotaxis sensory proteins for amino acids in Pseudomonas fluorescens Pf0-1 and their involvement in chemotaxis to tomato root exudate and root colonization. Microbes Environ. 2012, 27, 462–469. [Google Scholar] [CrossRef] [Green Version]
  135. Oku, S.; Komatsu, A.; Nakashimada, Y.; Tajima, T.; Kato, J. Identification of Pseudomonas fluorescens chemotaxis sensory proteins for malate, succinate, and fumarate, and their involvement in root colonization. Microbes Environ. 2014, 29, 413–419. [Google Scholar] [CrossRef] [Green Version]
  136. Schweinitzer, T.; Josenhans, C. Bacterial energy taxis: A global strategy? Arch. Microbiol. 2010, 192, 507–520. [Google Scholar] [CrossRef] [Green Version]
  137. Costerton, J. The role of bacterial exopolysaccharides in nature and disease. J. Ind. Microbiol. Biotechnol. 1999, 22, 551–563. [Google Scholar] [CrossRef]
  138. Høiby, N. A personal history of research on microbial biofilms and biofilm infections. Pathog. Dis. 2014, 70, 205–211. [Google Scholar] [CrossRef] [Green Version]
  139. Flemming, H.-C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice, S.A.; Kjelleberg, S. Biofilms: An emergent form of bacterial life. Nat. Rev. Microbiol. 2016, 14, 563. [Google Scholar] [CrossRef]
  140. Hall-Stoodley, L.; Costerton, J.W.; Stoodley, P. Bacterial biofilms: From the natural environment to infectious diseases. Nat. Rev. Microbiol. 2004, 2, 95–108. [Google Scholar] [CrossRef]
  141. Rudrappa, T.; Biedrzycki, M.L.; Bais, H.P. Causes and consequences of plant-associated biofilms. FEMS Microbiol. Ecol. 2008, 64, 153–166. [Google Scholar] [CrossRef]
  142. Karygianni, L.; Ren, Z.; Koo, H.; Thurnheer, T. Biofilm Matrixome: Extracellular Components in Structured Microbial Communities. Trends Microbiol. 2020, 28, 668–681. [Google Scholar] [CrossRef]
  143. Flemming, H.-C.; Wingender, J. The biofilm matrix. Nat. Rev. Microbiol. 2010, 8, 623–633. [Google Scholar] [CrossRef] [PubMed]
  144. Dragoš, A.; Kovács, Á.T. The peculiar functions of the bacterial extracellular matrix. Trends Microbiol. 2017, 25, 257–266. [Google Scholar] [CrossRef] [PubMed]
  145. Mann, E.E.; Wozniak, D.J. Pseudomonas biofilm matrix composition and niche biology. FEMS Microbiol. Rev. 2012, 36, 893–916. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Blanco-Romero, E.; Garrido-Sanz, D.; Rivilla, R.; Redondo-Nieto, M.; Martin, M. In Silico Characterization and Phylogenetic Distribution of Extracellular Matrix Components in the Model Rhizobacteria Pseudomonas fluorescens F113 and Other Pseudomonads. Microorganisms 2020, 8, 1740. [Google Scholar] [CrossRef]
  147. Ha, D.-G.; O’Toole, G.A. c-di-GMP and its effects on biofilm formation and dispersion: A Pseudomonas aeruginosa review. Microb. Biofilms 2015, 3, 301–317. [Google Scholar] [CrossRef] [Green Version]
  148. Fazli, M.; Almblad, H.; Rybtke, M.L.; Givskov, M.; Eberl, L.; Tolker-Nielsen, T. Regulation of biofilm formation in Pseudomonas and Burkholderia species. Environ. Microbiol. 2014, 16, 1961–1981. [Google Scholar] [CrossRef]
  149. Davies, J.A.; Harrison, J.J.; Marques, L.L.; Foglia, G.R.; Stremick, C.A.; Storey, D.G.; Turner, R.J.; Olson, M.E.; Ceri, H. The GacS sensor kinase controls phenotypic reversion of small colony variants isolated from biofilms of Pseudomonas aeruginosa PA14. FEMS Microbiol. Ecol. 2007, 59, 32–46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. O’Toole, G.A.; Kolter, R. Flagellar and twitching motility are necessary for Pseudomonas aeruginosa biofilm development. Mol. Microbiol. 1998, 30, 295–304. [Google Scholar] [CrossRef]
  151. Hickman, J.W.; Harwood, C.S. Identification of FleQ from Pseudomonas aeruginosa as a c-di-GMP-responsive transcription factor. Mol. Microbiol. 2008, 69, 376–389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Martínez-Gil, M.; Ramos-González, M.I.; Espinosa-Urgel, M. Roles of cyclic di-GMP and the Gac system in transcriptional control of the genes coding for the Pseudomonas putida adhesins LapA and LapF. J. Bacteriol. 2014, 196, 1484–1495. [Google Scholar] [CrossRef] [Green Version]
  153. Matsuyama, B.Y.; Krasteva, P.V.; Baraquet, C.; Harwood, C.S.; Sondermann, H.; Navarro, M.V. Mechanistic insights into c-di-GMP–dependent control of the biofilm regulator FleQ from Pseudomonas aeruginosa. Proc. Natl. Acad. Sci. USA 2016, 113, E209–E218. [Google Scholar] [CrossRef] [Green Version]
  154. Baraquet, C.; Murakami, K.; Parsek, M.R.; Harwood, C.S. The FleQ protein from Pseudomonas aeruginosa functions as both a repressor and an activator to control gene expression from the pel operon promoter in response to c-di-GMP. Nucleic Acids Res. 2012, 40, 7207–7218. [Google Scholar] [CrossRef] [PubMed]
  155. Navarro, M.V.; Newell, P.D.; Krasteva, P.V.; Chatterjee, D.; Madden, D.R.; O’Toole, G.A.; Sondermann, H. Structural basis for c-di-GMP-mediated inside-out signaling controlling periplasmic proteolysis. PLoS Biol. 2011, 9, e1000588. [Google Scholar] [CrossRef]
  156. Newell, P.D.; Yoshioka, S.; Hvorecny, K.L.; Monds, R.D.; O’Toole, G.A. Systematic analysis of diguanylate cyclases that promote biofilm formation by Pseudomonas fluorescens Pf0-1. J. Bacteriol. 2011, 193, 4685–4698. [Google Scholar] [CrossRef] [Green Version]
  157. Monds, R.D.; Newell, P.D.; Gross, R.H.; O’Toole, G.A. Phosphate-dependent modulation of c-di-GMP levels regulates Pseudomonas fluorescens Pf0-1 biofilm formation by controlling secretion of the adhesin LapA. Mol. Microbiol. 2007, 63, 656–679. [Google Scholar] [CrossRef]
  158. López-Sánchez, A.; Jiménez-Fernández, A.; Calero, P.; Gallego, L.D.; Govantes, F. New methods for the isolation and characterization of biofilm-persistent mutants in Pseudomonas putida. Environ. Microbiol. Rep. 2013, 5, 679–685. [Google Scholar]
  159. Valentini, M.; Filloux, A. Biofilms and cyclic di-GMP (c-di-GMP) signaling: Lessons from Pseudomonas aeruginosa and other bacteria. J. Biol. Chem. 2016, 291, 12547–12555. [Google Scholar] [CrossRef] [Green Version]
  160. Liang, F.; Zhang, B.; Yang, Q.; Zhang, Y.; Zheng, D.; Zhang, L.-q.; Yan, Q.; Wu, X. Cyclic-di-GMP Regulates the Quorum-Sensing System and Biocontrol Activity of Pseudomonas fluorescens 2P24 through the RsmA and RsmE Proteins. Appl. Environ. Microbiol. 2020, 86, e02016–e02020. [Google Scholar] [CrossRef] [PubMed]
  161. Flores-Bautista, E.; Hernández-Guerrero, R.; Huerta-Saquero, A.; Tenorio-Salgado, S.; Rivera-Gómez, N.; Romero, A.; Ibarra, J.A.; Pérez-Rueda, E. Deciphering the functional diversity of DNA-binding transcription factors in Bacteria and Archaea organisms. PLoS ONE 2020, 15, e0237135. [Google Scholar] [CrossRef]
  162. Pérez-Rueda, E.; Hernández-Guerrero, R.; Martínez-Núñez, M.A.; Armenta-Medina, D.; Sánchez, I.; Ibarra, J.A. Abundance, diversity and domain architecture variability in prokaryotic DNA-binding transcription factors. PLoS ONE 2018, 13, e0195332. [Google Scholar] [CrossRef] [Green Version]
  163. Sánchez, I.; Hernández-Guerrero, R.; Méndez-Monroy, P.E.; Martínez-Núñez, M.A.; Ibarra, J.A.; Pérez-Rueda, E. Evaluation of the abundance of DNA-binding transcription factors in prokaryotes. Genes 2020, 11, 52. [Google Scholar] [CrossRef] [Green Version]
  164. Fan, L.; Wang, T.; Hua, C.; Sun, W.; Li, X.; Grunwald, L.; Liu, J.; Wu, N.; Shao, X.; Yin, Y. A compendium of DNA-binding specificities of transcription factors in Pseudomonas syringae. Nat. Commun. 2020, 11, 4947. [Google Scholar] [CrossRef] [PubMed]
  165. Baynham, P.J.; Brown, A.L.; Hall, L.L.; Wozniak, D.J. Pseudomonas aeruginosa AlgZ, a ribbon–helix–helix DNA-binding protein, is essential for alginate synthesis and algD transcriptional activation. Mol. Microbiol. 1999, 33, 1069–1080. [Google Scholar] [CrossRef]
  166. Waligora, E.A.; Ramsey, D.M.; Pryor, E.E.; Lu, H.; Hollis, T.; Sloan, G.P.; Deora, R.; Wozniak, D.J. AmrZ beta-sheet residues are essential for DNA binding and transcriptional control of Pseudomonas aeruginosa virulence genes. J. Bacteriol. 2010, 192, 5390–5401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Jones, C.J.; Newsom, D.; Kelly, B.; Irie, Y.; Jennings, L.K.; Xu, B.; Limoli, D.H.; Harrison, J.J.; Parsek, M.R.; White, P. ChIP-Seq and RNA-Seq reveal an AmrZ-mediated mechanism for cyclic di-GMP synthesis and biofilm development by Pseudomonas aeruginosa. PLoS Pathog. 2014, 10, e1003984. [Google Scholar] [CrossRef]
  168. Wozniak, D.J.; Sprinkle, A.B.; Baynham, P.J. Control of Pseudomonas aeruginosa algZ expression by the alternative sigma factor AlgT. J. Bacteriol. 2003, 185, 7297–7300. [Google Scholar] [CrossRef] [Green Version]
  169. Baynham, P.J.; Wozniak, D.J. Identification and characterization of AlgZ, an AlgT-dependent DNA-binding protein required for Pseudomonas aeruginosa algD transcription. Mol. Microbiol. 1996, 22, 97–108. [Google Scholar] [CrossRef] [PubMed]
  170. Wozniak, D.J.; Ohman, D. Transcriptional analysis of the Pseudomonas aeruginosa genes algR, algB, and algD reveals a hierarchy of alginate gene expression which is modulated by algT. J. Bacteriol. 1994, 176, 6007–6014. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  171. Ramsey, D.M.; Wozniak, D.J. Understanding the control of Pseudomonas aeruginosa alginate synthesis and the prospects for management of chronic infections in cystic fibrosis. Mol. Microbiol. 2005, 56, 309–322. [Google Scholar] [CrossRef]
  172. Xu, B.; Ju, Y.; Soukup, R.J.; Ramsey, D.M.; Fishel, R.; Wysocki, V.H.; Wozniak, D.J. The Pseudomonas aeruginosa AmrZ C-terminal domain mediates tetramerization and is required for its activator and repressor functions. Environ. Microbiol. Rep. 2016, 8, 85–90. [Google Scholar] [CrossRef] [Green Version]
  173. Jones, C.J.; Ryder, C.R.; Mann, E.E.; Wozniak, D.J. AmrZ modulates Pseudomonas aeruginosa biofilm architecture by directly repressing transcription of the psl operon. J. Bacteriol. 2013, 195, 1637–1644. [Google Scholar] [CrossRef] [Green Version]
  174. Petrova, O.E.; Cherny, K.E.; Sauer, K. The Pseudomonas aeruginosa diguanylate cyclase GcbA, a homolog of P. fluorescens GcbA, promotes initial attachment to surfaces, but not biofilm formation, via regulation of motility. J. Bacteriol. 2014, 196, 2827–2841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Hou, L.; Debru, A.; Chen, Q.; Bao, Q.; Li, K. AmrZ Regulates Swarming Motility through c-di-GMP Dependent Motility Inhibition and Controlling Pel Polysaccharide Production in Pseudomonas aeruginosa PA14. Front. Microbiol. 2019, 10, 1847. [Google Scholar] [CrossRef] [Green Version]
  176. Garrett, E.S.; Perlegas, D.; Wozniak, D.J. Negative control of flagellum synthesis in Pseudomonas aeruginosa is modulated by the alternative sigma factor AlgT (AlgU). J. Bacteriol. 1999, 181, 7401–7404. [Google Scholar] [CrossRef] [Green Version]
  177. Tart, A.H.; Blanks, M.J.; Wozniak, D.J. The AlgT-dependent transcriptional regulator AmrZ (AlgZ) inhibits flagellum biosynthesis in mucoid, nonmotile Pseudomonas aeruginosa cystic fibrosis isolates. J. Bacteriol. 2006, 188, 6483–6489. [Google Scholar] [CrossRef] [Green Version]
  178. Xu, A.; Zhang, M.; Du, W.; Wang, D.; Ma, L.Z. A molecular mechanism for how sigma factor AlgT and transcriptional regulator AmrZ inhibit twitching motility in Pseudomonas aeruginosa. Environ. Microbiol. 2020, 23, 572–587. [Google Scholar] [CrossRef]
  179. Baynham, P.J.; Ramsey, D.M.; Gvozdyev, B.V.; Cordonnier, E.M.; Wozniak, D.J. The Pseudomonas aeruginosa ribbon-helix-helix DNA-binding protein AlgZ (AmrZ) controls twitching motility and biogenesis of type IV pili. J. Bacteriol. 2006, 188, 132–140. [Google Scholar] [CrossRef] [Green Version]
  180. Pryor, E.E., Jr.; Waligora, E.A.; Xu, B.; Dellos-Nolan, S.; Wozniak, D.J.; Hollis, T. The transcription factor AmrZ utilizes multiple DNA binding modes to recognize activator and repressor sequences of Pseudomonas aeruginosa virulence genes. PLoS Pathog. 2012, 8, e1002648. [Google Scholar] [CrossRef] [Green Version]
  181. Huang, H.; Shao, X.; Xie, Y.; Wang, T.; Zhang, Y.; Wang, X.; Deng, X. An integrated genomic regulatory network of virulence-related transcriptional factors in Pseudomonas aeruginosa. Nat. Commun. 2019, 10, 2931. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Allsopp, L.P.; Wood, T.E.; Howard, S.A.; Maggiorelli, F.; Nolan, L.M.; Wettstadt, S.; Filloux, A. RsmA and AmrZ orchestrate the assembly of all three type VI secretion systems in Pseudomonas aeruginosa. Proc. Natl. Acad. Sci. USA 2017, 114, 7707–7712. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Prada-Ramírez, H.A.; Pérez-Mendoza, D.; Felipe, A.; Martínez-Granero, F.; Rivilla, R.; Sanjuán, J.; Gallegos, M.T. AmrZ regulates cellulose production in Pseudomonas syringae pv. tomato DC 3000. Mol. Microbiol. 2016, 99, 960–977. [Google Scholar]
  184. Pérez-Mendoza, D.; Felipe, A.; Ferreiro, M.D.; Sanjuán, J.; Gallegos, M.T. AmrZ and FleQ Co-regulate cellulose production in Pseudomonas syringae pv. tomato DC3000. Front. Microbiol. 2019, 10, 746. [Google Scholar] [CrossRef] [Green Version]
  185. Baltrus, D.A.; Dougherty, K.; Díaz, B.; Murillo, R. Evolutionary Plasticity of AmrZ Regulation in Pseudomonas. mSphere 2018, 3, e00132-18. [Google Scholar] [CrossRef] [Green Version]
  186. Liu, H.; Yan, H.; Xiao, Y.; Nie, H.; Huang, Q.; Chen, W. The exopolysaccharide gene cluster pea is transcriptionally controlled by RpoS and repressed by AmrZ in Pseudomonas putida KT2440. Microbiol. Res. 2019, 218, 1–11. [Google Scholar] [CrossRef]
  187. Blanco-Romero, E.; Garrido-Sanz, D.; Duran, D.; Rivilla, R.; Redondo-Nieto, M.; Martin, M. Regulation of extracellular matrix components by AmrZ is mediated by c-di-GMP in Pseudomonas ogarae F113. Sci. Rep. 2022, 12, 11914. [Google Scholar] [CrossRef]
  188. Xu, B.; Wozniak, D.J. Development of a novel method for analyzing Pseudomonas aeruginosa twitching motility and its application to define the AmrZ regulon. PLoS ONE 2015, 10, e0136426. [Google Scholar] [CrossRef] [Green Version]
  189. Jyot, J.; Dasgupta, N.; Ramphal, R. FleQ, the major flagellar gene regulator in Pseudomonas aeruginosa, binds to enhancer sites located either upstream or atypically downstream of the RpoN binding site. J. Bacteriol. 2002, 184, 5251–5260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Baraquet, C.; Harwood, C.S. Cyclic diguanosine monophosphate represses bacterial flagella synthesis by interacting with the Walker A motif of the enhancer-binding protein FleQ. Proc. Natl. Acad. Sci. USA 2013, 110, 18478–18483. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Dasgupta, N.; Ramphal, R. Interaction of the Antiactivator FleN with the Transcriptional Activator FleQ Regulates Flagellar Number in Pseudomonas aeruginosa. J. Bacteriol. 2001, 183, 6636–6644. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  192. Su, T.; Liu, S.; Wang, K.; Chi, K.; Zhu, D.; Wei, T.; Huang, Y.; Guo, L.; Hu, W.; Xu, S. The REC domain mediated dimerization is critical for FleQ from Pseudomonas aeruginosa to function as a c-di-GMP receptor and flagella gene regulator. J. Struct. Biol. 2015, 192, 1–13. [Google Scholar] [CrossRef]
  193. Molina-Henares, M.A.; Ramos-González, M.I.; Daddaoua, A.; Fernández-Escamilla, A.M.; Espinosa-Urgel, M. FleQ of Pseudomonas putida KT2440 is a multimeric cyclic diguanylate binding protein that differentially regulates expression of biofilm matrix components. Res. Microbiol. 2017, 168, 36–45. [Google Scholar] [CrossRef]
  194. Marshall, B.; Robleto, E.A.; Wetzler, R.; Kulle, P.; Casaz, P.; Levy, S.B. The adnA transcriptional factor affects persistence and spread of Pseudomonas fluorescens under natural field conditions. Appl. Environ. Microbiol. 2001, 67, 852–857. [Google Scholar] [CrossRef] [Green Version]
  195. Casaz, P.; Happel, A.; Keithan, J.; Read, D.L.; Strain, S.R.; Levy, S.B. The Pseudomonas fluorescens transcription activator AdnA is required for adhesion and motility. Microbiology 2001, 147, 355–361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Mastropaolo, M.D.; Silby, M.W.; Nicoll, J.S.; Levy, S.B. Novel genes involved in Pseudomonas fluorescens Pf0-1 motility and biofilm formation. Appl. Environ. Microbiol. 2012, 78, 4318–4329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Baraquet, C.; Harwood, C.S. FleQ DNA binding consensus sequence revealed by studies of FleQ-dependent regulation of biofilm gene expression in Pseudomonas aeruginosa. J. Bacteriol. 2016, 198, 178–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Jiménez-Fernández, A.; López-Sánchez, A.; Jiménez-Díaz, L.; Navarrete, B.; Calero, P.; Platero, A.I.; Govantes, F. Complex interplay between FleQ, cyclic diguanylate and multiple σ factors coordinately regulates flagellar motility and biofilm development in Pseudomonas putida. PLoS ONE 2016, 11, e0163142. [Google Scholar] [CrossRef] [Green Version]
  199. Xiao, Y.; Nie, H.; Liu, H.; Luo, X.; Chen, W.; Huang, Q. C-di-GMP regulates the expression of lapA and bcs operons via FleQ in Pseudomonas putida KT2440. Environ. Microbiol. Rep. 2016, 8, 659–666. [Google Scholar] [CrossRef]
  200. Wang, Y.; Li, Y.; Wang, J.; Wang, X. FleQ regulates both the type VI secretion system and flagella in Pseudomonas putida. Biotechnol. Appl. Biochem. 2018, 65, 419–427. [Google Scholar] [CrossRef]
  201. Navarrete, B.; Leal-Morales, A.; Serrano-Ron, L.; Sarrió, M.; Jiménez-Fernández, A.; Jiménez-Díaz, L.; López-Sánchez, A.; Govantes, F. Transcriptional organization, regulation and functional analysis of flhF and fleN in Pseudomonas putida. PLoS ONE 2019, 14, e0214166. [Google Scholar] [CrossRef] [Green Version]
  202. Giddens, S.R.; Jackson, R.W.; Moon, C.D.; Jacobs, M.A.; Zhang, X.-X.; Gehrig, S.M.; Rainey, P.B. Mutational activation of niche-specific genes provides insight into regulatory networks and bacterial function in a complex environment. Proc. Natl. Acad. Sci. USA 2007, 104, 18247–18252. [Google Scholar] [CrossRef] [Green Version]
  203. Nie, H.; Xiao, Y.; Liu, H.; He, J.; Chen, W.; Huang, Q. FleN and FleQ play a synergistic role in regulating lapA and bcs operons in Pseudomonas putida KT2440. Environ. Microbiol. Rep. 2017, 9, 571–580. [Google Scholar] [CrossRef] [PubMed]
  204. Blanco-Romero, E.; Duran, D.; Garrido-Sanz, D.; Rivilla, R.; Martin, M.; Redondo-Nieto, M. Transcriptomic analysis of Pseudomonas ogarae F113 reveals the antagonistic roles of AmrZ and FleQ during rhizosphere adaption. Microb. Genom. 2022, 8, 000750. [Google Scholar] [CrossRef] [PubMed]
  205. Blanco-Romero, E.; Redondo-Nieto, M.; Martinez-Granero, F.; Garrido-Sanz, D.; Ramos-Gonzalez, M.I.; Martin, M.; Rivilla, R. Genome-wide analysis of the FleQ direct regulon in Pseudomonas fluorescens F113 and Pseudomonas putida KT2440. Sci. Rep. 2018, 8, 13145. [Google Scholar] [CrossRef] [PubMed]
  206. Duran, D.; Bernal, P.; Vazquez-Arias, D.; Blanco-Romero, E.; Garrido-Sanz, D.; Redondo-Nieto, M.; Rivilla, R.; Martin, M. Pseudomonas fluorescens F113 type VI secretion systems mediate bacterial killing and adaption to the rhizosphere microbiome. Sci. Rep. 2021, 11, 5772. [Google Scholar] [CrossRef]
Figure 1. Common mechanisms of plant growth-promotion by rhizobacteria (PGPR). Plants can attract beneficial bacteria able to elicit the ISR or promote its growth via the production of exudates that bacteria can use as a carbon and energy source. Plant growth-promoting mechanisms are typically divided into biofertilization, phytostimulation, or biocontrol whether they directly promote plant growth by supporting plant nutrition or modifying the hormonal balance of the plant, or indirectly prevent plant diseases by avoiding the presence of potential pathogens. Most common mechanisms are represented. ACC: 1-aminocyclopropane-1-carboxylate; ISR: induced systemic resistance; KB: ketobutyrate.
Figure 1. Common mechanisms of plant growth-promotion by rhizobacteria (PGPR). Plants can attract beneficial bacteria able to elicit the ISR or promote its growth via the production of exudates that bacteria can use as a carbon and energy source. Plant growth-promoting mechanisms are typically divided into biofertilization, phytostimulation, or biocontrol whether they directly promote plant growth by supporting plant nutrition or modifying the hormonal balance of the plant, or indirectly prevent plant diseases by avoiding the presence of potential pathogens. Most common mechanisms are represented. ACC: 1-aminocyclopropane-1-carboxylate; ISR: induced systemic resistance; KB: ketobutyrate.
Microorganisms 11 01037 g001
Figure 2. Phylogenetic tree of the Pseudomonas genus. Tree inferred by MLSA analysis. The P. fluorescens “complex” includes the subgroups P. protegens, P. chlororaphis, P. corrugata, P. koreensis, P. jessenii, P. mandelii, P. fragi, P. gessardii, and P. fluorescens. From Garrido-Sanz et al., 2016 [40].
Figure 2. Phylogenetic tree of the Pseudomonas genus. Tree inferred by MLSA analysis. The P. fluorescens “complex” includes the subgroups P. protegens, P. chlororaphis, P. corrugata, P. koreensis, P. jessenii, P. mandelii, P. fragi, P. gessardii, and P. fluorescens. From Garrido-Sanz et al., 2016 [40].
Microorganisms 11 01037 g002
Figure 3. Current model for motility environmental regulation in Pseudomonas ogarae F113. Several pathways have been identified in the regulation of motility in this bacterium. Blue circles represent c-di-GMP molecules. Gray dashed lines indicate transcriptional regulation, whereas regular black lines represent other levels of regulation. P depicts phosphate. Ray shapes represent environmental stimuli. Small RNA encoded by rsmXYZ and titrated by RsmAEI is drawn close to the Rsm boxes. C: cytosol; IM: inner membrane; OM: outer membrane; P: periplasm.
Figure 3. Current model for motility environmental regulation in Pseudomonas ogarae F113. Several pathways have been identified in the regulation of motility in this bacterium. Blue circles represent c-di-GMP molecules. Gray dashed lines indicate transcriptional regulation, whereas regular black lines represent other levels of regulation. P depicts phosphate. Ray shapes represent environmental stimuli. Small RNA encoded by rsmXYZ and titrated by RsmAEI is drawn close to the Rsm boxes. C: cytosol; IM: inner membrane; OM: outer membrane; P: periplasm.
Microorganisms 11 01037 g003
Figure 4. Schematic representation of the main components of the extracellular matrix (ECM) in Pseudomonas. ECM components include exopolysaccharides (EPSs), lipopolysaccharide (LPS), proteins, lipids, extracellular nucleic acids such as extracellular DNA (eDNA). C: cytosol; IM: inner membrane; OM: outer membrane; P: periplasm.
Figure 4. Schematic representation of the main components of the extracellular matrix (ECM) in Pseudomonas. ECM components include exopolysaccharides (EPSs), lipopolysaccharide (LPS), proteins, lipids, extracellular nucleic acids such as extracellular DNA (eDNA). C: cytosol; IM: inner membrane; OM: outer membrane; P: periplasm.
Microorganisms 11 01037 g004
Figure 5. AmrZ and FleQ form a central hub for environmental adaption in Pseudomonas ogarae F113. Proposed model of the AmrZ and FleQ interplay in the regulation of traits implicated in environmental adaption. According to this model, AmrZ and FleQ form an oscillator by their mutual transcriptional repression. AmrZ activates EPSs production genes and represses motility and iron homeostasis genes. Conversely, FleQ acts as an activator of motility and expression of iron homeostasis genes and as a repressor of EPSs genes. The second messenger c-di-GMP participates in this circuit since AmrZ activates the expression of diguanylate cyclases and FleQ transcriptional regulation is modulated by c-di-GMP binding.
Figure 5. AmrZ and FleQ form a central hub for environmental adaption in Pseudomonas ogarae F113. Proposed model of the AmrZ and FleQ interplay in the regulation of traits implicated in environmental adaption. According to this model, AmrZ and FleQ form an oscillator by their mutual transcriptional repression. AmrZ activates EPSs production genes and represses motility and iron homeostasis genes. Conversely, FleQ acts as an activator of motility and expression of iron homeostasis genes and as a repressor of EPSs genes. The second messenger c-di-GMP participates in this circuit since AmrZ activates the expression of diguanylate cyclases and FleQ transcriptional regulation is modulated by c-di-GMP binding.
Microorganisms 11 01037 g005
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

Blanco-Romero, E.; Durán, D.; Garrido-Sanz, D.; Redondo-Nieto, M.; Martín, M.; Rivilla, R. Adaption of Pseudomonas ogarae F113 to the Rhizosphere Environment—The AmrZ-FleQ Hub. Microorganisms 2023, 11, 1037. https://doi.org/10.3390/microorganisms11041037

AMA Style

Blanco-Romero E, Durán D, Garrido-Sanz D, Redondo-Nieto M, Martín M, Rivilla R. Adaption of Pseudomonas ogarae F113 to the Rhizosphere Environment—The AmrZ-FleQ Hub. Microorganisms. 2023; 11(4):1037. https://doi.org/10.3390/microorganisms11041037

Chicago/Turabian Style

Blanco-Romero, Esther, David Durán, Daniel Garrido-Sanz, Miguel Redondo-Nieto, Marta Martín, and Rafael Rivilla. 2023. "Adaption of Pseudomonas ogarae F113 to the Rhizosphere Environment—The AmrZ-FleQ Hub" Microorganisms 11, no. 4: 1037. https://doi.org/10.3390/microorganisms11041037

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