Skip to content
BY 4.0 license Open Access Published by De Gruyter October 2, 2020

Cyclic di-GMP signaling controlling the free-living lifestyle of alpha-proteobacterial rhizobia

  • Elizaveta Krol , Simon Schäper and Anke Becker ORCID logo EMAIL logo
From the journal Biological Chemistry

Abstract

Cyclic-di-GMP (c-di-GMP) is a ubiquitous bacterial second messenger which has been associated with a motile to sessile lifestyle switch in many bacteria. Here, we review recent insights into c-di-GMP regulated processes related to environmental adaptations in alphaproteobacterial rhizobia, which are diazotrophic bacteria capable of fixing nitrogen in symbiosis with their leguminous host plants. The review centers on Sinorhizobium meliloti, which in the recent years was intensively studied for its c-di-GMP regulatory network.

Introduction

In their natural environment, bacteria constantly face changes of conditions and need to adapt their behavior in order to maximize propagation and minimize the risk of damage. Signal perception and transmission are key processes for generating physiological, genetic and cellular adaptive responses. In bacteria, various signaling systems are based on direct sensing of the cue by regulatory proteins, for example transcription factors and sensors belonging to two-component regulatory systems (Stock et al. 2000). Other signaling pathways involve second messenger molecules. Nucleotide second messengers include the stringent response alarmones guanosine tetra- and pentaphosphate, and versatile function cyclic mono-, di and trinucleotides (Gomelsky 2011; Gründling and Lee 2016; Krasteva and Sondermann 2017; Severin and Waters 2019).

Bis-(3′,5′)-cyclic dimeric guanosine monophosphate (c-di-GMP) has emerged as one of the most ubiquitous and versatile bacterial second messengers. Activities of c-di-GMP metabolizing enzymes are subject to regulation by internal or environmental factors, whereas binding of the second messenger to the cognate receptors exerts action on the regulatory targets and triggers the physiological response. Computational prediction of c-di-GMP metabolizing proteins as well as experimental evidence revealed their broad distribution among the major phylogenetic branches of bacteria, such as Proteobacteria, Spirochetes, Cyanobacteria, Actinobacteria and Firmicutes (Galperin et al. 2001). Genes encoding these proteins are present in multiple copies per genome and their number can vary significantly even between species from the same genus (Bobrov et al. 2011). Comparative genomic analyses indicated that free-living bacteria with complex environmental lifestyles carry far more c-di-GMP metabolizing enzymes than obligate parasites (Galperin 2005), consistent with an important role of this second messenger in environmental adaptation. Regulation based on c-di-GMP signaling takes place at transcriptional, post-transcriptional and post-translational levels. It has key roles in physiological regulation determining bacterial lifestyle, including regulation of cellular processes such as biofilm formation and dispersal, motility, virulence, cell cycle and differentiation (Jenal et al. 2017). The number of identified targets and mechanisms of c-di-GMP-mediated regulation is constantly increasing.

Biological nitrogen fixation, performed exclusively by bacteria, is a primary source of combined nitrogen in any given ecosystem. It is catalyzed by the nitrogenase, an enzyme that requires microoxic conditions and has a concomitant high energy demand for the N2-fixing reaction. Along with diverse free-living water- and plant-associated soil bacteria, plant endosymbionts are capable of nitrogen fixation. Root nodule symbiosis between soil bacteria from the alpha-proteobacterial Rhizobiales order, collectively called alpha-rhizobia, and leguminous plants accomplishes a direct supply of bioavailable nitrogen to the host. The process begins with a signal exchange between the partners, leading to bacterial chemotaxis towards the plant roots, attachment to the root hair and their entrapment inside the curling root hair tip. Subsequently, an infection thread is formed inside the root hair, in which bacteria proliferate and progress towards the root cortex. In the root cortex, bacteria invade the plant cells and differentiate into nitrogen-fixing forms, called bacteroids. Symbiosis culminates in formation of a novel plant organ, a nitrogen fixing root nodule in which bacteria provide combined nitrogen to the plant (Roy et al. 2020).

During the free-living stage, rhizobia live as soil saprophytes and become exposed to adverse environmental conditions such as temperature, pH and osmotic pressure challenges, nutrient limitation and desiccation. Therefore, alternating phases of a motile lifestyle, characterized by proliferation and spreading in favorable conditions, and a sessile lifestyle, supporting bacterial survival under unfavorable conditions, are likely prevalent in bacterial soil communities. Processes typically controlled by c-di-GMP signaling, such as motility, exopolysaccharide production and surface attachment have been reported to be important for survival of rhizobia in the soil and for competitive establishment of symbiosis (Figure 1) (Caetano-Annolés et al. 1988; Cheng and Walker 1998; Downie 2010; Skorupska et al. 2006). In this review, we summarize the current knowledge on alpha-rhizobial c-di-GMP regulated processes related to environmental adaptation. We focus the review on Sinorhizobium meliloti as it provides the best studied c-di-GMP regulatory network in alpha-rhizobia to date. S. meliloti emerged as a model alpha-rhizobium for studying not only its symbiotic interaction with its host plants from the genus Medicago, but also properties associated with sessile or motile lifestyle (Jones et al. 2007; Janczarek 2011).

Figure 1: Physiological processes, influenced by c-di-GMP-mediated regulation in alpha-rhizobia.1: Sedentary lifestyle under unfavorable conditions. 2: Motile lifestyle, spreading. 3: Motile lifestyle, plant chemotaxis. 4: Symbiosis, potential role in attachment to the root hair. 5: Symbiosis, potential role in infection thread progression (EPS biosynthesis).
Figure 1:

Physiological processes, influenced by c-di-GMP-mediated regulation in alpha-rhizobia.

1: Sedentary lifestyle under unfavorable conditions. 2: Motile lifestyle, spreading. 3: Motile lifestyle, plant chemotaxis. 4: Symbiosis, potential role in attachment to the root hair. 5: Symbiosis, potential role in infection thread progression (EPS biosynthesis).

c-di-GMP metabolic enzymes

In bacteria, c-di-GMP is synthesized from two GTP molecules by diguanylate cyclases (DGCs). Their active site contains a GGD(E)EF signature motif, therefore the corresponding catalytic domains were named GGDEF domains (Figure 2). For c-di-GMP synthesis, a GGDEF domain homodimer is required (Chan et al. 2004; Paul et al. 2007). The glycine residues in the active site are involved in GTP binding, the aspartate/glutamate at the third position is required for the formation of a phosphodiester bond, and the glutamate in the fourth position coordinates an Mg2+ or Mn2+ ion, involved in GTP binding (Chan et al. 2004; Wassmann et al. 2007). In about half of all annotated GGDEF domains, an inhibitory c-di-GMP binding site with an RxxD signature motif is present (Seshasayee et al. 2010). Binding of c-di-GMP at this site prevents homodimerization of GGDEF domains, resulting in feedback inhibition (Chan et al. 2004; Christen et al. 2006).

Figure 2: C-di-GMP mediated regulation in bacteria.Abundance of the second messenger is defined by enzymatic activities of GGDEF, EAL and HD-GYP domains. These are subject to regulation by abiotic factors, small molecules and protein-protein interactions (signals). C-di-GMP binds to receptor molecules, which mediate regulation of the targets, leading to a physiological response. Dashed line represents negative feedback regulation of DGC activity by binding of c-di-GMP at the inhibitory site.
Figure 2:

C-di-GMP mediated regulation in bacteria.

Abundance of the second messenger is defined by enzymatic activities of GGDEF, EAL and HD-GYP domains. These are subject to regulation by abiotic factors, small molecules and protein-protein interactions (signals). C-di-GMP binds to receptor molecules, which mediate regulation of the targets, leading to a physiological response. Dashed line represents negative feedback regulation of DGC activity by binding of c-di-GMP at the inhibitory site.

Hydrolysis of c-di-GMP is catalyzed by EAL and HD-GYP domains, named after their signature active site motifs (Galperin et al. 2001) (Figure 2). Phosphodiesterase (PDE) cleavage of c-di-GMP by EAL domains results in the formation of the linear nucleotide pGpG. The glutamate residue of the active site is directly involved in coordination of a Mg2+ or Mn2+ metal ion required for c-di-GMP binding (Schmidt et al. 2005; Tchigvintsev et al. 2010). The vast majority of EAL domains form homodimers or higher-order oligomers in vitro, and a dimer is the most probable functional unit of the EAL domain in vivo (Barends et al. 2009; Bellini et al. 2017; Sundriyal et al. 2014). Unlike EAL domains, HD-GYP domains degrade c-di-GMP in a one-step reaction that yields two molecules of GMP (Bellini et al. 2014) (Figure 2).

Within a given protein, GGDEF and EAL domains are often arranged in tandem (Römling et al. 2013). This apparent contradiction is resolved by their differential regulation or sequence aberrations in the active site, resulting in conditionally or permanently determined single enzymatic activity. For instance, the enzymatic activity of Agrobacterium tumefaciens GGDEF-EAL domain protein DcpA is shifted from a DGC to a PDE in presence of pteridine reductase PruA (Feirer et al. 2015). In Vibrio parahaemolyticus, dual-function protein ScrC switches from DGC to PDE activity in response to autoinducer molecules at high cell densities (Trimble and McCarter 2011).

The alpha-rhizobial S. meliloti type strain Rm2011 carries 18 intact genes encoding c-di-GMP metabolizing proteins, with six containing only a GGDEF domain, one containing only an EAL domain and eleven tandem-type proteins containing both domains. 10 out of 17 GGDEF domains contain a canonical GG(D/E)EF catalytic site motif and 10 out of 11 EAL domains carry an intact active site (Schäper et al. 2016). This array of c-di-GMP metabolizing proteins is medium-size relative to other nitrogen-fixing alpha-rhizobial species, which contain up to 51 c-di-GMP metabolizing enzymes (Gao et al. 2014) or the enterobacterium Escherichia coli with 29 enzymes (Hengge et al. 2015). No proteins with HD-GYP domains were identified in S. meliloti. Like many c-di-GMP metabolizing proteins in other bacterial species, most of the S. meliloti GGDEF and EAL proteins contain additional cytoplasmic or periplasmic sensory input domains.

An effort towards understanding the roles of different S. meliloti enzymes was made employing gene overexpression coupled with c-di-GMP quantification (Schäper et al. 2016). This analysis identified five DGCs as enzymatically active. These included the cytoplasmic proteins PleD (SMc01370) and BgrR (SMb20447), and inner membrane proteins SMb20523, SMc01464, and SMc03178. Overproduction of active DGCs resulted in elevated cellular c-di-GMP levels, reduced swimming motility, alterations in exopolysaccharide production and increased biofilm formation (Schäper et al. 2016).

REC domains, which are known as receiver part of two-component systems using phosphorylation for signal transmission (Bourret 2010), are frequently encountered in c-di-GMP metabolizing proteins (Römling et al. 2013). Phosphorylation of the REC domain as a cue for DGC activation was demonstrated for PleD from alpha-proteobacterium Caulobacter crescentus and WspR from gamma-proteobacterium Pseudomonas aeruginosa (Hickman et al. 2005; Paul et al. 2007; Wassmann et al. 2007). C. crescentus PleD was activated by beryllium fluoride mimicking phosphorylation, resulting in dimerization and DGC activity (Paul et al. 2007). Similarly to its C. crescentus homolog, S. meliloti PleD contains two REC domains. C. crescentus cell division produces a surface-attached stalked mother cell and a flagellated swarmer daughter cell. Enzymatic activity of C. crescentus PleD is linked to its localization at the old cell pole of the sessile mother cell, which results in higher c-di-GMP content in this cell compared to the motile daughter cell (Christen et al. 2010; Paul et al. 2004). In S. meliloti, PleD transiently localized to the old pole of the daughter cell, equivalent to the C. crescentus motile swarmer cell (Schäper et al. 2016). It remains to be determined if the c-di-GMP content of the two S. meliloti progeny cells differ.

Sinorhizobium meliloti is a rod-shaped bacterium that proliferates by asymmetric cell division and exhibits unipolar cell wall growth at the new cell pole (Brown et al. 2012; Frage et al. 2016; Schäper et al. 2018). RgsP (SMc00074) is an active PDE responsible for about half of the net c-di-GMP degradation (Schäper et al. 2018). RgsP additionally contains PAS (regulatory Per-Arndt-Sim domain) and 7TMR-DISM (seven-transmembrane receptors with diverse intracellular signal modules) domains, and a non-consensus GGDEF domain. Moreover, RgsP is an essential protein, localized to sites of cell wall growth at one cell pole and the septal site. RgsP is involved in interactions with an array of other essential transmembrane or periplasmic proteins and the Tol-Pal system, localized at the new cell pole in both progeny cells and the septal site in a dividing cell (Krol et al. 2020; Schäper et al. 2018) (Figure 3). Although the enzymatic portion of RgsP is not essential for its cell growth-related function, the question if localized PDE activity of RgsP could generate a c-di-GMP pole-to-pole gradient in S. meliloti cells is the subject of future studies.

Figure 3: Regulatory network of c-di-GMP control on exopolysaccharide production and swimming motility in S. meliloti.Red circles: c-di-GMP, gray circles: GTP, black circles: pGpG. Blue boxes with question marks represent yet-unknown c-di-GMP receptors.
Figure 3:

Regulatory network of c-di-GMP control on exopolysaccharide production and swimming motility in S. meliloti.

Red circles: c-di-GMP, gray circles: GTP, black circles: pGpG. Blue boxes with question marks represent yet-unknown c-di-GMP receptors.

The GGDEF-EAL tandem protein SMc03178 is potentially the most active S. meliloti DGC, since its overexpression resulted in the highest c-di-GMP accumulation (Schäper et al. 2016). In addition to a GGDEF domain, it contains an EAL domain, and extracellular sensory CHASE (Cyclases/Histidine kinases Associated Sensory Extracellular) and regulatory PAS domains, implying direct regulation by a yet-unknown external factor and potential dual c-di-GMP synthesis and degradation functionality.

Although GG(D/E)EF is the canonical motif of the catalytic DGC domain, an active DGC with AGDEF motif was described in Vibrio cholerae suggesting that the first position of the motif is less conserved (Hunter et al. 2014). S. meliloti tandem protein BgrR represents another example of an active DGC with AGDEF motif (Schäper et al. 2016). The gene encoding BgrR is the first of the bgrRSTUWV operon (Baena et al. 2019). BgrUVW constitute a putative partner-switching system, composed of transmembrane phosphatase BgrU and protein kinase BgrW, both acting upon BgrV. If dephosphorylated by BgrU, BgrV inhibits DGC activity of BgrR (Figure 3) (Baena et al. 2019). A yet-unknown environmental signal is suggested to modulate the BgrU-mediated BgrV dephosphorylation, leading to changes in BgrR DGC activity. Additionally, BgrS and BgrT might modulate BgrU phosphatase activity via methylation (Baena et al. 2019).

Gene deletion analysis revealed that all S. meliloti GGDEF domain proteins except for RgsP, which has no DGC activity, can be eliminated from the cells (Schäper et al. 2016). The resulting c-di-GMP0 strain lacking 16 GGDEF domain proteins displayed attenuated growth in acidic conditions, however no further free-living or symbiotic defects were found (Schäper et al. 2016). This constitutes an important difference to other model bacteria such as Salmonella Enteritidis and C. crescentus, which showed defects in swimming motility and cell cycle progression in cells lacking DGC enzymes (Abel et al. 2011; Solano et al. 2009).

c-di-GMP receptors

c-di-GMP binding proteins are key components in the delivery of second messenger signals to the cognate targets at transcriptional, post-transcriptional and post-translational levels. Some of these proteins display distinct structural properties, like a PilZ domain (see below) and enzymatically inactive GGDEF and EAL domains, and possess defined c-di-GMP binding sites (Benach et al. 2007; Duerig et al. 2009; Petters et al. 2012; Ramelot et al. 2007). In addition, proteins with diverse non-conserved c-di-GMP binding sites were shown to receive and transmit the second messenger signal (Fang et al. 2014; Gallagher et al. 2020; Wang et al. 2016; Sprecher et al. 2017). Moreover, c-di-GMP can serve as a cofactor, which promotes enzymatic activity of the protein upon binding, in the absence of a specific c-di-GMP binding domain (Steiner et al. 2013).

The PilZ domain, named after the P. aeruginosa type IV pilus control protein, is the first discovered and most commonly known type of c-di-GMP receptors. PilZ domains are composed of approximately 100 amino acids and contain the c-di-GMP binding motifs RXXXR and (D/N)X(S/A)XXG (Benach et al. 2007; Ramelot et al. 2007). The c-di-GMP-regulated function discovered first, cellulose biosynthesis in Acetobacter xylinum, relies on binding of the second messenger to the PilZ domain of cellulose synthase CeSA (Fujiwara et al. 2013; Weinhouse et al. 1997). In addition to being part of multidomain proteins, PilZ domains can exist as stand-alone proteins, tandem dimers or imperfect tandems, consisting of one functional and one non-consensus PilZ domain (Galperin and Chou 2020). An example of an imperfect PilZ domain tandem is E. coli YcgR. Upon c-di-GMP binding, it undergoes structural rearrangement, which promotes its interaction with MotA and FliGMN, resulting in a negative effect on swimming motility (Boehm et al. 2010; Fang and Gomelsky 2010; Hou et al. 2020; Paul et al. 2010).

In the S. meliloti genome, two stand-alone PilZ domain proteins are encoded, SMc00999 and McrA (SMc00507). At elevated c-di-GMP levels, McrA was shown to mediate repression of swimming motility resulting from pleD overexpression (Figure 3), whereas SMc00999 did not mediate any phenotypic change in the conditions tested (Schäper et al. 2016). McrA bound c-di-GMP in vitro and underwent a conformational change upon c-di-GMP binding (Schäper et al. 2016). Factors mediating the repression of swimming motility by McrA remain unknown. Similarly to McrA, stand-alone PilZ domain proteins DgrA and DgrB regulate c-di-GMP-dependent motility in C. crescentus through a yet unknown mechanism (Christen et al. 2007).

An example of a c-di-GMP receptor protein with a non-conserved c-di-GMP binding site is S. meliloti (1 → 3)(1 → 4)-β-D-glucan (mixed linkage β-glucan) biosynthesis glycosyltransferase BgsA displaying similarities to cellulose synthases (Perez-Mendoza et al. 2015). Such cellulose synthases perceive the c-di-GMP signal via PilZ domains (Morgan et al. 2014). In contrast, allosteric activation of BgsA proceeds via c-di-GMP binding to its C-terminal portion, which is non-homologous to PilZ domains (Perez-Mendoza et al. 2017). A 139 amino acid long C-terminal portion of BgsA contains residues R599 and H615 crucial for both c-di-GMP binding and mixed linkage β-glucan biosynthesis in cells with elevated c-di-GMP (Perez-Mendoza et al., 2015, 2017).

c-di-GMP-binding transcription factors are c-di-GMP receptors that mediate second messenger-dependent regulation at the level of gene expression. Unlike PilZ or non-consensus c-di-GMP-metabolizing domains, they do not have a common conserved c-di-GMP binding motif at the amino acid sequence level. Transcriptional activators and repressors of various types like CRP/FNR, TetR-like, XRE, NtrC or FixJ-LuxR-CsgD were reported to regulate gene expression dependent on c-di-GMP (Chin et al. 2010; Hsieh et al. 2018; Li and He 2012; Tschowri et al. 2014).

In S. meliloti, the AraC-type transcription activator CuxR was identified as c-di-GMP receptor (Figure 3) (Schäper et al. 2017). The mode of c-di-GMP-CuxR interaction, involving an RxxxR motif and an additional distal binding site, is reminiscent of that of PilZ domains (Schäper et al. 2017). Binding of a c-di-GMP dimer to CuxR is proposed to promote structural rearrangements in CuxR favoring its dimerization and hence its DNA-binding ability, required for promoter activation (Schäper et al. 2017).

Regulatory network controlling processes, related to environmental adaptation in S. meliloti

Alpha-rhizobia are able to produce a complex array of glucidic molecules, such as exopolysaccharides (EPSs), lipopolysaccharide (LPS), capsular polysaccharide, K-antigen polysaccharide, cyclic glucans, glucomannan and gel-forming polysaccharide (Fraysse et al. 2003; Laus et al. 2006). Rhizobial surface polysaccharides play a role in the infection process and in the free-living state, contribute to nutrient gathering, surface attachment, biofilm formation, and protection against environmental stresses and antimicrobial compounds (Downie 2010; Skorupska et al. 2006). Noteworthy, alpha-rhizobia from the genus Rhizobium are able to produce cellulose, which plays an important role in surface attachment and biofilm formation (Perez-Mendoza et al. 2014; Robledo et al. 2012; Laus et al. 2005). Swimming or swarming motility is required for both spreading during the free-living state and efficient establishment of the contact with the plant host at the early stages of symbiotic interaction (Caetano-Anollés et al. 1988). Rhizobia are equipped with quorum sensing (QS) systems that use N-acyl-homoserine lactones (AHLs) as messenger molecules (Marketon et al. 2002). QS regulates polysaccharide production and motility and influences symbiotic interaction (Edwards et al. 2009; Hoang et al. 2004; Sanchez-Contreras et al. 2007).

Sinorhizobium meliloti forms peritrichous flagella enabling swimming motility and chemotactic response during exponential growth as well as chemotaxis towards the plant (Rotter et al. 2006; Schmitt 2002; Sourjik et al. 2000; Webb et al. 2014). Two major EPSs are produced by S. meliloti. EPS I, succinoglycan, is an acidic heteropolymer consisting of repeating octasaccharide subunits (Reinhold et al. 1994). EPS II, galactoglucan, consists of disaccharide repeating units (Her et al. 1990). The ability to produce at least one of the major EPS is crucial for infection thread progression (Battisti et al. 1992; Cheng and Walker 1998; González et al. 1996; Urzainqui and Walker 1992). Furthermore, recent research provided evidence that at elevated c-di-GMP levels, S. meliloti with intact quorum sensing system genes expR, sinI and sinR is able to produce mixed-linkage β-glucan (Pérez-Mendoza et al. 2015).

Moreover, c-di-GMP stimulates production of arabinose-containing polysaccharide (APS) and an adhesion polysaccharide of unknown composition, likely similar to A. tumefaciens unipolar adhesion polysaccharide UPP. These polysaccharides facilitate cell aggregation, surface attachment and biofilm formation, however they are not required for effective symbiotic interaction (Pérez-Mendoza et al. 2015; Schäper et al. 2016; Schäper et al. 2019; Xu et al. 2012).

C-di-GMP-mediated control of S. meliloti EPS biosynthesis and swimming motility is implicated into a complex regulatory network governing these processes (Janczarek 2011; Rotter et al. 2006; Scharf and Schmitt 2002; Sourjik et al. 2000). Of notice, opposing control of EPS I production and swimming motility is exerted by a regulatory system composed of ExoR (periplasmic repressor), ExoS (sensory histidine kinase) and ChvI (transcription regulator) (Wang et al. 2010; Yao et al. 2004). In general, stress factors, such as starvation, generally negatively affect expression of genes controlling swimming motility and simultaneously activate expression of EPS biosynthesis genes (Chao et al. 2005; Domínguez-Ferreras et al. 2006; Hoang et al. 2008; Krol and Becker, 2004, 2011).

The mechanisms of stress response regulation of EPS biosynthesis and swimming motility are only partially known. For example, expression of EPS II biosynthesis genes under phosphate limitation is stimulated by PhoB, the response regulator of the PhoR-PhoB two-component system (Bahlawane et al. 2008a; Krol and Becker 2004). Furthermore, the global regulators MucR and ExpR modulate S. meliloti exopolysaccharide production and swimming motility (Figure 3). MucR is a zinc-finger transcriptional regulator. It is transcriptionally autoregulated and couples both EPS biosynthetic pathways by positive regulation of EPS I production and negative regulation of EPS II production (Bertram-Drogatz et al. 1998; Rüberg et al. 1999). MucR also negatively regulates swimming motility by repressing the transcription of flagellar gene regulator rem (Bahlawane et al. 2008b; Rotter et al. 2006). ExpR is a LuxR-type regulator of the Sin/ExpR QS system. S. meliloti QS AHL molecules are produced by the synthase SinI (Gao et al. 2005; McIntosh et al. 2009). ExpR-AHL regulates various target processes at the transcription level, exerting a positive effect on polysaccharide biosynthesis and a negative effect on flagellar motility (Figure 3) (Charoenpanich et al. 2013; Mueller and González 2011; Zatakia et al. 2014).

c-di-GMP-mediated regulation of motility

Bacterial swimming motility is crucial at the early steps of plant-microbial interaction, which culminate in a physical contact between the partners. Bacteria sense the plant-derived compounds and respond with chemotaxis. Motility and chemotaxis quantitatively affect important traits that facilitate the initial contact and adsorption of symbiotic rhizobia to the host root surface (Caetano-Anollés et al. 1988).

Regulation of bacterial motility by c-di-GMP is complex. Low levels proved to be beneficial for this trait, whereas either artificial accumulation or removal of c-di-GMP resulted in motility inhibition (Abel et al. 2011; Bhasme et al. 2020; Pallegar et al. 2020; Yang et al. 2016). The mechanisms of c-di-GMP-mediated control of bacterial motility range from regulation of flagellar gene expression to direct binding to proteins interacting with flagellar components and regulation of chemotaxis (Boehm et al. 2010; Fang and Gomelsky 2010; Hou et al. 2020; Paul et al. 2010; Sun et al. 2019). Noteworthy, not only does c-di-GMP affect swimming motility, but also control of c-di-GMP levels is interwoven with regulation of flagellar motility. In A. tumefaciens, the master activator VisN of flagellar motility genes represses the DGC genes dgcB and dgcC, providing a pathway to inversely correlate flagellar gene expression with c-di-GMP levels (Xu et al. 2013). Upon c-di-GMP-binding, P. aeruginosa PilZ-domain protein MapZ affects chemoreceptor methylation via interaction with chemotaxis methyltransferase CheR1 (Xu et al. 2016). In turn, asymmetrical inheritance of the chemotaxis apparatus after cell division results in differential activation of PDE Pch via the phosphorylation status of CheA (Kulasekara et al. 2013).

C-di-GMP-mediated regulation of motility in S. meliloti shows both similarities and differences to the general paradigm. In contrast to c-di-GMP0 strains of Salmonella Enteritidis and C. crescentus, which show a flagellar motility defect, the S. meliloti c-di-GMP0 strain performs normally in swimming motility assays (Schäper et al. 2016). Unlike E. coli and P. aeruginosa, whose flagellar gene expression is repressed in c-di-GMP-dependent manner (Hickman and Harwood 2008; Nieto et al. 2019), no effect on flagellar gene transcript abundances was observed in S. meliloti overproducing DGC PleD (Schäper et al. 2017). Artificial increase of c-di-GMP levels upon DGC overproduction in S. meliloti and the two related alpha-rhizobia Rhizobium etli and Rhizobium leguminosarum resulted in inhibition of swimming motility, consistent with the general paradigm (Schäper et al. 2016; Perez-Mendoza et al. 2014). One possible pathway of c-di-GMP control of S. meliloti swimming motility is its binding to flagellar export ATPase FliI (Trampari et al. 2015). Moreover, PilZ domain protein McrA mediates inhibition of swimming motility upon overexpression of pleD (Figure 3).

Regulatory connections to QS

Quorum sensing (QS) is a powerful mechanism of global gene regulation in response to changes in population density which relies on secretion and uptake of small signal molecules. Functional QS is advantageous for bacterial performance in both pathogenic and symbiotic microbe-plant interactions (Calatrava-Morales et al. 2018; Lowe-Power et al. 2018; Sanchez-Contreras et al. 2007). Although S. meliloti QS is not essential for a successful symbiotic interaction, strains lacking AHL synthase SinI showed attenuated symbiotic performance (Gurich and Gonzalez 2009; Marketon et al. 2002).

Reciprocal interaction between QS and c-di-GMP regulatory circuits are well documented, with a general trend of dominant negative control of QS by c-di-GMP (Hochstrasser et al. 2019; Lin Chua et al. 2017; Waldron et al. 2019; Yang et al. 2017). In S. meliloti, elevated levels of c-di-GMP repressed transcription of AHL synthase enzyme sinI and lowered abundance of QS signal AHL molecules in the culture medium (Schäper et al. 2016) (Figure 3). However, absence of QS regulator ExpR had no effect on S. meliloti intracellular c-di-GMP content (Schäper et al. 2016). Inhibition of the QS system by c-di-GMP provides an additional level of control on swimming motility and polysaccharide production by this second messenger.

Regulation of surface attachment and biofilm formation by c-di-GMP

A positive regulatory effect of c-di-GMP on bacterial surface attachment and formation of higher-order three-dimensional structures designated biofilms is the most prominent and best understood function of this second messenger in promoting a sedentary lifestyle (for recent reviews, see Hengge 2020; Maunders and Welch 2017; Purcell and Tamayo 2016). c-di-GMP-dependent regulation of biofilm formation takes place in a non-uniform distribution across the biofilm and follows a defined spatiotemporal pattern during colony maturation (Nair et al. 2017). Biofilm formation is associated with survival in free-living conditions (Rinaudi and Giordano 2010) and diverse pathogenic interactions (Kumar et al. 2017).

Biofilms are composed of bacterial cells encased in extracellular matrix, which can contain cellulose and other exopolysaccharides, adhesive pili, non-fimbrial adhesins and extracellular DNA. Its composition defines multicellular morphotypes and cooperative bacterial movements (Gao et al. 2012; Serra et al. 2013; Whitchurch et al. 2002; Zogaj et al. 2001). c-di-GMP dependent regulation of polysaccharide biosynthesis is important for extracellular matrix formation (Liang 2015; McDougald et al. 2012).

In S. meliloti and the related plant pathogen A. tumefaciens, biofilm formation is enhanced at high c-di-GMP levels. One of the underlying molecular mechanisms comprises activation of A. tumefaciens adhesion polysaccharide UPP production in a yet-unknown manner (Xu et al. 2012). In A. tumefaciens, UPP biosynthesis is controlled by an uppABCDEF gene cluster encoding two glycosyltransferases, two polysaccharide transport proteins and two hypothetical proteins (Xu et al. 2012). The S. meliloti gene cluster SMc01796-SMc01790 is highly similar to uppABCDEF, however it contains an additional glycosyltransferase gene (Schäper et al. 2016). In both A. tumefaciens and S. meliloti, removal of EPS I due to knockout of exoY increased the biofilm forming capacity of cultures of strains with elevated c-di-GMP levels, indicating that EPS I diminishes cell aggregation and surface attachment (Schäper et al. 2016; Xu et al. 2013). Consistent with this finding, the global regulators ExpR and MucR, stimulating EPS I production, negatively affected biofilm formation by S. meliloti (Schäper et al. 2016). The molecular mechanism of c-di-GMP action on UPP biosynthesis and the UPP sugar composition are subjects of future studies.

Biosynthesis of mixed-linkage β-glucan in strains with elevated levels of c-di-GMP requires c-di-GMP binding glycosyltransferase BgsA and putative export protein BgsB (Figure 3). It contributes to cell aggregation and biofilm formation and is required for efficient attachment to plant roots (Perez-Mendoza et al., 2015, 2017).

Furthermore, at high c-di-GMP levels, S. meliloti produced arabinose-containing polysaccharide (APS) (Figure 3). Its biosynthesis is controlled by the uxs1-uxe-apsS-apsH1-apsE-apsH2 operon (Schäper et al. 2019). Expression of the uxs1-aspH2 operon is repressed by MucR and is activated by the cognate transcription regulator CuxR, encoded upstream of the uxs1-apsH2 operon (Schäper et al. 2019). CuxR requires c-di-GMP for DNA binding and therefore transcription of the operon is activated at high c-di-GMP levels (Schäper et al. 2017). The gene products of the first two genes in the operon, Uxs1 and Uxe, act as UDP-xylose synthase and UDP-xylose 4-epimerase, respectively (Gu et al. 2011). ApsS is a putative glycosyl transferase, ApsE is a putative endoglycanase and the two remaining proteins are of unknown function. APS production significantly increased biofilm formation in strains that lacked the ability to produce the putative polar adhesion polysaccharide, whose biosynthesis is controlled by the SMc01796-SMc01790 gene cluster (Schäper et al. 2019).

c-di-GMP-mediated regulation of symbiosis-relevant polysaccharide production

Artificial increase in c-di-GMP levels upon DGC overproduction in S. meliloti, R. etli and R. leguminosarum affected biosynthesis of EPSs. Cellulose production was enhanced in R. etli and R. leguminosarum (Perez-Mendoza et al. 2014). In S. meliloti Rm2011, not containing a functional expR gene, EPS I production was increased at high c-di-GMP levels (Schäper et al. 2016). This was not observed in S. meliloti Rm8530 carrying functional expR (Perez-Mendoza et al. 2015). These EPSs are required at the early stages of plant infection (Muszyński et al. 2016; Nieahus and Becker 1998; Staehelin et al. 2006).

At high c-di-GMP levels, transcription of S. meliloti exoY, encoding the first, production rate-determining enzyme of the EPS I biosynthesis pathway, was not altered, which implies a posttranscriptional control by c-di-GMP (Schäper et al. 2016). In contrast, production of EPS II, stimulated by phosphate starvation, was negatively affected by elevated c-di-GMP levels. This correlated with repression of wgeA, a gene from the EPS II biosynthesis cluster, indicating that c-di-GMP repressed EPS II biosynthesis at transcription level (Schäper et al. 2016). Moreover, negative regulation of AHL production by c-di-GMP constitutes an additional pathway of EPS II biosynthesis regulation via QS (Figure 3).

Interestingly, at elevated c-di-GMP levels, activation of uxs1-uxe-apsS-apsH1-apsE-apsH2 operon transcription can have an indirect positive effect on EPS I and EPS II biosynthesis. The UDP-sugar epimerase Uxe, initially reported to perform conversion of UDP-xylose and UDP-arabinose, appeared to be also able to convert UDP-glucose and UDP-galactose. UDP-galactose, a component of LPS and both symbiotically relevant EPS, is normally synthesized by S. meliloti ExoB, encoded within the EPS I biosynthesis gene cluster, and previously the only known UDP-glucose 4-epimerase in this S. meliloti strain. Activating uxs1-apsH2 operon expression complemented the symbiotic defect conferred by an exoB mutation (Schäper et al. 2019). Thus, c-di-GMP-mediated regulation can promote the symbiotically important EPS production by increasing nucleotide sugar precursor abundance.

Effect of c-di-GMP on the symbiotic interaction

In plant pathogenic bacteria, c-di-GMP regulates virulence traits in a positive as well as a negative fashion. In the soft rot pathogen Dickeya dadantii, production of plant cell wall lytic enzymes is repressed by c-di-GMP (Yuan et al. 2020). In the phytopathogen Xanthomonas campestris, DGCs were reported to inhibit virulence, whereas PDEs increased it, suggesting that low c-di-GMP levels are beneficial for plant infection (Su et al. 2016; Xue et al. 2018; Yang et al. 2016). In Erwinia amylovora, c-di-GMP promotes virulence determinants such as amylovoran production and biofilm formation. However, motility and the type III secretion system are expressed and functional when cyclic di-GMP is absent (Kharadi et al. 2018). Moreover, c-di-GMP-mediated modulation of protein secretion systems of virtually any known type was reported in plant pathogenic or beneficial bacteria (Lopez-Baena et al. 2019).

In alpha-rhizobia, only adverse effects of elevated c-di-GMP levels on symbiotic performance were described. Removal of c-di-GMP did not affect the symbiotic efficiency of S. meliloti (Schäper et al. 2016). In contrast, strongly increased c-di-GMP levels in the related rhizobia R. etli and R. leguminosarum enhanced bacterial attachment to the plant roots, at the same time reducing the ability of the bacteria to promote growth of the respective host plants, despite normal number, size and visual appearance of the nodules (Perez-Mendoza et al. 2014). The adverse effect of increased c-di-GMP levels during symbiotic interaction was corroborated by observation of a massive loss of the plasmid conferring high DGC activity during the symbiotic interaction (Perez-Mendoza et al. 2014). S. meliloti ML β-glucan, which is produced at high c-di-GMP levels, was required for increased bacterial attachment to the host plant roots (Perez-Mendoza et al. 2015). However, this did not provide an advantage for the symbiotic performance in general (Perez-Mendoza et al. 2015).

Concluding remarks

Since its discovery, the picture of bacterial c-di-GMP mediated regulation has become more and more fine-grained and widely conserved roles for this almost ubiquitous second messenger in controlling the switch between sessile and motile lifestyles have been recognized in diverse bacteria. In the recent years, several studies provided insights into the c-di-GMP regulatory network in alpha-rhizobial root nodule symbionts and pointed to a role in the free-living rather than in the symbiotic state. So far, mechanistic studies of c-di-GMP mediated regulation mostly focused on S. meliloti. Studies in S. meliloti contributed novel types of c-di-GMP binding proteins and thus broadened the repertoire of known receptors of this second messenger in bacteria. Yet, there is certainly much to discover in this and other alpha-rhizobia. The most important gap in our knowledge of c-di-GMP signaling in these bacteria comprise the environmental signals controlling c-di-GMP mediated regulation as well as perception and transduction of these signals.


Corresponding author: Anke Becker, Center for Synthetic Microbiology (SYNMIKRO), Philipps-Universität Marburg, D-35032Marburg, Germany; Department of Biology, Philipps-Universität Marburg, D-35032Marburg, Germany, E-mail:
Present address:Simon Schäper, Instituto de Tecnologia Química e Biológica António Xavier, Oeiras, Portugal.

Award Identifier / Grant number: SFB 987

Acknowledgment

Work by the authors was funded by the German Research Foundation in the framework of Collaborative Research Centre CRC 987.

  1. Author contribution: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This study was supported by German Research Foundation under grant CRC 987.

  3. Conflict of interest statement: The authors declare no conflicts of interest regarding this article.

References

Abel, S., Chien, P., Wassmann, P., Schirmer, T., Kaever, V., Laub, M. T., Baker, T.A., and Jenal, U. (2011). Regulatory cohesion of cell cycle and cell differentiation through interlinked phosphorylation and second messenger networks. Mol. Cell 43: 550–560, https://doi.org/10.1016/j.molcel.2011.07.018.Search in Google Scholar

Baena, I., Pérez-Mendoza, D., Sauviac, L., Francesch, K., Martín, M., Rivilla, R., Bonilla, I., Bruand, C., Sanjuán, J., and Lloret, J. (2019). A partner-switching system controls activation of mixed-linkage β-glucan synthesis by c-di-GMP in Sinorhizobium meliloti. Environ. Microbiol. 21: 3379–3391, https://doi.org/10.1111/1462-2920.14624.Search in Google Scholar

Bahlawane, C., Baumgarth, B., Serrania, J., Rüberg, S., and Becker, A. (2008a). Fine-tuning of galactoglucan biosynthesis in Sinorhizobium meliloti by differential WggR (ExpG)-, PhoB-, and MucR-dependent regulation of two promoters. J. Bacteriol. 190: 3456–3466, https://doi.org/10.1128/jb.00062-08.Search in Google Scholar

Bahlawane, C., McIntosh, M., Krol, E., and Becker, A. (2008b). Sinorhizobium meliloti regulator MucR couples exopolysaccharide synthesis and motility. Mol. Plant Microbe Interact. 21: 1498–1509, https://doi.org/10.1094/mpmi-21-11-1498.Search in Google Scholar

Barends, T.R., Hartmann, E., Griese, J.J., Beitlich, T., Kirienko, N.V., Ryjenkov, D.A., Reinstein, J., Shoeman, R. L., Gomelsky, M., and Schlichting, I. (2009). Structure and mechanism of a bacterial light-regulated cyclic nucleotide phosphodiesterase. Nature 459: 1015–1008, https://doi.org/10.1038/nature07966.Search in Google Scholar

Battisti, L., Lara, J.C., and Leigh, J.A. (1992). Specific oligosaccharide form of the Rhizobium meliloti exopolysaccharide promotes nodule invasion in alfalfa. Proc. Natl. Acad. Sci. U. S. A. 89: 5625–5629, https://doi.org/10.1073/pnas.89.12.5625.Search in Google Scholar

Bellini, D., Caly, D.L., McCarthy, Y., Bumann, M., An, S.Q., Dow, J. M., Ryan, R.P., and Walsh, M.A. (2014). Crystal structure of an HD-GYP domain cyclic-di-GMP phosphodiesterase reveals an enzyme with a novel trinuclear catalytic iron centre. Mol. Microbiol. 91: 26–38, https://doi.org/10.1111/mmi.12447.Search in Google Scholar

Bellini, D., Horrell, S., Hutchin, A., Phippen, C.W., Strange, R.W., Cai, Y., Wagner, A., Webb, J.S., Tews, I., and Walsh, M.A. (2017). Dimerisation induced formation of the active site and the identification of three metal sites in EAL-phosphodiesterases. Sci. Rep. 7: 42166, https://doi.org/10.1038/srep42166.Search in Google Scholar

Benach, J., Swaminathan, S.S., Tamayo, R., Handelman, S.K., Folta-Stogniew, E., Ramos, J.E., Forouhar, F., Neely, H., Seetharaman, J., Camilli, A., et al. (2007). The structural basis of cyclic diguanylate signal transduction by PilZ domains. EMBO J. 26: 5153–5166, https://doi.org/10.1038/sj.emboj.7601918.Search in Google Scholar

Bertram-Drogatz, P.A., Quester, I., Becker, A., and Pühler, A. (1998). The Sinorhizobium meliloti MucR protein, which is essential for the production of high-molecular-weight succinoglycan exopolysaccharide, binds to short DNA regions upstream of exoH and exoY. Mol. Gen. Genet. 257: 433–41, https://doi.org/10.1007/s004380050667.Search in Google Scholar

Bhasme, P., Wei, Q., Xu, A., Naqvi, S.T.A., Wang, D., and Ma, L.Z. (2020). Evaluation and characterization of the predicted diguanylate cyclase-encoding genes in Pseudomonas aeruginosa. Microbiologyopen 9: e975, https://doi.org/10.1002/mbo3.975.Search in Google Scholar

Bobrov, A.G., Kirillina, O., Ryjenkov, D.A., Waters, C.M., Price, P.A., Fetherston, J.D., Mack, D., Goldman, W.E., Gomelsky, M., and Perry, R.D. (2011). Systematic analysis of cyclic di-GMP signalling enzymes and their role in biofilm formation and virulence in Yersinia pestis. Mol. Microbiol. 79: 533–551, https://doi.org/10.1111/j.1365-2958.2010.07470.x.Search in Google Scholar

Boehm, A., Kaiser, M., Li, H., Spangler, C., Kasper, C.A., Ackermann, M., Kaever, V., Sourjik, V., Roth, V., and Jenal, U. (2010). Second messenger-mediated adjustment of bacterial swimming velocity. Cell 141: 107–116, https://doi.org/10.1016/j.cell.2010.01.018.Search in Google Scholar

Bourret, R.B. (2010). Receiver domain structure and function in response regulator proteins. Curr. Opin. Microbiol. 13: 142–149, https://doi.org/10.1016/j.mib.2010.01.015.Search in Google Scholar

Brown, P.J., de Pedro, M.A., Kysela, D. T., Van der Henst, C., Kim, J., De Bolle, X., Fuqua, C., and Brun, Y.V. (2012). Polar growth in the alphaproteobacterial order Rhizobiales. Proc. Natl. Acad. Sci. U. S. A. 109: 1697–1701, https://doi.org/10.1073/pnas.1114476109.Search in Google Scholar

Caetano-Anollés, G., Wall, L.G., De Micheli, A. T., Macchi, E.M., Bauer, W.D., and Favelukes, G. (1988). Role of motility and chemotaxis in efficiency of nodulation by Rhizobium meliloti. Plant Physiol. 86: 1228–1235, https://doi.org/10.1104/pp.86.4.1228.Search in Google Scholar

Calatrava-Morales, N., McIntosh, M., and Soto, M.J. (2018). Regulation mediated by N-acyl homoserine lactone quorum sensing signals in the rhizobium-legume symbiosis. Genes 9: E263, https://doi.org/10.3390/genes9050263.Search in Google Scholar

Chan, C., Paul, R., Samoray, D., Amiot, N. C., Giese, B., Jenal, U., and Schirmer, T. (2004). Structural basis of activity and allosteric control of diguanylate cyclase. Proc. Natl. Acad. Sci. U.S.A. 101: 17084–17089, https://doi.org/10.1073/pnas.0406134101.Search in Google Scholar

Chao, T.C., Buhrmester, J., Hansmeier, N., Pühler, A., and Weidner, S. (2005). Role of the regulatory gene rirA in the transcriptional response of Sinorhizobium meliloti to iron limitation. Appl. Environ. Microbiol. 71: 5969–5982, https://doi.org/10.1128/aem.71.10.5969-5982.2005.Search in Google Scholar

Charoenpanich, P., Meyer, S., Becker, A., and McIntosh, M. (2013). Temporal expression program of quorum sensing-based transcription regulation in Sinorhizobium meliloti. J. Bacteriol. 195: 3224–3236, https://doi.org/10.1128/jb.00234-13.Search in Google Scholar

Cheng, H.P. and Walker, G.C. (1998). Succinoglycan is required for initiation and elongation of infection threads during nodulation of alfalfa by Rhizobium meliloti. J. Bacteriol. 180: 5183–5191, https://doi.org/10.1128/jb.180.19.5183-5191.1998.Search in Google Scholar

Chin, K.H., Lee, Y.C., Tu, Z.L., Chen, C.H., Tseng, Y.H., Yang, J.M., Ryan, R.P., McCarthy, Y., Dow, J.M., Wang, A.H., et al. (2010). The cAMP receptor-like protein CLP is a novel c-di-GMP receptor linking cell-cell signaling to virulence gene expression in Xanthomonas campestris. J. Mol. Biol. 396: 646–662, https://doi.org/10.1016/j.jmb.2009.11.076.Search in Google Scholar

Christen, B., Christen, M., Paul, R., Schmid, F., Folcher, M., Jenoe, P., Meuwly, M., and Jenal, U. (2006). Allosteric control of cyclic di-GMP signaling. J. Biol. Chem. 281: 32015–32024, https://doi.org/10.1074/jbc.m603589200.Search in Google Scholar

Christen, M., Christen, B., Allan, M.G., Folcher, M., Jenö, P., Grzesiek, S., and Jenal, U. (2007). DgrA is a member of a new family of cyclic diguanosine monophosphate receptors and controls flagellar motor function in Caulobacter crescentus. Proc. Natl. Acad. Sci. U.S.A. 104: 4112–4117, https://doi.org/10.1073/pnas.0607738104.Search in Google Scholar

Christen, M., Kulasekara, H.D., Christen, B., Kulasekara, B.R., Hoffman, L.R., and Miller, S.I. (2010). Asymmetrical distribution of the second messenger c-di-GMP upon bacterial cell division. Science 328: 1295–1297, https://doi.org/10.1126/science.1188658.Search in Google Scholar

Domínguez-Ferreras, A., Pérez-Arnedo, R., Becker, A., Olivares, J., Soto, M.J., and Sanjuán, J. (2006). Transcriptome profiling reveals the importance of plasmid pSymB for osmoadaptation of Sinorhizobium meliloti. J. Bacteriol. 188: 7617–7625. https://doi.org/10.1128/JB.00719-06.Search in Google Scholar

Downie, J.A. (2010). The roles of extracellular proteins, polysaccharides and signals in the interactions of rhizobia with legume roots. FEMS Microbiol. Rev. 34: 150–170, https://doi.org/10.1111/j.1574-6976.2009.00205.x.Search in Google Scholar

Duerig, A., Abel, S., Folcher, M., Nicollier, M., Schwede, T., Amiot, N., Giese, B., and Jenal, U. (2009). Second messenger-mediated spatiotemporal control of protein degradation regulates bacterial cell cycle progression. Genes Dev. 23: 93–104, https://doi.org/10.1101/gad.502409.Search in Google Scholar

Edwards, A., Frederix, M., Wisniewski-Dyé, F., Jones, J., Zorreguieta, A., and Downie, J.A. (2009). The cin and rai quorum-sensing regulatory systems in Rhizobium leguminosarum are coordinated by ExpR and CinS, a small regulatory protein coexpressed with CinI. J. Bacteriol. 191: 3059–3067, https://doi.org/10.1128/jb.01650-08.Search in Google Scholar

Fang, X., Ahmad, I., Blanka, A., Schottkowski, M., Cimdins, A., Galperin, M.Y., Römling, U., and Gomelsky, M. (2014). GIL, a new c-di-GMP-binding protein domain involved in regulation of cellulose synthesis in enterobacteria. Mol. Microbiol. 93: 439–452, https://doi.org/10.1111/mmi.12672.Search in Google Scholar

Fang, X. and Gomelsky, M. (2010). A post-translational., c-di-GMP-dependent mechanism regulating flagellar motility. Mol. Microbiol. 76: 1295–1305, https://doi.org/10.1111/j.1365-2958.2010.07179.x.Search in Google Scholar

Feirer, N., Xu, J., Allen, K.D., Koestler, B.J., Bruger, E.L., Waters, C.M., White, R.H., and Fuqua, C. (2015). A pterin-dependent signaling pathway regulates a dual-function diguanylate cyclase-phosphodiesterase controlling surface attachment in Agrobacterium tumefaciens. MBio 6: e00156, https://doi.org/10.1128/mbio.00156-15.Search in Google Scholar

Frage, B., Döhlemann, J., Robledo, M., Lucena, D., Sobetzko, P., Graumann, P.L., and Becker, A. (2016). Spatiotemporal choreography of chromosome and megaplasmids in the Sinorhizobium meliloti cell cycle. Mol. Microbiol. 100: 808–823, https://doi.org/10.1111/mmi.13351.Search in Google Scholar

Fraysse, N., Couderc, F., and Poinsot, V. (2003). Surface polysaccharide involvement in establishing the rhizobium-legume symbiosis. Eur. J. Biochem. 270: 1365–1380, https://doi.org/10.1046/j.1432-1033.2003.03492.x.Search in Google Scholar

Fujiwara, T., Komoda, K., Sakurai, N., Tajima, K., Tanaka, I., and Yao, M. (2013). The c-di-GMP recognition mechanism of the PilZ domain of bacterial cellulose synthase subunit A. Biochem. Biophys. Res. Commun. 431: 802–807, https://doi.org/10.1016/j.bbrc.2012.12.103.Search in Google Scholar

Gallagher, K.A., Schumacher, M.A., Bush, M.J., Bibb, M.J., Chandra, G., Holmes, N.A., Zeng, W., Henderson, M., Zhang, H., Findlay, K. C., et al. (2020). c-di-GMP arms an anti-σ to control progression of multicellular differentiation in Streptomyces. Mol. Cell. 77: 586–599, https://doi.org/10.1016/j.molcel.2019.11.006.Search in Google Scholar

Galperin, M.Y. and Chou, S.H. (2020). Structural conservation and diversity of PilZ-related domains. J. Bacteriol. 202: e00664–19, https://doi.org/10.1128/jb.00664-19.Search in Google Scholar

Galperin, M.Y., Nikolskaya, A.N., and Koonin, E.V. (2001). Novel domains of the prokaryotic two-component signal transduction systems. FEMS Microbiol. Lett. 203: 11–21, https://doi.org/10.1111/j.1574-6968.2001.tb10814.x.Search in Google Scholar

Galperin, M.Y. (2005). A census of membrane-bound and intracellular signal transduction proteins in bacteria: bacterial IQ, extroverts and introverts. BMC Microbiol. 5: 35, https://doi.org/10.1186/1471-2180-5-35.Search in Google Scholar

Gao, M., Chen, H., Eberhard, A., Gronquist, M.R., Robinson, J.B., Rolfe, B.G., and Bauer, W.D. (2005). sinI- and expR-dependent quorum sensing in Sinorhizobium meliloti. J. Bacteriol. 187: 7931–7944, https://doi.org/10.1128/jb.187.23.7931-7944.2005.Search in Google Scholar

Gao, M., Coggin, A., Yagnik, K., and Teplitski, M. (2012). Role of specific quorum-sensing signals in the regulation of exopolysaccharide II production within Sinorhizobium meliloti spreading colonies. PLoS One 7: e42611, https://doi.org/10.1371/journal.pone.0042611.Search in Google Scholar

Gao, S., Romdhane, S.B., Beullens, S., Kaever, V., Lambrichts, I., Fauvart, M., and Michiels, J. (2014). Genomic analysis of cyclic-di-GMP-related genes in rhizobial type strains and functional analysis in Rhizobium etli. Appl. Microbiol. Biotechnol. 98: 4589–4602, https://doi.org/10.1007/s00253-014-5722-7.Search in Google Scholar

Gomelsky, M. (2011). cAMP, c-di-GMP, c-di-AMP and now cGMP: bacteria use them all!. Mol. Microbiol. 79: 562–565, https://doi.org/10.1111/j.1365-2958.2010.07514.x.Search in Google Scholar

González, J.E., Reuhs, B.L., and Walker, G.C. (1996). Low molecular weight EPS II of Rhizobium meliloti allows nodule invasion in Medicago sativa. Proc. Natl. Acad. Sci. U.S.A. 93: 8636–8641, https://doi.org/10.1073/pnas.93.16.8636.Search in Google Scholar

Gründling, A. and Lee, V.T. (2016). Old concepts, new molecules and current approaches applied to the bacterial nucleotide signalling field. Philos. Trans. R. Soc. Lond. B Biol. Sci. 371: 20150503, https://doi.org/10.1098/rstb.2015.0503.Search in Google Scholar

Gu, X., Lee, S.G., and Bar-Peled, M. (2011). Biosynthesis of UDP-xylose and UDP-arabinose in Sinorhizobium meliloti 1021: first characterization of a bacterial UDP-xylose synthase, and UDP-xylose 4-epimerase. Microbiology 157: 260–269, https://doi.org/10.1099/mic.0.040758-0.Search in Google Scholar

Gurich, N. and González, J.E. (2009). Role of quorum sensing in Sinorhizobium meliloti-Alfalfa symbiosis. J. Bacteriol. 191: 4372–4382, https://doi.org/10.1128/jb.00376-09.Search in Google Scholar

Hengge, R. (2020). Linking bacterial growth, survival, and multicellularity – small signaling molecules as triggers and drivers. Curr. Opin. Microbiol. 55: 57–66, https://doi.org/10.1016/j.mib.2020.02.007.Search in Google Scholar

Hengge, R., Galperin, M.Y., Ghigo, J. M., Gomelsky, M., Green, J., Hughes, K. T., Jenal, U., and Landini, P. (2015). Systematic nomenclature for GGDEF and EAL domain-containing cyclic di-GMP turnover proteins of Escherichia coli. J. Bacteriol. 198: 7–11, https://doi.org/10.1128/jb.00424-15.Search in Google Scholar

Her, G.-R., Glazebrook, J., Walker, G.C., and Reinhold, V.N. (1990). Structural studies of a novel exopolysaccharide produced by a mutant of Rhizobium meliloti Rm1021. Carbohydr. Res. 198: 305–312, https://doi.org/10.1016/0008-6215(90)84300-j.Search in Google Scholar

Hickman, J.W. and Harwood, C.S. (2008). Identification of FleQ from Pseudomonas aeruginosa as a c-di-GMP-responsive transcription factor. Mol. Microbiol. 69: 376–389, https://doi.org/10.1111/j.1365-2958.2008.06281.x.Search in Google Scholar

Hickman, J.W., Tifrea, D.F., and Harwood, C.S. (2005). A chemosensory system that regulates biofilm formation through modulation of cyclic diguanylate levels. Proc. Natl. Acad. Sci. U. S. A. 102: 14422–14427, https://doi.org/10.1073/pnas.0507170102.Search in Google Scholar

Hoang, H.H., Becker, A., and González, J.E. (2004). The LuxR homolog ExpR., in combination with the Sin quorum sensing system., plays a central role in Sinorhizobium meliloti gene expression. J. Bacteriol. 186: 5460–5472, https://doi.org/10.1128/jb.186.16.5460-5472.2004.Search in Google Scholar

Hoang, H.H., Gurich, N., and González, J.E. (2008). Regulation of motility by the ExpR/Sin quorum-sensing system in Sinorhizobium meliloti. J. Bacteriol. 190: 861–871, https://doi.org/10.1128/jb.01310-07.Search in Google Scholar

Hochstrasser, R., Kessler, A., Sahr, T., Simon, S., Schell, U., Gomez-Valero, L., Buchrieser, C., and Hilbi, H. (2019). The pleiotropic Legionella transcription factor LvbR links the Lqs and c-di-GMP regulatory networks to control biofilm architecture and virulence. Environ. Microbiol. 21: 1035–1053, https://doi.org/10.1111/1462-2920.14523.Search in Google Scholar

Hou, Y.J., Yang, W.S., Hong, Y., Zhang, Y., Wang, D.C., and Li, D.F. (2020). Structural insights into the mechanism of c-di-GMP-bound YcgR regulating flagellar motility in Escherichia coli. J. Biol. Chem. 295: 808–821, https://doi.org/10.1074/jbc.ra119.009739.Search in Google Scholar

Hsieh, M.L., Hinton, D.M., and Waters, C.M. (2018). VpsR and cyclic di-GMP together drive transcription initiation to activate biofilm formation in Vibrio cholerae. Nucleic Acids Res. 46: 8876–8887, https://doi.org/10.1093/nar/gky606.Search in Google Scholar

Hunter, J.L., Severin, G.B., Koestler, B.J., and Waters, C.M. (2014). The Vibrio cholerae diguanylate cyclase VCA0965 has an AGDEF active site and synthesizes cyclic di-GMP. BMC Microbiol. 14: 22, https://doi.org/10.1186/1471-2180-14-22.Search in Google Scholar

Janczarek, M. (2011). Environmental signals and regulatory pathways that influence exopolysaccharide production in rhizobia. Int. J. Mol. Sci. 12: 7898–7933, https://doi.org/10.3390/ijms12117898.Search in Google Scholar

Jenal, U., Reinders, A., and Lori, C. (2017). Cyclic di-GMP: second messenger extraordinaire. Nat. Rev. Microbiol. 15: 271–284, https://doi.org/10.1038/nrmicro.2016.190.Search in Google Scholar

Jones, K.M., Kobayashi, H., Davies, B.W., Taga, M.E., and Walker, G.C. (2007). How rhizobial symbionts invade plants: the Sinorhizobium-Medicago model. Nat. Rev. Microbiol. 5: 619–633, https://doi.org/10.1038/nrmicro1705.Search in Google Scholar

Kharadi, R.R., Castiblanco, L.F., Waters, C.M., and Sundin, G.W. (2018). Phosphodiesterase genes regulate amylovoran production, biofilm formation, and virulence in Erwinia amylovora. Appl. Environ. Microbiol. 85: e02233–18, https://doi.org/10.1128/aem.02233-18.Search in Google Scholar

Krasteva, P.V. and Sondermann, H. (2017). Versatile modes of cellular regulation via cyclic dinucleotides. Nat. Chem. Biol. 13: 350–359, https://doi.org/10.1038/nchembio.2337.Search in Google Scholar

Krol, E. and Becker, A. (2004). Global transcriptional analysis of the phosphate starvation response in Sinorhizobium meliloti strains 1021 and 2011. Mol. Genet. Genom. 272: 1–17, https://doi.org/10.1007/s00438-004-1030-8.Search in Google Scholar

Krol, E. and Becker, A. (2011). ppGpp in Sinorhizobium meliloti: biosynthesis in response to sudden nutritional downshifts and modulation of the transcriptome. Mol. Microbiol. 81: 1233–1254, https://doi.org/10.1111/j.1365-2958.2011.07752.x.Search in Google Scholar

Krol, E., Yau, H.C.L., Lechner, M., Schäper, S., Bange, G., Vollmer, W., and Becker, A. (2020). Tol-Pal system and Rgs proteins interact to promote unipolar growth and cell division in Sinorhizobium meliloti. MBio. 11: e00306–e00320. https://doi.org/10.1128/mBio.00306-20.Search in Google Scholar

Kulasekara, B.R., Kamischke, C., Kulasekara, H.D., Christen, M., Wiggins, P.A., and Miller, S.I. (2013). c-di-GMP heterogeneity is generated by the chemotaxis machinery to regulate flagellar motility. eLife 2: e01402, https://doi.org/10.7554/elife.01402.Search in Google Scholar

Kumar, A., Alam, A., Rani, M., Ehtesham, N.Z., and Hasnain, S.E. (2017). Biofilms: survival and defense strategy for pathogens. Int J. Med. Microbiol. 307: 481–489, https://doi.org/10.1016/j.ijmm.2017.09.016.Search in Google Scholar

Laus, M.C., Logman, T.J., Lamers, G.E., Van Brussel, A.A., Carlson, R.W., and Kijne, J.W. (2006). A novel polar surface polysaccharide from Rhizobium leguminosarum binds host plant lectin. Mol. Microbiol. 59: 1704–1713, https://doi.org/10.1111/j.1365-2958.2006.05057.x.Search in Google Scholar

Laus, M.C., van Brussel, A.A., and Kijne, J.W. (2005). Role of cellulose fibrils and exopolysaccharides of Rhizobium leguminosarum in attachment to and infection of Vicia sativa root hairs. Mol. Plant Microbe Interact. 18: 533–538, https://doi.org/10.1094/mpmi-18-0533.Search in Google Scholar

Li, W. and He, Z.G. (2012). LtmA., a novel cyclic di-GMP-responsive activator, broadly regulates the expression of lipid transport and metabolism genes in Mycobacterium smegmatis. Nucleic Acids Res. 40: 11292–11307, https://doi.org/10.1093/nar/gks923.Search in Google Scholar

Liang, Z.X. (2015). The expanding roles of c-di-GMP in the biosynthesis of exopolysaccharides and secondary metabolites. Nat. Prod. Rep. 32: 663–683, https://doi.org/10.1039/c4np00086b.Search in Google Scholar

Lin Chua, S., Liu, Y., Li, Y., Jun Ting, H., Kohli, G.S., Cai, Z., Suwanchaikasem, P., Kau Kit Goh, K., Pin Ng, S., Tolker-Nielsen, T., et al. (2017). Reduced intracellular c-di-GMP content increases expression of QS-regulated genes in Pseudomonas aeruginosa. Front. Cell. Infect. Microbiol. 7: 451, https://doi.org/10.3389/fcimb.2017.00451.Search in Google Scholar

López-Baena, F.J., Vinardell, J.M., and Medina, C. (2019). Regulation of protein secretion systems mediated by cyclic diguanylate in plant-interacting bacteria. Front. Microbiol. 10: 1289, https://doi.org/10.3389/fmicb.2019.01289.Search in Google Scholar

Lowe-Power, T.M., Khokhani, D., and Allen, C. (2018). How Ralstonia solanacearum exploits and thrives in the flowing plant xylem environment. Trends Microbiol. 26: 929–942, https://doi.org/10.1016/j.tim.2018.06.002.Search in Google Scholar

Marketon, M.M., Gronquist, M.R., Eberhard, A., and González, J.E. (2002). Characterization of the Sinorhizobium meliloti sinR/sinI locus and the production of novel N-acyl homoserine lactones. J. Bacteriol. 184: 5686–5695, https://doi.org/10.1128/jb.184.20.5686-5695.2002.Search in Google Scholar

Maunders, E. and Welch, M. (2017). Matrix exopolysaccharides; the sticky side of biofilm formation. FEMS Microbiol. Lett. 364: fnx120, https://doi.org/10.1093/femsle/fnx120.Search in Google Scholar

McDougald, D., Rice, S.A., Barraud, N., Steinberg, P.D., and Kjelleberg, S. (2012). Should we stay or should we go: mechanisms and ecological consequences for biofilm dispersal. Nat. Rev. Microbiol. 10: 39–50, https://doi.org/10.1038/nrmicro2695.Search in Google Scholar

McIntosh, M., Meyer, S., and Becker, A. (2009). Novel Sinorhizobium meliloti quorum sensing positive and negative regulatory feedback mechanisms respond to phosphate availability. Mol. Microbiol. 74: 1238–1256, https://doi.org/10.1111/j.1365-2958.2009.06930.x.Search in Google Scholar

Morgan, J.L.W., McNamara, J.T., and Zimmer, J. (2014). Mechanism of activation of bacterial cellulose synthase by cyclic-di-GMP. Nat. Struct. Mol. Biol. 21: 489–496, https://doi.org/10.1038/nsmb.2803.Search in Google Scholar

Mueller, K. and González, J.E. (2011). Complex regulation of symbiotic functions is coordinated by MucR and quorum sensing in Sinorhizobium meliloti. J. Bacteriol. 193: 485–496, https://doi.org/10.1128/jb.01129-10.Search in Google Scholar

Muszyński, A., Heiss, C., Hjuler, C.T., Sullivan, J.T., Kelly, S.J., Thygesen, M.B., Stougaard, J., Azadi, P., Carlson, R.W., and Ronson, C.W. (2016). Structures of exopolysaccharides involved in receptor-mediated perception of Mesorhizobium loti by Lotus japonicus. J. Biol. Chem. 291: 20946–20961, https://doi.org/10.1074/jbc.M116.743856.Search in Google Scholar

Nair, H.A., Periasamy, S., Yang, L., Kjelleberg, S., and Rice, S.A. (2017). Real time, spatial, and temporal mapping of the distribution of c-di-GMP during biofilm development. J. Biol. Chem. 292: 477–487, https://doi.org/10.1074/jbc.m116.746743.Search in Google Scholar

Niehaus, K. and Becker, A. (1998). The role of microbial surface polysaccharides in the Rhizobium-legume interaction. Subcell. Biochem. 29: 73–116, https://doi.org/10.1007/978-1-4899-1707-2_3.Search in Google Scholar

Nieto, V., Partridge, J.D., Severin, G.B., Lai, R.Z., Waters, C.M., Parkinson, J.S., and Harshey, R.M. (2019). Under elevated c-di-GMP in Escherichia coli, YcgR alters flagellar motor bias and speed sequentially, with additional negative control of the flagellar regulon via the adaptor protein RssB. J. Bacteriol. 202: e00578–19, https://doi.org/10.1128/jb.00578-19.Search in Google Scholar

Pallegar, P., Peña-Castillo, L., Langille, E., Gomelsky, M., and Lang, AS (2020). Cyclic di-GMP-mediated regulation of gene transfer and motility in Rhodobacter capsulatus. J. Bacteriol. 202: e00554–19, https://doi.org/10.1128/jb.00554-19.Search in Google Scholar

Paul, K., Nieto, V., Carlquist, W.C., Blair, D.F., and Harshey, R.M. (2010). The c-di-GMP binding protein YcgR controls flagellar motor direction and speed to affect chemotaxis by a “backstop brake” mechanism. Mol. Cell 38: 128–139, https://doi.org/10.1016/j.molcel.2010.03.001.Search in Google Scholar

Paul, R., Abel, S., Wassmann, P., Beck, A., Heerklotz, H., and Jenal, U. (2007). Activation of the diguanylate cyclase PleD by phosphorylation-mediated dimerization. J. Biol. Chem. 282: 29170–20177, https://doi.org/10.1074/jbc.m704702200.Search in Google Scholar

Paul, R., Weiser, S., Amiot, N.C., Chan, C., Schirmer, T., Giese, B., and Jenal, U. (2004). Cell cycle-dependent dynamic localization of a bacterial response regulator with a novel di-guanylate cyclase output domain. Genes Dev. 18: 715–27, https://doi.org/10.1101/gad.289504.Search in Google Scholar

Pérez-Mendoza, D., Aragón, IM., Prada-Ramírez, HA., Romero-Jiménez, L., Ramos, C., Gallegos, M.T., and Sanjuán, J. (2014). Responses to elevated c-di-GMP levels in mutualistic and pathogenic plant-interacting bacteria. PLoS One 9: e91645, https://doi.org/10.1371/journal.pone.0091645.Search in Google Scholar

Pérez-Mendoza, D., Bertinetti, D., Lorenz, R., Gallegos, M.T., Herberg, F.W., and Sanjuán, J. (2017). A novel c-di-GMP binding domain in glycosyltransferase BgsA is responsible for the synthesis of a mixed-linkage β-glucan. Sci. Rep. 7: 8997, https://doi.org/10.1038/s41598-017-09290-2.Search in Google Scholar

Pérez-Mendoza, D., Rodríguez-Carvajal, M., Romero-Jiménez, L., de Arajuo Farias, G., Lloret, J., Gallegos, M.T., and Sanjuán, J. (2015). Novel mixed-linkage β-glucan activated by c-di-GMP in Sinorhizobium meliloti. Proc. Natl. Acad. Sci. U.S.A. 112: E757–765, https://doi.org/10.1073/pnas.1421748112.Search in Google Scholar

Petters, T., Zhang, X., Nesper, J., Treuner-Lange, A., Gomez-Santos, N., Hoppert, M., Jenal, U., and Søgaard-Andersen, L. (2012). The orphan histidine protein kinase SgmT is a c-di-GMP receptor and regulates composition of the extracellular matrix together with the orphan DNA binding response regulator DigR in Myxococcus xanthus. Mol. Microbiol. 84: 147–165, https://doi.org/10.1111/j.1365-2958.2012.08015.x.Search in Google Scholar

Purcell, E.B. and Tamayo, R. (2016). Cyclic diguanylate signaling in gram-positive bacteria. FEMS Microbiol. Rev. 40: 753–773, https://doi.org/10.1093/femsre/fuw013.Search in Google Scholar

Ramelot, T.A., Yee, A., Cort, J.R., Semesi, A., Arrowsmith, C.H., and Kennedy, M.A. (2007). NMR structure and binding studies confirm that PA4608 from Pseudomonas aeruginosa is a PilZ domain and a c-di-GMP binding protein. Proteins 66: 266–271, https://doi.org/10.1002/prot.21199.Search in Google Scholar

Reinhold, B.B., Chan, S.Y., Reuber, T.L., Marra, A., Walker, G.C., and Reinhold, V.N. (1994). Detailed structural characterization of succinoglycan, the major exopolysaccharide of Rhizobium meliloti Rm1021. J. Bacteriol. 176: 1997–2002, https://doi.org/10.1128/jb.176.7.1997-2002.1994.Search in Google Scholar

Rinaudi, L.V. and Giordano, W. (2010). An integrated view of biofilm formation in rhizobia. FEMS Microbiol. Lett. 304: 1–11, https://doi.org/10.1111/j.1574-6968.2009.01840.x.Search in Google Scholar

Robledo, M., Rivera, L., Jiménez-Zurdo, J.I., Rivas, R., Dazzo, F., Velázquez, E., Martínez-Molina, E., Hirsch, A.M., and Mateos, P.F. (2012). Role of Rhizobium endoglucanase CelC2 in cellulose biosynthesis and biofilm formation on plant roots and abiotic surfaces. Microb. Cell Fact. 12: 125, https://doi.org/10.1186/1475-2859-11-125.Search in Google Scholar

Römling, U., Galperin, M.Y., and Gomelsky, M. (2013). Cyclic di-GMP: the first 25 years of a universal bacterial second messenger. Microbiol. Mol. Biol. Rev. 77: 1–52, https://doi.org/10.1128/mmbr.00043-12.Search in Google Scholar

Rotter, C., Mühlbacher, S., Salamon, D., Schmitt, R., and Scharf, B. (2006). Rem., a new transcriptional activator of motility and chemotaxis in Sinorhizobium meliloti. J. Bacteriol. 188: 6932–6942, https://doi.org/10.1128/jb.01902-05.Search in Google Scholar

Roy, S., Liu, W., Nandety, R.S., Crook, A., Mysore, K.S., Pislariu, C.I., Frugoli, J., Dickstein, R., and Udvardi, M.K. (2020). Celebrating 20 years of genetic discoveries in legume nodulation and symbiotic nitrogen fixation. Plant Cell 32: 15–41, https://doi.org/10.1105/tpc.19.00279.Search in Google Scholar

Rüberg, S., Pühler, A., and Becker, A. (1999). Biosynthesis of the exopolysaccharide galactoglucan in Sinorhizobium meliloti is subject to a complex control by the phosphate-dependent regulator PhoB and the proteins ExpG and MucR. Microbiology 145: 603–611, https://doi.org/10.1099/13500872-145-3-603.Search in Google Scholar

Sanchez-Contreras, M., Bauer, W.D., Gao, M., Robinson, J.B., and Allan Downie, J. (2007). Quorum-sensing regulation in rhizobia and its role in symbiotic interactions with legumes. Philos. Trans. R. Soc. Lond. B Biol. Sci. 362: 1149–1163, https://doi.org/10.1098/rstb.2007.2041.Search in Google Scholar

Schäper, S., Krol, E., Skotnicka, D., Kaever, V., Hilker, R., Søgaard-Andersen, L., and Becker, A. (2016). Cyclic di-GMP regulates multiple cellular functions in the symbiotic alphaproteobacterium Sinorhizobium meliloti. J. Bacteriol. 198: 521–535, https://doi.org/10.1128/jb.00795-15.Search in Google Scholar

Schäper, S., Steinchen, W., Krol, E., Altegoer, F., Skotnicka, D., Søgaard-Andersen, L., Bange, G., and Becker, A. (2017). AraC-like transcriptional regulator CuxR binds c-di-GMP by a PilZ-like mechanism to regulate extracellular polysaccharide production. Proc. Natl. Acad. Sci. U.S.A. 114: E4822–4831, https://doi.org/10.1073/pnas.1702435114.Search in Google Scholar

Schäper, S., Wendt, H., Bamberger, J., Sieber, V., Schmid, J., and Becker, A. (2019). A bifunctional UDP-sugar 4-epimerase supports biosynthesis of multiple cell surface polysaccharides in Sinorhizobium meliloti. J. Bacteriol. 201: e00801–18, https://doi.org/10.1128/jb.00801-18.Search in Google Scholar

Schäper, S., Yau, H.C.L., Krol, E., Skotnicka, D., Heimerl, T., Gray, J., Kaever, V., Søgaard-Andersen, L., Vollmer, W., and Becker, A. (2018). Seven-transmembrane receptor protein RgsP and cell wall-binding protein RgsM promote unipolar growth in Rhizobiales. PLoS Genet. 14: e1007594, https://doi.org/10.1371/journal.pgen.1007594.Search in Google Scholar

Scharf, B. and Schmitt, R. (2002). Sensory transduction to the flagellar motor of Sinorhizobium meliloti. J. Mol. Microbiol. Biotechnol. 4: 183–186.Search in Google Scholar

Schmidt, A.J., Ryjenkov, D.A., and Gomelsky, M. (2005). The ubiquitous protein domain EAL is a cyclic diguanylate-specific phosphodiesterase: enzymatically active and inactive EAL domains. J. Bacteriol. 187: 4774–4781, https://doi.org/10.1128/jb.187.14.4774-4781.2005.Search in Google Scholar

Schmitt, R. (2002). Sinorhizobial chemotaxis: a departure from the enterobacterial paradigm. Microbiology 148: 627–631, https://doi.org/10.1099/00221287-148-3-627.Search in Google Scholar

Serra, D.O., Richter, A.M., and Hengge, R. (2013). Cellulose as an architectural element in spatially structured Escherichia coli biofilms. J. Bacteriol. 195: 5540–5554, https://doi.org/10.1128/jb.00946-13.Search in Google Scholar

Seshasayee, A.S., Fraser, G.M., and Luscombe, N.M. (2010). Comparative genomics of cyclic-di-GMP signalling in bacteria: post-translational regulation and catalytic activity. Nucleic Acids Res. 38: 5970–5981, https://doi.org/10.1093/nar/gkq382.Search in Google Scholar

Severin, G.B. and Waters, C.M. (2019). Pyrimidines and cyclic trinucleotides join the second messenger symphony. Cell Host Microbe 25: 471–473, https://doi.org/10.1016/j.chom.2019.03.016.Search in Google Scholar

Skorupska, A., Janczarek, M., Marczak, M., Mazur, A., and Król, J. (2006). Rhizobial exopolysaccharides: genetic control and symbiotic functions. Microb. Cell Fact. 5: 7, https://doi.org/10.1186/1475-2859-5-7.Search in Google Scholar

Solano, C., García, B., Latasa, C., Toledo-Arana, A., Zorraquino, V., Valle, J., Casals, J., Pedroso, E., and Lasa, I. (2009). Genetic reductionist approach for dissecting individual roles of GGDEF proteins within the c-di-GMP signaling network in Salmonella. Proc. Natl. Acad. Sci. U.S.A. 106: 7997–8002, https://doi.org/10.1073/pnas.0812573106.Search in Google Scholar

Sourjik, V., Muschler, P., Scharf, B., and Schmitt, R. (2000). VisN and VisR are global regulators of chemotaxis, flagellar, and motility genes in Sinorhizobium (Rhizobium) meliloti. J. Bacteriol. 182: 782–788, https://doi.org/10.1128/jb.182.3.782-788.2000.Search in Google Scholar

Sprecher, K.S., Hug, I., Nesper, J., Potthoff, E., Mahi, M.A., Sangermani, M., Kaever, V., Schwede, T., Vorholt, J., and Jenal, U. (2017). Cohesive properties of the caulobacter crescentus holdfast adhesin are regulated by a novel c-di-GMP effector protein. MBio 8: e00294–17, https://doi.org/10.1128/mbio.00294-17.Search in Google Scholar

Staehelin, C., Forsberg, L.S., D’Haeze, W., Gao, M.Y., Carlson, R.W., Xie, Z.P., Pellock, B.J., Jones, K.M., Walker, G.C., Streit, W.R., et al. (2006). Exo-oligosaccharides of Rhizobium sp. strain NGR234 are required for symbiosis with various legumes. J. Bacteriol. 188: 6168–6178, https://doi.org/10.1128/jb.00365-06.Search in Google Scholar

Steiner, S., Lori, C., Boehm, A., and Jenal, U. (2013). Allosteric activation of exopolysaccharide synthesis through cyclic di-GMP-stimulated protein-protein interaction. EMBO J. 32: 354–368, https://doi.org/10.1038/emboj.2012.315.Search in Google Scholar

Stock, A.M., Robinson, V.L., and Goudreau, P.N. (2000). Two-component signal transduction. Annu. Rev. Biochem. 69: 183–215, https://doi.org/10.1146/annurev.biochem.69.1.183.Search in Google Scholar

Su, J., Zou, X., Huang, L., Bai, T., Liu, S., Yuan, M., Chou, S.H., He, Y.W., Wang, H., and He, J. (2016). DgcA, a diguanylate cyclase from Xanthomonas oryzae pv. oryzae regulates bacterial pathogenicity on rice. Sci. Rep. 6: 25978, https://doi.org/10.1038/srep25978.Search in Google Scholar

Sun, Y., Xie, Z., Sui, F., Liu, X., and Cheng, W. (2019). Identification of Cbp1, a c-di-GMP binding chemoreceptor in Azorhizobium caulinodans ORS571 involved in chemotaxis and nodulation of the host plant. Front. Microbiol. 10: 638, https://doi.org/10.3389/fmicb.2019.00638.Search in Google Scholar

Sundriyal, A., Massa, C., Samoray, D., Zehender, F., Sharpe, T., Jenal, U., and Schirmer, T. (2014). Inherent regulation of EAL domain-catalyzed hydrolysis of second messenger cyclic di-GMP. J. Biol. Chem. 289: 6978–6990, https://doi.org/10.1074/jbc.m113.516195.Search in Google Scholar

Tchigvintsev, A., Xu, X., Singer, A., Chang, C., Brown, G., Proudfoot, M., Cui, H., Flick, R., Anderson, W.F., Joachimiak, A., et al. (2010). Structural insight into the mechanism of c-di-GMP hydrolysis by EAL domain phosphodiesterases. J. Mol. Biol. 402: 524–538, https://doi.org/10.1016/j.jmb.2010.07.050.Search in Google Scholar

Trampari, E., Stevenson, C.E., Little, R.H., Wilhelm, T., Lawson, D.M., and Malone, J.G. (2015). Bacterial rotary export ATPases are allosterically regulated by the nucleotidesecond messenger cyclic-di-GMP. J. Biol. Chem. 290: 24470–24483, https://doi.org/10.1074/jbc.M115.661439.Search in Google Scholar

Trimble, M.J. and McCarter, L.L. (2011). Bis-(3’-5’)-cyclic dimeric GMP-linked quorum sensing controls swarming in Vibrio parahaemolyticus. Proc. Natl. Acad. Sci. U.S.A. 108: 18079–18084, https://doi.org/10.1073/pnas.1113790108.Search in Google Scholar

Tschowri, N., Schumacher, M.A., Schlimpert, S., Chinnam, N.B., Findlay, K.C., Brennan, R.G., and Buttner, M.J. (2014). Tetrameric c-di-GMP mediates effective transcription factor dimerization to control Streptomyces development. Cell 158: 1136–1147, https://doi.org/10.1016/j.cell.2014.07.022.Search in Google Scholar

Urzainqui, A. and Walker, G.C. (1992). Exogenous suppression of the symbiotic deficiencies of Rhizobium meliloti exo mutants. J. Bacteriol. 174: 3403–3406, https://doi.org/10.1128/jb.174.10.3403-3406.1992.Search in Google Scholar

Waldron, E.J., Snyder, D., Fernandez, N.L., Sileo, E., Inoyama, D., Freundlich, J.S., Waters, C.M., Cooper, V.S., and Neiditch, M.B. (2019). Structural basis of DSF recognition by its receptor RpfR and its regulatory interaction with the DSF synthase RpfF. PLoS Biol. 17: e3000123, https://doi.org/10.1371/journal.pbio.3000123.Search in Google Scholar

Wang, Y.C., Chin, K.H., Tu, Z.L., He, J., Jones, C.J., Sanchez, D.Z., Yildiz, F.H., Galperin, M.Y., and Chou, S.H. (2016). Nucleotide binding by the widespread high-affinity cyclic di-GMP receptor MshEN domain. Nat. Commun. 7: 12481, https://doi.org/10.1038/ncomms12481.Search in Google Scholar

Wang, C., Kemp, J., Da Fonseca, I.O., Equi, R.C., Sheng, X., Charles, T.C., and Sobral, B.W.S. (2010). Sinorhizobium meliloti 1021 loss‐of‐function deletion mutation in chvI and its phenotypic characteristics. Mol. Plant‐Microbe Interact. 23: 153–160, https://doi.org/10.1094/mpmi-23-2-0153.Search in Google Scholar

Wassmann, P., Chan, C., Paul, R., Beck, A., Heerklotz, H., Jenal, U., and Schirmer, T. (2007). Structure of BeF3-modified response regulator PleD: implications for diguanylate cyclase activation, catalysis, and feedback inhibition. Structure 15: 915–927, https://doi.org/10.1016/j.str.2007.06.016.Search in Google Scholar

Webb, B.A., Hildreth, S., Helm, R.F., and Scharf, B.E. (2014). Sinorhizobium meliloti chemoreceptor McpU mediates chemotaxis toward host plant exudates through direct proline sensing. Appl. Environ. Microbiol. 80: 3404–3415, https://doi.org/10.1128/aem.00115-14.Search in Google Scholar

Weinhouse, H., Sapir, S., Amikam, D., Shilo, Y., Volman, G., Ohana, P., and Benziman, M. (1997). c-di-GMP-binding protein., a new factor regulating cellulose synthesis in Acetobacter xylinum. FEBS Lett. 416: 207–211, https://doi.org/10.1016/s0014-5793(97)01202-7.Search in Google Scholar

Whitchurch, C.B., Tolker-Nielsen, T., Ragas, P.C., and Mattick, J.S. (2002). Extracellular DNA required for bacterial biofilm formation. Science 295: 1487, https://doi.org/10.1126/science.295.5559.1487.Search in Google Scholar

Xu, J., Kim, J., Danhorn, T., Merritt, P.M., and Fuqua, C. (2012). Phosphorus limitation increases attachment in Agrobacterium tumefaciens and reveals a conditional functional redundancy in adhesin biosynthesis. Res. Microbiol. 163: 674–684, https://doi.org/10.1016/j.resmic.2012.10.013.Search in Google Scholar

Xu, J., Kim, J., Koestler, B.J., Choi, J.H., Waters, C.M., and Fuqua, C. (2013). Genetic analysis of Agrobacterium tumefaciens unipolar polysaccharide production reveals complex integrated control of the motile-to-sessile switch. Mol. Microbiol. 89: 929–948, https://doi.org/10.1111/mmi.12321.Search in Google Scholar

Xu, L., Xin, L., Zeng, Y., Yam, JK., Ding, Y., Venkataramani, P., Cheang, Q.W., Yang, X., Tang, X., Zhang, L.H., et al. (2016). A cyclic di-GMP-binding adaptor protein interacts with a chemotaxis methyltransferase to control flagellar motor switching. Sci. Signal. 9: ra102, https://doi.org/10.1126/scisignal.aaf7584.Search in Google Scholar

Xue, D. R., Tian, F., Yang, F.H., Chen, H.M., Yuan, X.C., Yang, C.H., Chen, Y., Wang, Q., and He, C. (2018). Phosphodiesterase EdpX1 promotes virulence, exopolysaccharide production and biofilm formation in Xanthomonas oryzae pv. oryzae. Appl. Environ. Microbiol. 84: e01717–e01718, https://doi.org/10.1128/aem.01717-18.Search in Google Scholar

Yang, C., Cui, C., Ye, Q., Kan, J., Fu, S., Song, S., Huang, Y., He, F., Zhang, L.H., Jia, Y., et al. (2017). Burkholderia cenocepacia integrates cis-2-dodecenoic acid and cyclic dimeric guanosine monophosphate signals to control virulence. Proc. Natl. Acad. Sci. U. S. A. 114: 13006–13011, https://doi.org/10.1073/pnas.1709048114.Search in Google Scholar

Yang, F.H., Qian, S.S., Tian, F., Chen, H.M., Hutchins, W., Yang, C.-H., and He, C. (2016). The GGDEF domain protein GdpX1 attenuates motility, exopolysaccharide production, and virulence in Xanthomonas oryzae pv. oryzae. J. Appl. Microbiol. 120: 1646–1657, https://doi.org/10.1111/jam.13115.Search in Google Scholar

Yao, S.Y., Luo, L., Har, K.J., Becker, A., Rüberg, S., Yu, G.Q., Zhu, J.B., and Cheng, H.P. (2004). Sinorhizobium meliloti ExoR and ExoS proteins regulate both succinoglycan and flagellum production. J. Bacteriol. 186: 6042–6049, https://doi.org/10.1128/jb.186.18.6042-6049.2004.Search in Google Scholar

Yuan, X., Zeng, Q., Xu, J., Severin, G.B., Zhou, X., Waters, C.M., Sundin, G.W., Ibekwe, A.M., Liu, F., and Yang, C.H. (2020). Tricarboxylic acid (TCA) cycle enzymes and intermediates modulate intracellular cyclic di-GMP levels and the production of plant cell wall-degrading enzymes in soft rot pathogen Dickeya dadantii. Mol. Plant Microbe Interact. 33: 296–307, https://doi.org/10.1094/mpmi-07-19-0203-r.Search in Google Scholar

Zatakia, H.M., Nelson, C.E., Syed, U.J., and Scharf, B.E. (2014). ExpR coordinates the expression of symbiotically important, bundle-forming Flp pili with quorum sensing in Sinorhizobium meliloti. Appl. Environ. Microbiol. 80: 2429–2439, https://doi.org/10.1128/aem.04088-13.Search in Google Scholar

Zogaj, X., Nimtz, M., Rohde, M., Bokranz, W., and Römling, U. (2001). The multicellular morphotypes of Salmonella typhimurium and Escherichia coli produce cellulose as the second component of the extracellular matrix. Mol. Microbiol. 39: 1452–1463, https://doi.org/10.1046/j.1365-2958.2001.02337.x.Search in Google Scholar

Received: 2020-06-29
Accepted: 2020-09-14
Published Online: 2020-10-02
Published in Print: 2020-11-26

© 2020 Elizaveta Krol et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 28.3.2024 from https://www.degruyter.com/document/doi/10.1515/hsz-2020-0232/html
Scroll to top button