Background: The Aromatoleum/Azoarcus/Thauera (AAT) cluster comprises anaerobic degradation specialists (Aromatoleum, Thauera) and N2-fixing endophytes (Azoarcus). Omics-based and genetic studies with associated model strains implicate stringent response (SR) in adaptation to nutrient limitation and plant colonization. SR is well-studied in standard bacteria such as Escherichia coli and known as adaptive strategy to nutrient limitations by adjusting e.g., transcription and stress response. SR involves the alarmone (p)ppGpp, whose cellular level is controlled by the synthetases/hydrolases RelA/SpoT and the noncanonical transcription factor DksA, whose interaction with RNA polymerase (RNAP) binding of (p)ppGpp enhances. Summary: DksA-mediated SR occurs across Proteobacteria and other phylogenetic groups such as Myxococcia and Spirochaetia, mostly applying to pathogens. Furthermore, all three DksA variants (four, two or one cysteine residue(s) for Zn2+-binding) were found. Genes encoding SR components are present in all studied 37 genomes representing 31 species from the AAT cluster. Each genome encodes a synthesizing RelA, a hydrolyzing SpoT, a four cysteine-containing DksA, and mostly also a one cysteine-containing DksA. Opposing functions of RelA and SpoT in Aromatoleum aromaticum EbN1T, Aromatoleum sp. strain CIB, Azoarcus olearius BH72, and Thauera aromatica K172T (also entire AAT cluster) are implicated by full conservation of amino acids (E and D vs. 2H2D motif and ED diad) essential for catalysis by their synthetase versus hydrolase domains. Likewise, functionality of the predicted C4-type DksAs from these four model strains was visually assessed by structural modeling and comparison of key features (binding sites for Zn2+/(p)ppGpp; CC tip for RNAP interaction) to those of the available E. coli DksA cryo-EM structure. Key Messages: SR as a global adaptation strategy should contribute to the success of the AAT cluster in its distinct habitats: complex and highly variable soils/sediments (high molecular and microbial diversity, fluctuating nutrient availabilities and redox states) of the free-living degradation specialists versus the defined endorhizosphere (more stable conditions, less complex community) of the endophytes. A noteworthy exception is Aromatoleum sp. strain CIB by combining degradation and endophytic features. Thus, future investigations into the role of SR in the habitat success of such bacteria reflecting their divergent environmental niches are needed as well as promising.

The betaproteobacterial family Rhodocyclaceae harbors a distinct cluster of physiologically specialized, facultative anaerobic bacteria distributed across the three genera Aromatoleum, Azoarcus, and Thauera [1]. While members of the genera Aromatoleum [2] and Thauera [3] comprise anaerobic degradation specialists for hydrocarbons, aromatic compounds, and terpenes, those of the genus Azoarcus stand out for their N2-fixing, endophytic lifestyle in association with plant roots, e.g., of rice [4]. The diversity of the Aromatoleum/Azoarcus/Thauera (AAT) cluster is evident from its 9/4/18 taxonomically described species and 10/7/20 available genomes, with the first complete genomes reported for Aromatoleum aromaticum EbN1T [5] and Azoarcus olearius BH72 [6]. Moreover, a recent biogeographic meta-analysis showed that members of this cluster are globally distributed and occur in a large variety of terrestrial and aquatic habitats [7]. In particular, 4 members of the AAT cluster were developed over more than 20 years into well-studied model systems by investigating them with different as well as complementary foci of research topics and experimental approaches, as briefly summarized in the following.

Aromatoleum aromaticum EbN1T is the model for architecture and regulation of the catabolic network towards a systems biology-level understanding, with combination of physiology, molecular genetics, and proteogenomics. The strain was originally isolated from freshwater mud samples with ethylbenzene under nitrate-reducing conditions [8]. The research on its complete catabolic network - the degradation part comprises 31 peripheral and 4 central routes, overall recruiting 279 different proteins (178 identified) [5] - integrates several perspectives, such as: (i) reaction sequences and substrate-specific regulation of (novel) peripheral degradation routes [9, 10]; (ii) function of sensory/regulatory proteins and nano-molar thresholds for responsiveness towards aromatic substrates [11]; (iii) inducer-independent formation of uptake systems and degradation routes in response to nutrient limitation (“preparedness for nutritional opportunities”) [12]; (iv) systems biology-level studies including a genome-wide metabolic model [13]; (v) horizontal gene transfer as a shaping force for the strain’s genome and likely also the genus’s pan-genome [7]; and (vi) novel anaerobic enzymes, including C—H bond cleaving ethylbenzene dehydrogenase [14] and ATP-dependent acetophenone carboxylase [15].

Aromatoleum sp. strain CIB is the model for transcriptional regulation of aromatic compound degradation, with application of profound molecular genetics. The strain (formerly Azoarcus sp. strain CIB; taxonomic description is still awaited) was isolated with benzoate under nitrate-reducing conditions [16] from a culture derived from a fuel contaminated laboratory column [17]. Major insights into transcriptional regulation include: (i) the repressors BzdR and MbdR, which are responsive to benzoyl- versus 3-methylbenzoyl-CoA and hence control the bzd/box gene clusters (anaerobic/aerobic benzoate degradation) versus mbd genes (anaerobic 3-methylbenzoate degradation) [18, 19]; and (ii) the master regulator AccR, which represses bzd gene expression in the context of carbon catabolite repression [20, 21]. Recently, the capacity for colonizing the roots of rice was recognized [22], with the endophytic lifestyle possibly governed by the second messenger cyclic di-GMP (c-di-GMP) [23].

Azoarcus olearius BH72 is the original model for the N2-fixing, endophytic lifestyle within the AAT cluster, with integration of advanced microscopy, physiology, transcriptomics, and molecular genetics. The strain was isolated from surface-sterilized roots of Kallar grass from saline-sodic soil in Pakistan, using nitrogen-free, malate-containing mineral medium [4]; species name according to [24]. Insights into the strategies for endophytic diazotroph-cereal interactions include: (i) colonization of rice roots in high numbers [25], with microoxia in root aerenchyma favoring nif gene expression (nitrogenase) [26‒28]; (ii) adhesion to the root surface requires the main structural pilAB genes (type IV pili) [29], minor pilins [30], and several adhesion proteins [31]; (iii) ingress and systemic spreading inside the plant is facilitated by the endoglucanase EglA [32], as well as twitching (type VI pili) [33] and flagellar motility [34]; (iv) two signal transduction proteins (c-di-GMP signaling) and components of type VI protein secretion systems (T6SS) are involved in endophytic competence [30]; and (v) N-limitation induces downregulation of genetic information processing versus upregulation of flagella and pili formation [35], which is reminiscent of responses earlier linked with stringent response (SR) in E. coli [36].

Thauera aromatica K172T is the model for the intricate enzyme mechanisms of the novel reactions discovered in aromatic compound degradation, with application of sophisticated biochemical approaches. The strain was isolated from anaerobic sludge with phenol under nitrate-reducing conditions [37]. Examples of novel degradative enzymes and reactions are: (i) benzylsuccinate synthase, which homolytically cleaves a benzylic C—H bond of toluene to facilitate the stereoselective addition to fumarate yielding (R)-benzylsuccinate [38, 39]; (ii) a modified β-oxidation converts the latter to central benzoyl-CoA [40, 41]; (iii) phosphorylation [42] followed by carboxylation [43] initiate anaerobic degradation of phenol; (iv) α-oxidation is involved in phenylalanine conversion to phenylacetate [44]; (iv) reductive dearomatization via an ATP-dependent, Birch-like reaction converts central benzoyl-CoA to a cyclic dienoyl-CoA [45, 46], which is then further degraded (v) via a specific, modified β-oxidation [47, 48].

The strategies of a “preparedness for nutritional opportunities” apparently implemented by Ar. aromaticum EbN1T in response to nutrient limitation [12], as well as the involvement of c-di-GMP in response to exposure to rice plant (Oryza sativa) exudates [30] of and “preparedness for alternative N-source opportunities under N2-fixation” by Az. olearius BH72 [35] already suggested SR and/or similar global regulatory processes to play a role in adaptation to niche-specific environmental conditions. This notion was recently substantiated by studies with Aromatoleum sp. strain CIB on how its endophytic lifestyle could be controlled by c-di-GMP levels [23]. Furthermore, the disparate habitats of the free-living degradation specialists (high molecular and microbial diversity and fluctuations) and the N2-fixing endophytes (more stable conditions, less complex community) suggest distinct strategies for niche adaptation in the phylogenetically coherent AAT cluster. Against this background, the present study aims at briefly summarizing the state-of-the-art of SR to then bioinformatically define the genetic repertoire for conducting SR across the AAT cluster, in order to lay a foundation for future experimental investigations.

Role of (p)ppGpp ‒ Historical Perspective

Nearly all bacteria have to deal with the omnipresent risk of nutrient limitation. To overcome such periods of scarcity, bacteria have evolved regulatory mechanisms to globally adjust their metabolism. This involves downregulating processes such as DNA replication, transcription, translation, and monomer biosynthesis to maintain cellular balance while simultaneously activating amino acid biosynthesis, stress responses, and other cellular processes. This adaptive strategy, known as SR, is a conserved feature across bacterial species and has been extensively characterized in standard model microorganisms such as Escherichia coli. SR is regulated by the alarmones (second messengers) guanosine tetraphosphate (ppGpp) and guanosine pentaphosphate (pppGpp), collectively referred to as (p)ppGpp (Fig. 1a) [49]. These molecules were first detected as chromatographic signals, termed “Magic Spots” (MS I and MS II), in amino acid-starved auxotrophic E. coli strains and later identified [50‒52]. A key regulatory mechanism of (p)ppGpp involves its binding to RNA polymerase (RNAP) at two distinct sites: site 1 at the interface of the ω and β′ subunits of RNAP and site 2 at the interface of RNAP and the transcription factor DksA. (p)ppGpp-binding at both sites regulates RNAP through allosteric inhibition: in site 1 by hindering relative movement of RNAP subunits and in site 2 by stabilizing RNAP-DksA interactions in the secondary channel of RNAP, which leads to interference with its promoter interactions (Fig. 2a) [49, 53‒55]. Beyond RNAP, (p)ppGpp has multiple additional regulatory targets in the form of proteins that bind it directly [56].

Fig. 1.

Multifaceted regulatory and global metabolic control by SR in E. coli. a Structure of the alarmone (second messenger) (p)ppGpp. b Synthesis/hydrolysis of (p)ppGpp and affected processes. c Domain structures of the (p)ppGpp synthetases/hydrolases Rel, RelA, and SpoT. Abbreviations (alphabetic order): ACT, aspartate kinase, chorismate mutase, TyrA (prephenate dehydrogenase); AH/RIS, α-helical/ribosomal intersubunit; CTD, C-terminal domain; NTD, N-terminal domain; and TGS, ThrRS (threonine-tRNA synthetase), GTPase, SpoT. Light blue and dashed line indicate nonfunctional pseudoHD (hydrolase domain) in RelA; pale pink marks largely inactive SD (synthetase domain) in SpoT. Gray section within SD marks the α11-loop, which interlinks it to (pseudo)HD and carries the allosteric site of (p)ppGpp (red dot) in Rel and RelA. d Regulation of the synthetase activity of RelA off the ribosome versus by ternary interaction with deacylated tRNA and stalled ribosome. Subfigures were redrawn/modified from [49].

Fig. 1.

Multifaceted regulatory and global metabolic control by SR in E. coli. a Structure of the alarmone (second messenger) (p)ppGpp. b Synthesis/hydrolysis of (p)ppGpp and affected processes. c Domain structures of the (p)ppGpp synthetases/hydrolases Rel, RelA, and SpoT. Abbreviations (alphabetic order): ACT, aspartate kinase, chorismate mutase, TyrA (prephenate dehydrogenase); AH/RIS, α-helical/ribosomal intersubunit; CTD, C-terminal domain; NTD, N-terminal domain; and TGS, ThrRS (threonine-tRNA synthetase), GTPase, SpoT. Light blue and dashed line indicate nonfunctional pseudoHD (hydrolase domain) in RelA; pale pink marks largely inactive SD (synthetase domain) in SpoT. Gray section within SD marks the α11-loop, which interlinks it to (pseudo)HD and carries the allosteric site of (p)ppGpp (red dot) in Rel and RelA. d Regulation of the synthetase activity of RelA off the ribosome versus by ternary interaction with deacylated tRNA and stalled ribosome. Subfigures were redrawn/modified from [49].

Close modal
Fig. 2.

Role of DksA in SR in E. coli. a (p)ppGpp/DksA interactions with RNA polymerase (RNAP) and effects on transcription; site 1 of RNAP binds only (p)ppGpp, while site 2 accepts DksA and (p)ppGpp. b RNAP/DksA/(p)ppGpp indirectly affect the divisome and thereby the cell length. Subfigures (a) and (b) were redrawn/modified from [49, 53], respectively.

Fig. 2.

Role of DksA in SR in E. coli. a (p)ppGpp/DksA interactions with RNA polymerase (RNAP) and effects on transcription; site 1 of RNAP binds only (p)ppGpp, while site 2 accepts DksA and (p)ppGpp. b RNAP/DksA/(p)ppGpp indirectly affect the divisome and thereby the cell length. Subfigures (a) and (b) were redrawn/modified from [49, 53], respectively.

Close modal

Role of Rel, RelA, and SpoT

The levels of SR-triggering (p)ppGpp in E. coli fluctuate depending on environmental changes, starvation, and stress [57]. The turnover between (p)ppGpp and the nucleotides GTP and GDP is catalyzed by synthetase (SYNTH) domain (SD)-containing enzymes (formation) and by hydrolase domain (HD)-containing enzymes (degradation). The synthetases form (p)ppGpp by an ATP-dependent 3′-hydroxyl group phosphorylation of GTP or GDP, respectively, while hydrolases cleave off this inorganic diphosphate (Fig. 1a, b). These enzymes belong to the RelA/SpoT homolog (RSH) superfamily, which includes three main types: (i) long-RSH enzymes with both SD and HD in their N-terminal regions (NTD) and additional regulatory subdomains in the C-terminal region (CTD), such as the TGS domain for interaction with uncharged tRNA; (ii) small alarmone synthetases with only the SD; and (iii) small alarmone hydrolases with only the HD (Fig. 1c) [49, 58]. The inherent sequence similarity between the name-giving RelAs/SpoTs of the RSH superfamily and Rels can be attributed to the evolutionary origin of RelA and SpoT as differentiations of Rel after a gene duplication. The archetypical Rel contains both functional SD and HD where either one is exclusively activated by the C-terminal regulatory domains in response to starvation versus nutrient abundance. RelA and SpoT also contain both SD and HD; however, they play contrasting roles in E. coli. While RelA contains an inactive pseudoHD and mainly acts as a switched synthetase, SpoT has only weak synthetase activity and acts mainly as a switched hydrolase [59, 60]. The cellular homeostasis of (p)ppGpp is not only required for aligning SR with multifaceted physiological demands but also needs to be strictly maintained, since over-accumulation of (p)ppGpp is detrimental to the cell [61]. Several appropriate regulation systems, which are based on controlling the activities of RelA and SpoT and reveal also the complex interactions of SR with metabolism, are described in the following.

RelA Regulation

Control of the (p)ppGpp-synthesizing activity of RelA (Fig. 1d) starts from its closed conformation, where pseudoHD and CTD block spatially the catalytic SD such that its substrates can be accommodated only poorly. The ramping up of SD activity can proceed via two distinct and intriguing routes that are either ribosome-independent or, on the contrary, involve the stalled ribosome (with empty A-site): (i) upon binding of (p)ppGpp to its allosteric site at the α-helix, which interlinks pseudoHD and SD, the NTD adopts an open conformation releasing its spatial inhibition. This allows efficient entry of substrates and thereby turning on of the SD activity. The spatial interaction with the CTD remains preserved, however. (ii) Under conditions of amino acid starvation, closed RelAs interact with deacetylated tRNAs and stalled ribosomes, forming the so-called “starved” ribosomal complex. Here, the spatial inhibition of SD by CTD but not that by pseudoHD is released yielding an extended conformation of RelA primed for (p)ppGpp synthesis. Subsequently, binding of (p)ppGpp to its allosteric site lifts the inhibition by pseudoHD, which leads to the full activation of SD [60, 62‒65].

Activity of RelA and consequently (p)ppGpp homeostasis in E. coli can furthermore be controlled on two different levels with opposing effects, whereby SR is linked to N-metabolism: (i) in response to N-starvation, the hexameric NtrC response regulator of the NtrBC two-component system (controlling nitrogen assimilation) binds to an enhancer-like region upstream of the relA gene. This activates an until then inactive σ54-RNAP-promoter complex, which in turn amplifies the constitutive σ70-dependent relA gene expression and as a result formation of (p)ppGpp [36]. (ii) By contrast, the NirD protein (small subunit of cytoplasmic nitrite reductase, NirBD) can bind to the catalytic NTD of RelA, resulting in inhibition of (p)ppGpp synthesis [66].

SpoT Regulation

The activity of SpoT can be modulated by various means: (i) interaction of the SpoT-TGS domain with the acyl carrier protein results in turning off of the hydrolase function for the benefit of the synthesis function, such that under conditions of impaired fatty acid metabolism (acyl carrier protein loaded with short- rather than long-chain fatty acids), SR is switched on [67]. (ii) Conversely, under the condition of nutrient richness, the 50S ribosomal subunit associated GTPase (CgtA/ObgE) interacts with SpoT to maintain its (p)ppGpp-hydrolyzing activity, thereby suppressing SR and ensuring normal growth [68‒70]. (iii) When glucose depletion is succeeded by a shift to an alternative substrate, the Rsd protein (anti-σ activity) is released from the PTS component HPr to then associate with SpoT and stimulate its HD. This harnesses excessive formation of (p)ppGpp by RelA during such transitions in C-source utilization [71]. (iv) During fatty acid and phosphate starvation, the small YtfK protein interacts with the catalytic domains of SpoT shifting its activity from hydrolysis to synthesis [72]. Notably, expression of the ytfK gene is positively regulated by the cAMP-CRP complex relating it to glucose availability and linked catabolite repression [73]. (v) The YmgB protein binds directly to the TGS and helical domains of SpoT, resulting again in SpoT-dependent accumulation of (p)ppGpp [74]. Earlier structural-functional analyses of YmgB indicated a role as DNA-binding protein and involvement in acid resistance [75].

Role of DksA

DksA is an independent regulator of RNAP, whose function is significantly enhanced in the presence of (p)ppGpp. It belongs to a family of structurally similar proteins that regulate transcription not by binding to DNA but by interacting within the secondary channel of RNAP. Despite sharing a conserved coiled-coil (CC) core motif with similar amino acid residues at functional key positions, other notable members of this family, such as GreA and GreB, exhibit no overall sequence similarity with DksA. These proteins bind competitively to the RNAP secondary channel, where they fulfill both redundant and unique regulatory functions [76, 77]. One regulatory effect on transcription by DksA and Gre factors arises from their ability to bind to RNAP prior to transcription initiation (Fig. 2a), destabilizing the RNAP open complex and thus acting as negative regulators of transcription [78‒80]. Notably, binding of (p)ppGpp at site 2 between DksA and the β′rim helix of RNAP, forming a ternary complex (RNAP-DksA/(p)ppGpp), alters the location of DksA therein, which enhances the destabilizing effect of DksA on the interaction of RNAP open with rRNA gene promoters [81]. Additionally, DksA and Gre factors influence transcription by targeting non-transcribing elongation complexes. Specifically, Gre factors bind to elongation complexes, which are arrested because of misincorporation, and cleave nascent RNA to rescue the complex. Similarly, DksA can bind to stalled elongation complexes during starvation and rescue the complex through an as-yet-unknown mechanism [76, 82]. In the context of SR, DksA employs a second fundamental regulatory mechanism where it, together with (p)ppGpp, exerts either positive or negative control of transcription depending on the GC content of target promoters. Specifically, it enhances transcription from promoters with low GC content while repressing those with high GC content [83, 84]. In sum, the cooperative regulatory effects of DksA and site 2-bound (p)ppGpp outperform the effects of (p)ppGpp bound to site 1, giving rise to diverse regulatory answers to varying (p)ppGpp levels [55]. Regulatory targets of DksA/(p)ppGpp in SR help the cell to manage starvation stress and beyond that influence cell wall homeostasis and cell division (Fig. 2b) [53, 85].

Notably, the intracellular concentration of DksA remains constant under different conditions and throughout the cell’s life cycle [75]. However, oxidative stress can lead to oxidation of Zn2+-binding cysteines (yielding disulfide bonds), distortion of the protein’s globular structure, and thereby to dysfunctional DksA. Notably, in Salmonella (closely related to E. coli), the chaperone DnaJ was shown to interact with impaired DksA and restore its functionality by reducing the disulfide bonds [86].

Since its initial discovery and mechanistical elucidation in E. coli, SR has been recognized as a widespread regulatory process, found in symbiotic and pathogenic bacteria-host interactions, plants, and animals [87‒90]. While these examples are predominantly mediated by (p)ppGpp, DksA-mediated SR appears to be largely restricted to Proteobacteria, with a few notable exceptions. In Firmicutes such as B. subtilis rising (p)ppGpp levels correlate with reduced GTP levels, which in turn impede RNA synthesis through this limitation in the NTP pool [91, 92]. In Archaea stringent-like phenotypes have been observed; however, these cases are mediated by neither DksA or (p)ppGpp [93, 94]. In contrast to these SR mechanisms in Firmicutes, Archaea, and eukaryotes, DksA-mediated SR in Proteobacteria represents a more direct and specific transcriptional control mechanism (see above paragraph). Table 1 [95‒122] shows a selection of organisms in which the phenotypic response of DksA-mediated SR has been studied. The diverse phenotypic responses show the lifestyle-dependent role of DksA and (p)ppGpp beyond nutrient limitation-triggered SR. It ranges from virulence control in pathogenic bacteria to facilitating symbiosis in plant symbionts, general stress response and resistance, phage-host interactions, and to being essential for viability in Myxococcus xanthus. A particularity in some organisms is the presence of two or more DksA paralogs that differ in the cysteine content of the Zn2+-binding site. DksA phylotypes with four, two, or one conserved cysteine(s) have been reported, where the differing Zn2+-binding kinetics directly influence the regulatory response to oxidative and nitrosative stress [104, 123, 124]. Rhodobacter sphaeroides encodes a four cysteine (C4) and a one cysteine (C1) DksA paralog. While the C1 DksA is formed during aerobic and anaerobic growth, the C4 DksA is only induced under anoxic conditions. Furthermore, the C1 DksA is necessary for photosynthetic growth, and its function can be complemented by the C4 DksA of E. coli and vice versa. The regulatory role of the C4 paralog, which cannot supplement DksA in E. coli, is unclear with deletion mutants having no specific phenotype [97].

Table 1.

Dissemination of DksA-mediated SR and phenotypic response

OrganismLifestyleDksA typeaExper. evidencePhenotypic responseRef.
Alphaproteobacteria 
Bartonella henselae Pathogen (human) C1 Gene KO 
  • Type IV secretion system

  • Overall virulence

 
[95
Sinorhizobium meliloti Plant symbiont C4 + C1 Gene KO 
  • Formation of root nodules

  • Symbiosis

 
[96
Rhodobacter sphaeroides Aquatic, photosynthetic C4 + C1 Gene KO 
  • Photosynthetic growth

  • Fatty acid biosynthesis

  • Amino acid uptake

 
[97
Gammaproteobacteria 
Escherichia coli Opportunistic pathogen (human) C4 Crystal structure 
  • Cell division

  • Inorganic polyphosphate accumulation

  • Antibiotic resistance

  • Glucose-phosphate stress recovery

 
[98
[99
[100
[101
E. coli – phage P1 interaction Bacteriophage   
  • Lysis genes (holin, lysozyme)

  • Lysis prevention (antiholin)

  • Replication genes (repA)

  • Lytic phage development

 
[102
Salmonella spp. Pathogen (human) C4 Gene KO 
  • Virulence pathways

  • Metabolic pathways

  • Type 1 fimbriae

  • Quorum sensing (AI-2 system)

  • Nitrosative stress response

  • Intestinal colonization

  • Biofilm formation

 
[103
[104
[105
Acinetobacter baumannii Opportunistic pathogen (human) C4 Gene KO 
  • Zn resistance

  • Cu resistance

  • Aromatic compound degradation

  • Oxidative stress tolerance

  • Virulence

  • Antibiotic resistance

  • Efflux pump genes

 
[106
[107
Yersinia enterocolitica Gastrointestinal foodborne pathogen (human) C4 Gene KO 
  • Temperature-dependent cell wall maintenance through fatty acid metabolism

  • Antibiotic resistance

  • Ribosomal gene expression

  • Uptake and utilization of amino acids

  • Chemotaxis and flagella formation

 
[108
[109
Vibrio cholerae Pathogen (human) C4 Gene KO 
  • Toxin production

 
[110
Haemophilus ducreyi Pathogen (human) C4 + C4 Gene KO 
  • Oxidative stress resistance (phagocyte evasion)

  • Attachment to host cells

  • Overall virulence

 
[111
Shigella flexneri Pathogen (human) C4 Gene KO 
  • Type III secretion system

  • Overall virulence

 
[112
Pseudomonas aeruginosa Broad host pathogen C4 + C2 Crystal structure 
  • Oxidative stress tolerance, virulence and biofilm formation

  • Swarming motility

  • Anaerobic respiration

  • Polyamine synthesis

 
[113
[114
PDB: 4IJJ 
Pseudomonas plecoglossicida Pathogen (fish) C4 Gene silencing (RNA interference) 
  • Flagellum assembly

  • Ribosome assembly

 
[115
Erwinia amylovora Pathogen (plant) C4 Gene KO 
  • Type III secretion system

  • Hypersensitive response (HR)

  • Overall virulence

 
[116
Xanthomonas citri Pathogen (plant) C1 Gene KO 
  • Type II and III secretion system

  • Virulence

  • Nutrient uptake (TonB-dependent transporters)

  • Siderophore production

 
[98
Pseudomonas putida Soil C4 Gene KO 
  • Starvation-induced biofilm dispersal

  • c-di-GMP metabolism

  • Planktonic lifestyle

 
[117
Azotobacter vinelandii Soil, N2-fixing C4 + C4(+C2) Gene KO 
  • PHB synthesis

 
[118
Myxococcia 
Myxococcus xanthus Saprophytic C4 (+ C4 + C3 + C2) Artificial regulation 
  • Essential for growth

 
[119
Chlamydiaceae 
Chlamydia trachomatis Pathogen (human) C1 Crystal structure 
  • Reticulate body – elementary body transition (virulence)

  • Overall virulence

 
[120
PDB: 6PTG 
Spirochaetia 
Borrelia burgdorferi Pathogen (human) C4 Gene KO 
  • Sigma factor RpoS degradation

  • pH inducible expression of genes under RpoS control

  • Expression of virulence factors

  • Osmotic stress response

  • Virulence

 
[121
[122
OrganismLifestyleDksA typeaExper. evidencePhenotypic responseRef.
Alphaproteobacteria 
Bartonella henselae Pathogen (human) C1 Gene KO 
  • Type IV secretion system

  • Overall virulence

 
[95
Sinorhizobium meliloti Plant symbiont C4 + C1 Gene KO 
  • Formation of root nodules

  • Symbiosis

 
[96
Rhodobacter sphaeroides Aquatic, photosynthetic C4 + C1 Gene KO 
  • Photosynthetic growth

  • Fatty acid biosynthesis

  • Amino acid uptake

 
[97
Gammaproteobacteria 
Escherichia coli Opportunistic pathogen (human) C4 Crystal structure 
  • Cell division

  • Inorganic polyphosphate accumulation

  • Antibiotic resistance

  • Glucose-phosphate stress recovery

 
[98
[99
[100
[101
E. coli – phage P1 interaction Bacteriophage   
  • Lysis genes (holin, lysozyme)

  • Lysis prevention (antiholin)

  • Replication genes (repA)

  • Lytic phage development

 
[102
Salmonella spp. Pathogen (human) C4 Gene KO 
  • Virulence pathways

  • Metabolic pathways

  • Type 1 fimbriae

  • Quorum sensing (AI-2 system)

  • Nitrosative stress response

  • Intestinal colonization

  • Biofilm formation

 
[103
[104
[105
Acinetobacter baumannii Opportunistic pathogen (human) C4 Gene KO 
  • Zn resistance

  • Cu resistance

  • Aromatic compound degradation

  • Oxidative stress tolerance

  • Virulence

  • Antibiotic resistance

  • Efflux pump genes

 
[106
[107
Yersinia enterocolitica Gastrointestinal foodborne pathogen (human) C4 Gene KO 
  • Temperature-dependent cell wall maintenance through fatty acid metabolism

  • Antibiotic resistance

  • Ribosomal gene expression

  • Uptake and utilization of amino acids

  • Chemotaxis and flagella formation

 
[108
[109
Vibrio cholerae Pathogen (human) C4 Gene KO 
  • Toxin production

 
[110
Haemophilus ducreyi Pathogen (human) C4 + C4 Gene KO 
  • Oxidative stress resistance (phagocyte evasion)

  • Attachment to host cells

  • Overall virulence

 
[111
Shigella flexneri Pathogen (human) C4 Gene KO 
  • Type III secretion system

  • Overall virulence

 
[112
Pseudomonas aeruginosa Broad host pathogen C4 + C2 Crystal structure 
  • Oxidative stress tolerance, virulence and biofilm formation

  • Swarming motility

  • Anaerobic respiration

  • Polyamine synthesis

 
[113
[114
PDB: 4IJJ 
Pseudomonas plecoglossicida Pathogen (fish) C4 Gene silencing (RNA interference) 
  • Flagellum assembly

  • Ribosome assembly

 
[115
Erwinia amylovora Pathogen (plant) C4 Gene KO 
  • Type III secretion system

  • Hypersensitive response (HR)

  • Overall virulence

 
[116
Xanthomonas citri Pathogen (plant) C1 Gene KO 
  • Type II and III secretion system

  • Virulence

  • Nutrient uptake (TonB-dependent transporters)

  • Siderophore production

 
[98
Pseudomonas putida Soil C4 Gene KO 
  • Starvation-induced biofilm dispersal

  • c-di-GMP metabolism

  • Planktonic lifestyle

 
[117
Azotobacter vinelandii Soil, N2-fixing C4 + C4(+C2) Gene KO 
  • PHB synthesis

 
[118
Myxococcia 
Myxococcus xanthus Saprophytic C4 (+ C4 + C3 + C2) Artificial regulation 
  • Essential for growth

 
[119
Chlamydiaceae 
Chlamydia trachomatis Pathogen (human) C1 Crystal structure 
  • Reticulate body – elementary body transition (virulence)

  • Overall virulence

 
[120
PDB: 6PTG 
Spirochaetia 
Borrelia burgdorferi Pathogen (human) C4 Gene KO 
  • Sigma factor RpoS degradation

  • pH inducible expression of genes under RpoS control

  • Expression of virulence factors

  • Osmotic stress response

  • Virulence

 
[121
[122

KO, knock out.

aDksA type refers to the number of conserved cysteines in the Zn2+-binding motive.

RelA and SpoT

Analyzing the genomes of representative members of the AAT cluster revealed the seamless presence of the hydrolases/synthetases RelA and SpoT with one coding gene each per genome (Fig. 3). The orthologs of the cluster share similar, high sequence identities with their counterparts in E. coli (RelA, 41–44%; SpoT, 48–51%). As RelA and SpoT are the only RSH family proteins detected in the analyzed genomes of the AAT cluster, it can be assumed that RelA acts as the predominant (p)ppGpp synthetase with a presumable inactive pseudoHD and, conversely, SpoT as the predominant (p)ppGpp hydrolase with only weak synthetase activity (see below section Phylogeny, Domain Architecture, Similarity, and Catalytic Residues of RelA and SpoT). As expected for betaproteobacteria, no truly bifunctional Rel or monofunctional small alarmone synthetase and small alarmone hydrolase [49] are encoded by our studied organisms.

Fig. 3.

Dissemination of the SR components (DksA, RelA, and SpoT) across genome-sequenced members of the AAT cluster. Phylotype color coding reflects the phylogenetic clustering of the DksAs (online suppl. Fig. S1), correlating with the number of cysteine residues involved in binding of the Zn2+ cation.

Fig. 3.

Dissemination of the SR components (DksA, RelA, and SpoT) across genome-sequenced members of the AAT cluster. Phylotype color coding reflects the phylogenetic clustering of the DksAs (online suppl. Fig. S1), correlating with the number of cysteine residues involved in binding of the Zn2+ cation.

Close modal

DksA

The studied genomes of the AAT cluster contain at least one copy of the dksA gene, with many species harboring additional paralogs (Fig. 3). These DksA paralogs correspond to three phylotypes (online suppl. Fig. S1; for all online suppl. material, see https://doi.org/10.1159/000546200), differentiated by the cysteine composition of their Zn2+-binding sites and sequence similarity to well-characterized DksA proteins:

The four cysteine-containing C4 phylotype is conserved across all studied genomes and the representatives share 72–83% sequence similarity to DksA of E. coli (DksAEco). The C4-type retains all functionally relevant amino acids of the Zn2+- and (p)ppGpp-binding sites as well as the RNAP interaction site (CC tip) described for the DksAEco, suggesting functional equivalence (for details see below section Visualizing Structural Features of DksA Serving Key Functions). This phylotype is presumed to play a similar role in transcriptional regulation and SR across species. In Figures 3 and online supplementary S1, DksAs belonging to this phylotype are indicated in blue.

The C1 phylotype, characterized by a single cysteine in the Zn2+-binding site, clusters with DksA proteins from Bartonella henselae and Rhodobacter sphaeroides (see above section General Dissemination of DksA-Mediated SR). Sequence identity to DksAEco lies between 48 and 65%. It is the second most prevalent type found in the majority of studied species from the AAT cluster. Members of this phylotype exhibit a characteristic lysine-to-glutamine substitution at the equivalent K139 position within the (p)ppGpp-binding site of DksAEco, while the RNAP-interacting loop residues remain conserved. Functional evidence suggests that this phylotype performs similar regulatory roles. For instance, the single C1-type DksA protein in Bartonella henselae mediates virulence by regulating the type IV secretion system (T4SS) in response to starvation [95]. Similarly, one of the two DksA homologs in Rhodobacter sphaeroides (C1-type) is essential for photosynthetic growth and is functionally interchangeable with DksAEco [97]. In Figures 3 and online supplementary S1, DksAs belonging to this phylotype are labeled in green.

A second C4-type DksA occurs exclusively in species of the genus Azoarcus within the studied AAT cluster. Despite retaining functionally relevant amino acids within the Zn2+-binding and RNAP interaction site, key amino acids involved in (p)ppGpp-binding are not conserved. Functional assignment of this phylotype is inconclusive, as the function of the representative second C4-type DksA of Rhodobacter sphaeroides is not elucidated [97]. They share 31–34% sequence identity to DksAEco. In Figures 3 and 4, DksAs belonging to this phylotype are labeled in red.

Fig. 4.

Comparison of selected RelAs and SpoTs. a Phylogenetic clustering and domain architecture of proteins from Ar. aromaticum EbN1T, Aromatoleum sp. strain CIB, Az. olearius BH72, and T. aromatica K172T, compared to the orthologs from E. coli K12. Accession numbers of proteins are indicated. For abbreviations of domains see legend to Fig. 1c. b Percent identity matrix of HD (top) and SD (bottom) from the same selected four strains of the AAT cluster compared to those of E. coli. The nonfunctional RelA-HD of E. coli is indicated by light gray background.

Fig. 4.

Comparison of selected RelAs and SpoTs. a Phylogenetic clustering and domain architecture of proteins from Ar. aromaticum EbN1T, Aromatoleum sp. strain CIB, Az. olearius BH72, and T. aromatica K172T, compared to the orthologs from E. coli K12. Accession numbers of proteins are indicated. For abbreviations of domains see legend to Fig. 1c. b Percent identity matrix of HD (top) and SD (bottom) from the same selected four strains of the AAT cluster compared to those of E. coli. The nonfunctional RelA-HD of E. coli is indicated by light gray background.

Close modal

As outlined in the Introduction, within the AAT cluster, the four species Ar. aromaticum EbN1T, Aromatoleum sp. strain CIB, Az. olearius BH72, and T. aromatica K172T are well-studied model organisms that represent the phylogenetic and phenotypic spectrum of the cluster. Moreover, their seeming “preparedness for nutritional/nitrogen opportunities” and use of c-di-GMP for plant colonization suggest second messenger-dependent regulatory processes such as SR to be also employed by these nonstandard model bacteria. Therefore, the predicted structures of their RelA, SpoT, and DksA proteins were inspected more closely to assess their functionality and thereby the capacity of the strains to perform SR.

Phylogeny, Domain Architecture, Similarity, and Catalytic Residues of RelA and SpoT

The RelAs and SpoTs from Ar. aromaticum EbN1T, Aromatoleum sp. strain CIB, Az. olearius BH72, and T. aromatica K172T mirror the phylogenetic relations of the strains and share the same composition and order of domains (hydrolase, synthetase, TGS, AH/RIS, and ACT) as known from the experimentally studied E. coli orthologs (Fig. 4a) [49]. Accordingly, percent identity matrices (on the amino acid level) in relation to the E. coli orthologs revealed solid-to-high values for the HDs (63.3–66.0% for SpoTs vs. 34.0–35.3% for RelAs) and the SDs (58.8–62.7% for SpoTs vs. 50.0–54.5% for RelAs) (Fig. 4b). Thus, these predicted RelAs and SpoTs in the four selected model strains should be functional.

To further substantiate this potential functionality, we aligned the sequences of HDs/SDs of SpoTs (Fig. 5a, [125‒127]) and RelAs (Fig. 5b), respectively, from the four model strains with those from E. coli. Thereby, we could verify the full conservation of all residues that are essential for catalysis (Fig. 5a bottom). In the catalytic center of the SpoT-HD, the 2H2D (His, Asp) motif and ED (Glu-Asp) diad coordinate Mn2+ and Mg2+, respectively, which jointly interact with the δε-PP group at the 3′-position of (p)ppGpp. Furthermore, the Mg2+ ion activates a H2O molecule for a nucleophilic attack, resulting in the hydrolytic removal of this diphosphate [125]. In the SpoT-SD, the conserved D and E residues provide the essential carboxyl groups that coordinate the catalytic Mg2+ ion. Here, the metal ion catalyzes the transfer of the βγ-PP from ATP to the 3′-hydroxyl group of the ribose moiety in GTP or GDP forming (p)ppGpp [126]. Taken together, the SpoTs from our model strains should indeed act as bifunctional hydrolase/synthetases, as typical for Beta- and Gammaproteobacteria but not for monofunctional hydrolase SpoTs in Moraxellaceae [58]. By contrast, the RelAs of the four model strains lack the 2H2D motif and ED diad generating a pseudoHD, while they harbor the conserved D and E residues in the SDs. Thus, the RelAs should serve the role of monofunctional synthetases. Noteworthy, multiple protein sequence alignments (online suppl. Fig. S2) revealed that all of these residues, which are indispensable for the catalytic activity of RelAs and SpoTs, are also fully conserved across the aforementioned 31 genome-sequenced members of the AAT cluster. However, the regulatory proteins Rsd, YtfK, and YmbG, which modulate the activities of RelA and SpoT in E. coli, are not encoded in any of the studied genomes of the AAT cluster.

Fig. 5.

Multiple sequence alignment of bi/monofunctional SpoTs (a) and RelAs (b) from E. coli K12, Ar. aromaticum EbN1T, Aromatoleum sp. strain CIB, Az. olearius BH72, and T. aromatica K172T (all truncated after SD). Order of strains is according to phylogenic relation of proteins. Role of essential residues in the catalytic centers of the HDs and SDs is illustrated in the schemes below the SpoT alignments; red arrows indicate nucleophilic attack (modified from refs. [125‒127]).

Fig. 5.

Multiple sequence alignment of bi/monofunctional SpoTs (a) and RelAs (b) from E. coli K12, Ar. aromaticum EbN1T, Aromatoleum sp. strain CIB, Az. olearius BH72, and T. aromatica K172T (all truncated after SD). Order of strains is according to phylogenic relation of proteins. Role of essential residues in the catalytic centers of the HDs and SDs is illustrated in the schemes below the SpoT alignments; red arrows indicate nucleophilic attack (modified from refs. [125‒127]).

Close modal

Visualizing Structural Features of DksA Serving Key Functions

The crystal and cryo-EM structures of DksAEco revealed that this all α-helical protein is composed of three major structural elements [128, 129] (Fig. 6, left panel, top): (i) the globular (G) domain, which comprises the N- and C-terminal regions; (ii) the central, CC domain consisting of two α-helices connected via an α-turn, which harbors the two highly conserved, functionally indispensable tip residues (D74, A76) that interact with the trigger loop/helix of RNAP [55, 130]; and (iii) a C-terminal α-helix (CT domain), which is loosely connected to the G and CC domains. The G and CT domains each contribute two cysteine residues to the Zn2+-binding site, which likely constrains the mobility of the α-helix of the CT domain. Binary and ternary 3D structures of DksAEco (with and without (p)ppGpp) in complex with the RNAP holoenzyme binding to the rRNA gene promoter from E. coli were determined based on X-ray crystal analysis [81] and cryo-EM [129]. DksAEco apparently uses in total four conserved residues from its CT (K139) and CC (L95, K98, and R129) domains to establish contacts with the 5′-phosphates and the guanine moiety of (p)ppGpp. This interaction with (p)ppGpp induces a rotation of DksAEco, which drives the CC tip deeper into the secondary channel of RNAP strengthening the DksA-RNAP interaction and leading to an interference of the CC tip with the bridging helix of RNAP. This in turn alters the kinetics of DNA loading, impedes open complex formation, and favors the less stable early stage of the RNAP-promoter complex [81, 129].

Fig. 6.

3D-structural visualization of selected DksAs. The left panel displays the overall 3D structure of DksA: top, from E. coli as determined by cryo-EM [129]; bottom, overlaid with those from Ar. aromaticum EbN1T and Az. olearius BH72 as predicted by AlphaFold. The right panel presents close-up views of the Zn2+- and (p)ppGpp-binding sites as well as the CC tip interacting with RNAP for E. coli [129], compared to those of the AlphaFold-predicted DksAs from Ar. aromaticum EbN1T, Aromatoleum sp. strain CIB, Az. olearius BH72, and T. aromatica K172T.

Fig. 6.

3D-structural visualization of selected DksAs. The left panel displays the overall 3D structure of DksA: top, from E. coli as determined by cryo-EM [129]; bottom, overlaid with those from Ar. aromaticum EbN1T and Az. olearius BH72 as predicted by AlphaFold. The right panel presents close-up views of the Zn2+- and (p)ppGpp-binding sites as well as the CC tip interacting with RNAP for E. coli [129], compared to those of the AlphaFold-predicted DksAs from Ar. aromaticum EbN1T, Aromatoleum sp. strain CIB, Az. olearius BH72, and T. aromatica K172T.

Close modal

The AlphaFold-predicted structures of the DksAs from our four model strains show the same overall architectural features as the DksAEco cryo-EM structure, as shown for the DksAs from Ar. aromaticum EbN1T and Az. olearius BH72 (Fig. 6, left panel, bottom). Moreover, close-up views of the Zn2+- and (p)ppGpp-binding sites as well of the CC tips of the DksAs from all four model bacteria correlated also very well with those of DksAEco (Fig. 6, right panel). Thus, these key structural DksA features from the four selected members of the AAT cluster can be assumed to function similarly in negatively regulating initiation of transcription.

Albeit SR has been studied in-depth for E. coli, this regulatory mechanism was observed in a taxonomically broad range of bacteria. As most of the functional studies were derived for pathogens, it would be rewarding to study functions and targets of SR for completely different environmental niches. Bioinformatic analyses allowed us to postulate that functional modules of SR are widespread across the AAT cluster, in which SR is only poorly investigated let alone understood. Unraveling SR responses in model bacteria of this group could help clarify some fundamental questions: Since SR is commonly regarded as stress response in comparison to our rich laboratory conditions, are soils and sediments generally stressful for bacteria in their environmental niche? Are roots as habitats for beneficial endophytes really the nutrient-rich niche as commonly thought? What roles does SR play in such scenarios beyond stress response?

Phylogenetic and Bioinformatic Analyses

Sequences of RelAs, SpoTs, and DksAs were obtained from RefSeq [131] and aligned using Clustal Omega [132] while simultaneously obtaining phylogenetic trees.

Modeling of DksA Structures

Structural modeling of DksAs from Ar. aromaticum EbN1T, Aromatoleum sp. CIB, Az. olearius BH72, and T. aromatica K172T were performed using AlphaFold Server version 3 [133]. To improve the accuracy and relevance of the predicted structures, co-structure modeling of DksA with all RNAP subunits of the respective organisms was employed. Including the RNAP constituents in the modeling query enhanced the structural context of the interaction site and the overall conformational fidelity of DksA. Post-prediction, the generated 3D models were analyzed to assess their structural quality and accuracy by evaluating the predicted aligned errors, predicted template modeling scores, and interface predicted template modeling scores provided by AlphaFold. The structural models with the highest interface predicted template modeling scores (all 0.81) and lowest predicted aligned errors (14.29‒15.84) were selected for further analysis. Visualizations and structural alignments were performed using UCSF ChimeraX and the integrated MatchMaker tool [134, 135].

Barbara Reinhold-Hurek and Ralf Rabus declare Editorial Board membership with Microbial Physiology. Patrick Becker, Jakob Ruickoldt, and Petra Wendler have no conflicts of interest to declare.

Bioinformatic analyses and manuscript preparation were conducted with no extramural funding.

P.B., B.R.-H., and R.R. conceived the study and finalized the manuscript and all authors agreed to its final version. P.B. performed the bioinformatic analysis and literature searches and drafted the manuscript; and J.R., P.W., and P.B. conducted the structural modeling.

1.
Garrity
GM
,
Bell
JA
,
Lilburn
T
.
Family I. Rhodocyclaceae fam. nov
.
In: Brenner DJ, Krieg NR, Staley JT, Garrity GM (eds), Bergey’s Manual of Systematic Bacteriology, 2nd edn, vol2
,
(The Proteobacteria), part C (The Alpha-, Beta-, Delta-, and Epsilonbacteria)
.
New York
:
Springer
;
2005
. p.
887
.
2.
Rabus
R
,
Wöhlbrand
L
,
Thies
D
,
Meyer
M
,
Reinhold-Hurek
B
,
Kämpfer
P
.
Aromatoleum gen. nov., a novel genus accommodating the phylogenetic lineage including Azoarcus evansii and related species, and proposal of Aromatoleum aromaticum sp. nov., Aromatoleum petrolei sp. nov., Aromatoleum bremense sp. nov., Aromatoleum toluolicum sp. nov. and Aromatoleum diolicum sp. nov
.
Int J Syst Evol Microbiol
.
2019
;
69
(
4
):
982
97
.
3.
Anders
H-J
,
Kaetzke
A
,
Kämpfer
P
,
Ludwig
W
,
Fuchs
G
.
Taxonomic position of aromatic-degrading denitrifying pseudomonad strains K 172 and KB 740 and their description as new members of the genera Thauera, as Thauera aromatica sp. nov., and Azoarcus, as Azoarcus evansii sp. nov., respectively, members of the beta subclass of the Proteobacteria
.
Int J Syst Bacteriol
.
1995
;
45
(
2
):
327
33
.
4.
Reinhold-Hurek
B
,
Hurek
T
,
Gillis
M
,
Hoste
B
,
Vancanneyt
M
,
Kersters
K
, et al
.
Azoarcus gen. nov., nitrogen-fixing proteobacteria associated with roots of Kallar grass (Leptochloa fusca (L.) Kunth), and description of two species, Azoarcus indigens sp. nov. and Azoarcus communis sp. nov
.
Int J Syst Bacteriol
.
1993
;
43
(
3
):
574
84
.
5.
Becker
P
,
Wünsch
D
,
Wöhlbrand
L
,
Neumann-Schaal
M
,
Schomburg
D
,
Rabus
R
.
The catabolic network of Aromatoleum aromaticum EbN1T
.
Microb Physiol
.
2024
;
34
:
1
77
.
6.
Krause
A
,
Ramakumar
A
,
Bartels
D
,
Battistoni
F
,
Bekel
T
,
Boch
J
, et al
.
Complete genome of the mutualistic, N2-fixing grass endophyte Azoarcus sp. strain BH72
.
Nat Biotechnol
.
2006
;
24
(
11
):
1385
91
.
7.
Becker
P
,
Döhmann
A
,
Wöhlbrand
L
,
Thies
D
,
Hinrichs
C
,
Buschen
R
, et al
.
Complex and flexible catabolism in Aromatoleum aromaticum pCyN1
.
Environ Microbiol
.
2022
;
24
(
7
):
3195
211
.
8.
Rabus
R
,
Widdel
F
.
Anaerobic degradation of ethylbenzene and other aromatic hydrocarbons by new denitrifying bacteria
.
Arch Microbiol
.
1995
;
163
(
2
):
96
103
.
9.
Wöhlbrand
L
,
Kallerhoff
B
,
Lange
D
,
Hufnagel
P
,
Thiermann
J
,
Reinhardt
R
, et al
.
Functional proteomic view of metabolic regulation in “Aromatoleum aromaticum” strain EbN1
.
Proteomics
.
2007
;
7
(
13
):
2222
39
.
10.
Trautwein
K
,
Wilkes
H
,
Rabus
R
.
Proteogenomic evidence for β-oxidation of plant-derived 3-phenylpropanoids in “Aromatoleum aromaticum” EbN1
.
Proteomics
.
2012
;
12
(
9
):
1402
13
.
11.
Buschen
R
,
Lambertus
P
,
Scheve
S
,
Horst
S
,
Song
F
,
Wöhlbrand
L
, et al
.
Sensitive and selective phenol sensing in denitrifying Aromatoleum aromaticum EbN1T
.
Microbiol Spectr
.
2023
;
11
(
6
):
e0210023
23
.
12.
Trautwein
K
,
Lahme
S
,
Wöhlbrand
L
,
Feenders
C
,
Mangelsdorf
K
,
Harder
J
, et al
.
Physiological and proteomic adaptation of “Aromatoleum aromaticum” EbN1 to low growth rates in benzoate-limited, anoxic chemostats
.
J Bacteriol
.
2012
;
194
(
9
):
2165
80
.
13.
Becker
P
,
Kirstein
S
,
Wünsch
D
,
Koblitz
J
,
Buschen
R
,
Wöhlbrand
L
, et al
.
Systems biology of aromatic compound catabolism in facultative anaerobic Aromatoleum aromaticum EbN1T
.
mSystems
.
2022
;
7
(
6
):
e0068522
22
.
14.
Kloer
DP
,
Hagel
C
,
Heider
J
,
Schulz
GE
.
Crystal structure of ethylbenzene dehydrogenase from Aromatoleum aromaticum
.
Structure
.
2006
;
14
(
9
):
1377
88
.
15.
Weidenweber
S
,
Schühle
K
,
Demmer
U
,
Warkentin
E
,
Ermler
U
,
Heider
J
.
Structure of the acetophenone carboxylase core complex: prototype of a new class of ATP-dependent carboxylases/hydrolases
.
Sci Rep
.
2017
;
7
:
39674
.
16.
López Barragán
MJ
,
Carmona
M
,
Zamarro
MT
,
Thiele
B
,
Boll
M
,
Fuchs
G
, et al
.
The bzd gene cluster, coding for anaerobic benzoate catabolism, in Azoarcus sp. strain CIB
.
J Bacteriol
.
2004
;
186
(
17
):
5762
74
.
17.
Hess
A
,
Zarda
B
,
Hahn
D
,
Häner
A
,
Stax
D
,
Höhener
P
, et al
.
In situ analysis of denitrifying toluene- and m-xylene-degrading bacteria in a diesel fuel-contaminated laboratory aquifer column
.
Appl Environ Microbiol
.
1997
;
63
(
6
):
2136
41
.
18.
Barragán
MJL
,
Blázquez
B
,
Zamarro
MT
,
Mancheño
JM
,
García
JL
,
Díaz
E
, et al
.
BzdR, a repressor that controls the anaerobic catabolism of benzoate in Azoarcus sp. CIB, is the first member of a new subfamily of transcriptional regulators
.
J Biol Chem
.
2005
;
280
(
11
):
10683
94
.
19.
Durante-Rodríguez
G
,
Valderrama
JA
,
Mancheño
JM
,
Rivas
G
,
Alfonso
C
,
Arias-Palomo
E
, et al
.
Biochemical characterization of the transcriptional regulator BzdR from Azoarcus sp. CIB
.
J Biol Chem
.
2010
;
285
(
46
):
35694
705
.
20.
Valderrama
JA
,
Shingler
V
,
Carmona
M
,
Díaz
E
.
AccR is a master regulator involved in carbon catabolite repression of the anaerobic catabolism of aromatic compounds in Azoarcus sp
.
J Biol Chem
.
2014
;
289
(
4
):
1892
904
.
21.
Valderrama
JA
,
Gómez-Álvarez
H
,
Martín-Moldes
Z
,
Berbís
,
Cañada
FJ
,
Durante-Rodríguez
G
, et al
.
A novel redox-sensing histidine kinase that controls carbon catabolite repression in Azoarcus sp
.
mBio
.
2019
;
10
(
2
):
e00059-19
.
22.
Fernández
H
,
Prandoni
N
,
Fernández-Pascual
M
,
Fajardo
S
,
Morcillo
C
,
Díaz
E
, et al
.
Azoarcus sp. CIB, an anaerobic biodegrader of aromatic compounds shows an endophytic lifestyle
.
PLoS One
.
2014
;
9
(
10
):
e110771
.
23.
Fernández-Llamosas
H
,
Díaz
E
,
Carmona
M
.
Motility, adhesion and c-di-GMP influence the endophytic colonization of rice by Azoarcus sp. CIB
.
Microorganisms
.
2021
;
9
(
3
):
554
.
24.
Chen
M-H
,
Sheu
S-Y
,
James
EK
,
Young
C-C
,
Chen
W-M
.
Azoarcus olearius sp. nov., a nitrogen-fixing bacterium isolated from oil-contaminated soil
.
Int J Syst Evol Microbiol
.
2013
;
63
(
Pt 10
):
3755
61
.
25.
Reinhold-Hurek
B
,
Hurek
T
.
Living inside plants: bacterial endophytes
.
Curr Opin Plant Biol
.
2011
;
14
(
4
):
435
43
.
26.
Hurek
T
,
Egener
T
,
Reinhold-Hurek
B
.
Divergence in nitrogenases of Azoarcus spp., Proteobacteria of the β subclass
.
J Bacteriol
.
1997
;
179
(
13
):
4172
8
.
27.
Egener
T
,
Hurek
T
,
Reinhold-Hurek
B
.
Use of green fluorescent protein to detect expression of nif genes of Azoarcus sp. BH72, a grass-associated diazotroph, on rice roots
.
Mol Plant-Microbe Interact
.
1998
;
11
(
1
):
71
5
.
28.
Jiang
X
,
Pees
T
,
Reinhold-Hurek
B
.
Deep-learning-based removal of autofluorescence and fluorescence quantification in plant-colonizing bacteria in vivo
.
New Phytol
.
2022
;
235
(
6
):
2481
95
.
29.
Dörr
J
,
Hurek
T
,
Reinhold‐Hurek
B
.
Type IV pili are involved in plant-microbe and fungus-microbe interactions
.
Mol Microbiol
.
1998
;
30
(
1
):
7
17
.
30.
Shidore
T
,
Dinse
T
,
Öhrlein
J
,
Becker
A
,
Reinhold-Hurek
B
.
Transcriptomic analysis of responses to exudates reveal genes required for rhizosphere competence of the endophyte Azoarcus sp. strain BH72
.
Environ Microbiol
.
2012
;
14
(
10
):
2775
87
.
31.
Sarkar
A
,
Marszalkowska
M
,
Schäfer
M
,
Pees
T
,
Klingenberg
H
,
Macht
F
, et al
.
Global expression analysis of the response to microaerobiosis reveals an important cue for endophytic establishment of Azoarcus sp. BH72
.
Environ Microbiol
.
2017
;
19
(
1
):
198
217
.
32.
Reinhold-Hurek
B
,
Maes
T
,
Gemmer
S
,
van Montagu
M
,
Hurek
T
.
An endoglucanase is involved in infection of rice roots by the not-cellulose-metabolizing endophyte Azoarcus sp. strain BH72
.
Mol Plant-Microbe Interact
.
2006
;
19
(
2
):
181
8
.
33.
Böhm
M
,
Hurek
T
,
Reinhold-Hurek
B
.
Twitching motility is essential for endophytic rice colonization by the N2-fixing endophyte Azoarcus sp. strain BH72
.
Mol Plant-Microbe Interact
.
2007
;
20
(
5
):
526
33
.
34.
Buschart
A
,
Sachs
S
,
Chen
X
,
Herglotz
J
,
Krause
A
,
Reinhold-Hurek
B
.
Flagella mediate endophytic competence rather than act as MAMPS in rice-Azoarcus sp. strain BH72 interactions
.
Mol Plant-Microbe Interact
.
2012
;
25
(
2
):
191
9
.
35.
Solaiyappan Mani
S
,
Reinhold-Hurek
B
.
RNA-Seq provides new insights into the gene expression changes in Azoarcus olearius BH72 under nitrogen-deficient and replete conditions beyond the nitrogen fixation process
.
Microorganisms
.
2021
;
9
:
1888
.
36.
Brown
DR
,
Barton
G
,
Pan
Z
,
Buck
M
,
Wigneshweraraj
S
.
Nitrogen stress response and stringent response are coupled in Escherichia coli
.
Nat Commun
.
2014
;
5
:
4115
.
37.
Tschech
A
,
Fuchs
G
.
Anaerobic degradation of phenol by pure cultures of newly isolated denitrifying pseudomonads
.
Arch Microbiol
.
1987
;
148
(
3
):
213
7
.
38.
Leuthner
B
,
Leutwein
C
,
Schulz
H
,
Hörth
P
,
Haehnel
W
,
Schiltz
E
, et al
.
Biochemical and genetic characterization of benzylsuccinate synthase from Thauera aromatica: a new glycyl radical enzyme catalysing the first step in anaerobic toluene metabolism
.
Mol Microbiol
.
1998
;
28
(
3
):
615
28
.
39.
Seyhan
D
,
Friedrich
P
,
Szaleniec
M
,
Hilberg
M
,
Buckel
W
,
Golding
BT
, et al
.
Elucidating the stereochemistry of enzymatic benzylsuccinate synthesis with chirally labeled toluene
.
Angew Chem Int Ed Engl
.
2016
;
55
(
38
):
11664
7
.
40.
Leutwein
C
,
Heider
J
.
Anaerobic toluene-catabolic pathway in denitrifying Thauera aromatica: activation and β-oxidation of the first intermediate, (R)-(+)-benzylsuccinate
.
Microbiology
.
1999
;
145 (Pt 11)
:
3265
71
.
41.
Weidenweber
S
,
Schühle
K
,
Lippert
M-L
,
Mock
J
,
Seubert
A
,
Demmer
U
, et al
.
Finis tolueni: a new type of thiolase with an integrated Zn-finger subunit catalyzes the final step of anaerobic toluene metabolism
.
FEBS J
.
2022
;
289
(
18
):
5599
616
.
42.
Schmeling
S
,
Narmandakh
A
,
Schmitt
O
,
Gad’on
N
,
Schühle
K
,
Fuchs
G
.
Phenylphosphate synthase: a new phosphotransferase catalyzing the first step in anaerobic phenol metabolism in Thauera aromatica
.
J Bacteriol
.
2004
;
186
(
23
):
8044
57
.
43.
Schühle
K
,
Fuchs
G
.
Phenylphosphate carboxylase: a new C-C lyase involved in anaerobic phenol metabolism in Thauera aromatica
.
J Bacteriol
.
2004
;
186
(
14
):
4556
67
.
44.
Schneider
S
,
Fuchs
G
.
Phenylacetyl-CoA:acceptor oxidoreductase, a new α-oxidizing enzyme that produces phenylglyoxylate. Assay, membrane localization, and differential production in Thauera aromatica
.
Arch Microbiol
.
1998
;
169
(
6
):
509
16
.
45.
Unciuleac
M
,
Boll
M
.
Mechanism of ATP-driven electron transfer catalyzed by the benzene ring-reducing enzyme benzoyl-CoA reductase
.
Proc Natl Acad Sci USA
.
2001
;
98
:
13619
24
.
46.
Möbitz
H
,
Friedrich
T
,
Boll
M
.
Substrate binding and reduction of benzoyl-CoA reductase: evidence for nucleotide-dependent conformational changes
.
Biochemistry
.
2004
;
43
(
5
):
1376
85
.
47.
Boll
M
,
Laempe
D
,
Eisenreich
W
,
Bacher
A
,
Mittelberger
T
,
Heinze
J
, et al
.
Nonaromatic products from anoxic conversion of benzoyl-CoA with benzoyl-CoA reductase and cyclohexa-1,5-diene-1-carbonyl-CoA hydratase
.
J Biol Chem
.
2000
;
275
(
29
):
21889
95
.
48.
Dörner
E
,
Boll
M
.
Properties of 2-oxo-glutarate:ferredoxin oxidoreductase from Thauera aromatica and its role in enzymatic reduction of the aromatic ring
.
J Bacteriol
.
2002
;
184
(
14
):
3975
83
.
49.
Bange
G
,
Brodersen
DE
,
Liuzzi
A
,
Steinchen
W
.
Two P or not two P: understanding regulation by the bacterial second messengers (p)ppGpp
.
Annu Rev Microbiol
.
2021
;
75
:
383
406
.
50.
Cashel
M
.
The control of ribonucleic acid synthesis in Escherichia coli. IV. Relevance of unusual phosphorylated compounds from amino acid-starved stringent strains
.
J Biol Chem
.
1969
;
244
(
12
):
3133
41
.
51.
Cashel
M
,
Gallant
J
.
Two compounds implicated in the function of the RC gene of Escherichia coli
.
Nature
.
1969
;
221
(
5183
):
838
41
.
52.
Cashel
M
,
Kalbacher
B
.
The control of ribonucleic acid synthesis in Escherichia coli. V. Characterization of a nucleotide associated with the stringent response
.
J Biol Chem
.
1970
;
245
(
9
):
2309
18
.
53.
Anderson
SE
,
Vadia
SE
,
McKelvy
J
,
Levin
PA
.
The transcription factor DksA exerts opposing effects on cell division depending on the presence of ppGpp
.
mBio
.
2023
;
14
(
6
):
e0242523
23
.
54.
Ross
W
,
Vrentas
CE
,
Sanchez-Vazquez
P
,
Gaal
T
,
Gourse
RL
.
The magic spot: a ppGpp binding site on E. coli RNA polymerase responsible for regulation of transcription initiation
.
Mol Cell
.
2013
;
50
(
3
):
420
9
.
55.
Ross
W
,
Sanchez-Vazquez
P
,
Chen
AY
,
Lee
J-H
,
Burgos
HL
,
Gourse
RL
.
ppGpp binding to a site at the RNAP-DksA interface accounts for its dramatic effects on transcription initiation during the stringent response
.
Mol Cell
.
2016
;
62
(
6
):
811
23
.
56.
Gourse
RL
,
Chen
AY
,
Gopalkrishnan
S
,
Sanchez-Vazquez
P
,
Myers
A
,
Ross
W
.
Transcriptional responses to ppGpp and DksA
.
Annu Rev Microbiol
.
2018
;
72
:
163
84
.
57.
Dworkin
J
,
Harwood
CS
.
Metabolic reprogramming and longevity in quiescence
.
Annu Rev Microbiol
.
2022
;
76
:
91
111
.
58.
Atkinson
GC
,
Tenson
T
,
Hauryliuk
V
.
The RelA/SpoT homolog (RSH) superfamily: distribution and functional evolution of ppGpp synthetases and hydrolases across the tree of life
.
PLoS OnE
.
2011
;
6
(
8
):
e23479
.
59.
Hauryliuk
V
,
Atkinson
GC
,
Murakami
KS
,
Tenson
T
,
Gerdes
K
.
Recent functional insights into the role of (p)ppGpp in bacterial physiology
.
Nat Rev Microbiol
.
2015
;
13
(
5
):
298
309
.
60.
Sinha
AK
,
Winther
KS
.
The RelA hydrolase domain acts as a molecular switch for (p)ppGpp synthesis
.
Commun Biol
.
2021
;
4
(
1
):
434
.
61.
Tedin
K
,
Bremer
H
.
Toxic effects of high levels of ppGpp in Escherichia coli are relieved by rpoB mutations
.
J Biol Chem
.
1992
;
267
(
4
):
2337
44
.
62.
Brown
A
,
Fernández
IS
,
Gordiyenko
Y
,
Ramakrishnan
V
.
Ribosome-dependent activation of stringent control
.
Nature
.
2016
;
534
(
7606
):
277
80
.
63.
Loveland
AB
,
Bah
E
,
Madireddy
R
,
Zhang
Y
,
Brilot
AF
,
Grigorieff
N
, et al
.
Ribosome RelA structures reveal the mechanism of stringent response activation
.
Elife
.
2016
;
5
:
e17029
.
64.
Pausch
P
,
Abdelshahid
M
,
Steinchen
W
,
Schäfer
H
,
Gratani
FL
,
Freibert
S-A
, et al
.
Structural basis for regulation of the opposing (p)ppGpp synthetase and hydrolase within the stringent response orchestrator Rel
.
Cell Rep
.
2020
;
32
(
11
):
108157
.
65.
Roghanian
M
,
Van Nerom
K
,
Takada
H
,
Caballero-Montes
J
,
Tamman
H
,
Kudrin
P
, et al
.
(p)ppGpp controls stringent factors by exploiting antagonistic allosteric coupling between catalytic domains
.
Mol Cell
.
2021
;
81
(
16
):
3310
22.e6
.
66.
Léger
L
,
Byrne
D
,
Guiraud
P
,
Germain
E
,
Maisonneuve
E
.
NirD curtails the stringent response by inhibiting RelA activity in Escherichia coli
.
Elife
.
2021
;
10
:
e64092
.
67.
Battesti
A
,
Bouveret
E
.
Acyl carrier protein/SpoT interaction, the switch linking SpoT-dependent stress response to fatty acid metabolism
.
Mol Microbiol
.
2006
;
62
(
4
):
1048
63
.
68.
Wout
P
,
Pu
K
,
Sullivan
SM
,
Reese
V
,
Zhou
S
,
Lin
B
, et al
.
The Escherichia coli GTPase CgtAE cofractionates with the 50S ribosomal subunit and interacts with SpoT, a ppGpp synthetase/hydrolase
.
J Bacteriol
.
2004
;
186
(
16
):
5249
57
.
69.
Raskin
DM
,
Judson
N
,
Mekalanos
JJ
.
Regulation of the stringent response is the essential function of the conserved bacterial G protein CgtA in Vibrio cholerae
.
Proc Natl Acad Sci USA
.
2007
;
104
(
11
):
4636
41
.
70.
Persky
NS
,
Ferullo
DJ
,
Cooper
DL
,
Moore
HR
,
Lovett
ST
.
The ObgE/CgtA GTPase influences the stringent response to amino acid starvation in Escherichia coli
.
Mol Microbiol
.
2009
;
73
(
2
):
253
66
.
71.
Lee
J-W
,
Park
Y-H
,
Seok
Y-J
.
Rsd balances (p)ppGpp level by stimulating the hydrolase activity of SpoT during carbon source downshift in Escherichia coli
.
Proc Natl Acad Sci USA
.
2018
;
115
(
29
):
E6845
54
.
72.
Germain
E
,
Guiraud
P
,
Byrne
D
,
Douzi
B
,
Djendli
M
,
Maisonneuve
E
.
YtfK activates the stringent response by triggering the alarmone synthetase SpoT in Escherichia coli
.
Nat Commun
.
2019
;
10
(
1
):
5763
.
73.
Meyer
L
,
Germain
E
,
Maisonneuve
E
.
Regulation of ytfK by cAMP-CRP contributes to SpoT-dependent accumulation of (p)ppGpp in response to carbon starvation
.
Front Microbiol
.
2021
;
12
:
775164
.
74.
Guiraud
P
,
Germain
E
,
Byrne
D
,
Maisonneuve
E
.
The YmgB-SpoT interaction triggers the stringent response in Escherichia coli
.
J Biol Chem
.
2023
;
299
(
12
):
105429
.
75.
Lee
J
,
Page
R
,
García-Contreras
R
,
Palermino
J-M
,
Zhang
X-S
,
Doshi
O
, et al
.
Structure and function of the Escherichia coli protein YmgB: a protein critical for biofilm formation and acid-resistance
.
J Mol Biol
.
2007
;
373
(
1
):
11
26
.
76.
Furman
R
,
Sevostyanova
A
,
Artsimovitch
I
.
Transcription initiation factor DksA has diverse effects on RNA chain elongation
.
Nucleic Acids Res
.
2012
;
40
(
8
):
3392
402
.
77.
Vinella
D
,
Potrykus
K
,
Murphy
H
,
Cashel
M
.
Effects on growth by changes of the balance between GreA, GreB, and DksA suggest mutual competition and functional redundancy in Escherichia coli
.
J Bacteriol
.
2012
;
194
(
2
):
261
73
.
78.
Paul
BJ
,
Barker
MM
,
Ross
W
,
Schneider
DA
,
Webb
C
,
Foster
JW
, et al
.
DksA: a critical component of the transcription initiation machinery that potentiates the regulation of rRNA promoters by ppGpp and the initiating NTP
.
Cell
.
2004
;
118
(
3
):
311
22
.
79.
Rutherford
ST
,
Lemke
JJ
,
Vrentas
CE
,
Gaal
T
,
Ross
W
,
Gourse
RL
.
Effects of DksA, GreA, and GreB on transcription initiation: insights into the mechanisms of factors that bind in the secondary channel of RNA polymerase
.
J Mol Biol
.
2007
;
366
(
4
):
1243
57
.
80.
Stumper
SK
,
Ravi
H
,
Friedman
LJ
,
Mooney
RA
,
Corrêa
IR
Jnr
,
Gershenson
A
, et al
.
Delayed inhibition mechanism for secondary channel factor regulation of ribosomal RNA transcription
.
Elife
.
2019
;
8
:
e40576
.
81.
Molodtsov
V
,
Sineva
E
,
Zhang
L
,
Huang
X
,
Cashel
M
,
Ades
SE
, et al
.
Allosteric effector ppGpp potentiates the inhibition of transcript initiation by DksA
.
Mol Cell
.
2018
;
69
(
5
):
828
39.e5
.
82.
Toulmé
F
,
Mosrin-Huaman
C
,
Sparkowski
J
,
Das
A
,
Leng
M
,
Rahmouni
AR
.
GreA and GreB proteins revive backtracked RNA polymerase in vivo by promoting transcript trimming
.
EMBO J
.
2000
;
19
(
24
):
6853
9
.
83.
Haugen
SP
,
Ross
W
,
Gourse
RL
.
Advances in bacterial promoter recognition and its control by factors that do not bind DNA
.
Nat Rev Microbiol
.
2008
;
6
(
7
):
507
19
.
84.
Sanchez-Vazquez
P
,
Dewey
CN
,
Kitten
N
,
Ross
W
,
Gourse
RL
.
Genome-wide effects on Escherichia coli transcription from ppGpp binding to its two sites on RNA polymerase
.
Proc Natl Acad Sci USA
.
2019
;
116
(
17
):
8310
9
.
85.
Brückner
S
,
Müller
F
,
Schadowski
L
,
Kalle
T
,
Weber
S
,
Marino
EC
, et al
.
(p)ppGpp and moonlighting RNases influence the first step of lipopolysaccharide biosynthesis in Escherichia coli
.
microLife
.
2023
;
4
:
uqad031
.
86.
Kim
J-S
,
Liu
L
,
Fitzsimmons
LF
,
Wang
Y
,
Crawford
MA
,
Mastrogiovanni
M
, et al
.
DksA-DnaJ redox interactions provide a signal for the activation of bacterial RNA polymerase
.
Proc Natl Acad Sci USA
.
2018
;
115
(
50
):
E11780
9
.
87.
Prusa
J
,
Zhu
DX
,
Stallings
CL
.
The stringent response and Mycobacterium tuberculosis pathogenesis
.
Pathog Dis
.
2018
;
76
(
5
):
fty054
.
88.
Ito
D
,
Kawamura
H
,
Oikawa
A
,
Ihara
Y
,
Shibata
T
,
Nakamura
N
, et al
.
ppGpp functions as an alarmone in metazoa
.
Commun Biol
.
2020
;
3
(
1
):
671
.
89.
Pérez-Giménez
J
,
Iturralde
ET
,
Torres Tejerizo
G
,
Quelas
JI
,
Krol
E
,
Borassi
C
, et al
.
A stringent-response-defective Bradyrhizobium diazoefficiens strain does not activate the type 3 secretion system, elicits an early plant defense response, and circumvents NH4NO3-induced inhibition of nodulation
.
Appl Environ Microbiol
.
2021
;
87
(
9
):
e02989
20
.
90.
Mehrez
M
,
Romand
S
,
Field
B
.
New perspectives on the molecular mechanisms of stress signalling by the nucleotide guanosine tetraphosphate (ppGpp), an emerging regulator of photosynthesis in plants and algae
.
New Phytol
.
2023
;
237
(
4
):
1086
99
.
91.
Lopez
JM
,
Dromerick
A
,
Freese
E
.
Response of guanosine 5'-triphosphate concentration to nutritional changes and its significance for Bacillus subtilis sporulation
.
J Bacteriol
.
1981
;
146
(
2
):
605
13
.
92.
Pulschen
AA
,
Sastre
DE
,
Machinandiarena
F
,
Crotta Asis
A
,
Albanesi
D
,
de Mendoza
D
, et al
.
The stringent response plays a key role in Bacillus subtilis survival of fatty acid starvation
.
Mol Microbiol
.
2017
;
103
(
4
):
698
712
.
93.
Cimmino
C
,
Scoarughi
GL
,
Donini
P
.
Stringency and relaxation among the halobacteria
.
J Bacteriol
.
1993
;
175
(
20
):
6659
62
.
94.
Cellini
A
,
Scoarughi
GL
,
Poggiali
P
,
Santino
I
,
Sessa
R
,
Donini
P
, et al
.
Stringent control in the archaeal genus Sulfolobus
.
Res Microbiol
.
2004
;
155
(
2
):
98
104
.
95.
Québatte
M
,
Dick
MS
,
Kaever
V
,
Schmidt
A
,
Dehio
C
.
Dual input control: activation of the Bartonella henselae VirB/D4 type IV secretion system by the stringent sigma factor RpoH1 and the BatR/BatS two-component system
.
Mol Microbiol
.
2013
;
90
(
4
):
756
75
.
96.
Wippel
K
,
Long
SR
.
Contributions of Sinorhizobium meliloti transcriptional regulator DksA to bacterial growth and efficient symbiosis with Medicago sativa
.
J Bacteriol
.
2016
;
198
(
9
):
1374
83
.
97.
Lennon
CW
,
Lemmer
KC
,
Irons
JL
,
Sellman
MI
,
Donohue
TJ
,
Gourse
RL
, et al
.
A Rhodobacter sphaeroides protein mechanistically similar to Escherichia coli DksA regulates photosynthetic growth
.
mBio
.
2014
;
5
(
3
):
e01105
1114
.
98.
Zhang
Y
,
Teper
D
,
Xu
J
,
Wang
N
.
Stringent response regulators (p)ppGpp and DksA positively regulate virulence and host adaptation of Xanthomonas citri
.
Mol Plant Pathol
.
2019
;
20
(
11
):
1550
65
.
99.
Gray
MJ
.
Interactions between DksA and stress-responsive alternative sigma factors control inorganic polyphosphate accumulation in Escherichia coli
.
J Bacteriol
.
2020
;
202
(
14
):
e00133-20
.
100.
Wang
J
,
Cao
L
,
Yang
X
,
Wu
Q
,
Lu
L
,
Wang
Z
.
Transcriptional analysis reveals the critical role of RNA polymerase-binding transcription factor, DksA, in regulating multi-drug resistance of Escherichia coli
.
Int J Antimicrob Agents
.
2018
;
52
(
1
):
63
9
.
101.
Kessler
JR
,
Cobe
BL
,
Richards
GR
.
Stringent response regulators contribute to recovery from glucose phosphate stress in Escherichia coli
.
Appl Environ Microbiol
.
2017
;
83
(
24
):
e01636-17
.
102.
Cech
GM
,
Szalewska-Pałasz
A
,
Potrykus
K
,
Kloska
A
.
Virus-host interaction gets curiouser and curiouser. PART II: functional transcriptomics of the E. coli DksA-deficient cell upon phage P1vir infection
.
Int J Mol Sci
.
2021
;
22
(
11
):
6159
.
103.
Cohen
H
,
Adani
B
,
Cohen
E
,
Piscon
B
,
Azriel
S
,
Desai
P
, et al
.
The ancestral stringent response potentiator, DksA has been adapted throughout Salmonella evolution to orchestrate the expression of metabolic, motility, and virulence pathways
.
Gut Microbes
.
2022
;
14
(
1
):
1997294
.
104.
Crawford
MA
,
Henard
CA
,
Tapscott
T
,
Porwollik
S
,
McClelland
M
,
Vázquez-Torres
A
.
DksA-dependent transcriptional regulation in Salmonella experiencing nitrosative stress
.
Front Microbiol
.
2016
;
7
:
444
.
105.
Azriel
S
,
Goren
A
,
Rahav
G
,
Gal-Mor
O
.
The stringent response regulator DksA is required for Salmonella enterica serovar typhimurium growth in minimal medium, motility, biofilm formation, and intestinal colonization
.
Infect Immun
.
2016
;
84
(
1
):
375
84
.
106.
Maharjan
RP
,
Sullivan
GJ
,
Adams
FG
,
Shah
BS
,
Hawkey
J
,
Delgado
N
, et al
.
DksA is a conserved master regulator of stress response in Acinetobacter baumannii
.
Nucleic Acids Res
.
2023
;
51
(
12
):
6101
19
.
107.
Kim
N
,
Son
J-H
,
Kim
K
,
Kim
H-J
,
Shin
M
,
Lee
J-C
.
DksA modulates antimicrobial susceptibility of Acinetobacter baumannii
.
Antibiotics
.
2021
;
10
(
12
):
1472
.
108.
Huang
C
,
Li
W
,
Chen
J
.
Stringent response factor DksA contributes to fatty acid degradation function to influence cell membrane stability and polymyxin B resistance of Yersinia enterocolitica
.
Int J Mol Sci
.
2023
;
24
(
15
):
11951
.
109.
Huang
C
,
Li
W
,
Chen
J
.
Transcriptomic analysis reveals key roles of (p)ppGpp and DksA in regulating metabolism and chemotaxis in Yersinia enterocolitica
.
Int J Mol Sci
.
2023
;
24
(
8
):
7612
.
110.
Basu
P
,
Bhadra
RK
.
Post-transcriptional regulation of cholera toxin production in Vibrio cholerae by the stringent response regulator DksA
.
Microbiology
.
2019
;
165
(
1
):
102
12
.
111.
Holley
CL
,
Zhang
X
,
Fortney
KR
,
Ellinger
S
,
Johnson
P
,
Baker
B
, et al
.
DksA and (p)ppGpp have unique and overlapping contributions to Haemophilus ducreyi pathogenesis in humans
.
Infect Immun
.
2015
;
83
(
8
):
3281
92
.
112.
Sharma
AK
,
Payne
SM
.
Induction of expression of hfq by DksA is essential for Shigella flexneri virulence
.
Mol Microbiol
.
2006
;
62
(
2
):
469
79
.
113.
Fortuna
A
,
Collalto
D
,
Schiaffi
V
,
Pastore
V
,
Visca
P
,
Ascenzioni
F
, et al
.
The Pseudomonas aeruginosa DksA1 protein is involved in H2O2 tolerance and within-macrophages survival and can be replaced by DksA2
.
Sci Rep
.
2022
;
12
(
1
):
10404
.
114.
Min
KB
,
Yoon
SS
.
Transcriptome analysis reveals that the RNA polymerase-binding protein DksA1 has pleiotropic functions in Pseudomonas aeruginosa
.
J Biol Chem
.
2020
;
295
(
12
):
3851
64
.
115.
Wang
L-Y
,
Liu
Z-X
,
Zhao
L-M
,
Huang
L-X
,
Qin
Y-X
,
Su
Y-Q
, et al
.
Dual RNA-seq provides novel insight into the roles of dksA from Pseudomonas plecoglossicida in pathogen-host interactions with large yellow croakers (Larimichthys crocea)
.
Zool Res
.
2020
;
41
(
4
):
410
22
.
116.
Ancona
V
,
Lee
JH
,
Chatnaparat
T
,
Oh
J
,
Hong
J-I
,
Zhao
Y
.
The bacterial alarmone (p)ppGpp activates the type III secretion system in Erwinia amylovora
.
J Bacteriol
.
2015
;
197
(
8
):
1433
43
.
117.
Díaz-Salazar
C
,
Calero
P
,
Espinosa-Portero
R
,
Jiménez-Fernández
A
,
Wirebrand
L
,
Velasco-Domínguez
MG
, et al
.
The stringent response promotes biofilm dispersal in Pseudomonas putida
.
Sci Rep
.
2017
;
7
(
1
):
18055
.
118.
Ortiz-Vasco
CC
,
Moreno
S
,
Quintero-Navarro
LA
,
Rojo-Rodríguez
JB
,
Espín
G
.
The stringent response regulates the poly-β-hydroxybutyrate (PHB) synthesis in Azotobacter vinelandii
.
PLoS One
.
2024
;
19
(
4
):
e0299640
.
119.
García-Moreno
D
,
Polanco
MC
,
Navarro-Avilés
G
,
Murillo
FJ
,
Padmanabhan
S
,
Elías-Arnanz
M
.
A vitamin B12-based system for conditional expression reveals dksA to be an essential gene in Myxococcus xanthus
.
J Bacteriol
.
2009
;
191
(
9
):
3108
19
.
120.
Mandel
C
,
Yang
H
,
Buchko
GW
,
Abendroth
J
,
Grieshaber
N
,
Chiarelli
T
, et al
.
Expression and structure of the Chlamydia trachomatis DksA ortholog
.
Pathog Dis
.
2022
;
80
(
1
):
ftac007
.
121.
Boyle
WK
,
Richards
CL
,
Dulebohn
DP
,
Zalud
AK
,
Shaw
JA
,
Lovas
S
, et al
.
DksA-dependent regulation of RpoS contributes to Borrelia burgdorferi tick-borne transmission and mammalian infectivity
.
PLoS Pathog
.
2021
;
17
(
2
):
e1009072
.
122.
Mason
C
,
Thompson
C
,
Ouyang
Z
.
DksA plays an essential role in regulating the virulence of Borrelia burgdorferi
.
Mol Microbiol
.
2020
;
114
(
1
):
172
83
.
123.
Henard
CA
,
Tapscott
T
,
Crawford
MA
,
Husain
M
,
Doulias
P-T
,
Porwollik
S
, et al
.
The 4-cysteine zinc-finger motif of the RNA polymerase regulator DksA serves as a thiol switch for sensing oxidative and nitrosative stress
.
Mol Microbiol
.
2014
;
91
(
4
):
790
804
.
124.
Crawford
MA
,
Tapscott
T
,
Fitzsimmons
LF
,
Liu
L
,
Reyes
AM
,
Libby
SJ
, et al
.
Redox-active sensing by bacterial DksA transcription factors is determined by cysteine and zinc content
.
mBio
.
2016
;
7
(
2
):
e02161
.
125.
Zhou
RW
,
Manisa
B
,
Wang
B
.
A binuclear metallohydrolase model for RelA/SpoT-Homolog (RSH) hydrolases
.
J Biol Chem
.
2024
;
300
(
11
):
107841
.
126.
Patil
PR
,
Vithani
N
,
Singh
V
,
Kumar
A
,
Prakash
B
.
A revised mechanism for (p)ppGpp synthesis by Rel proteins: the critical role of the 2’-OH of GTP
.
J Biol Chem
.
2020
;
295
(
37
):
12851
67
.
127.
Hogg
T
,
Mechold
U
,
Malke
H
,
Cashel
M
,
Hilgenfeld
R
.
Conformational antagonism between opposing active sites in a bifunctional RelA/SpoT homolog modulates (p)ppGpp metabolism during the stringent response
.
Cell
.
2004
;
117
(
1
):
57
68
.
128.
Perederina
A
,
Svetlov
V
,
Vassylyeva
MN
,
Tahirov
TH
,
Yokoyama
S
,
Artsimovitch
I
, et al
.
Regulation through the secondary channel-structural framework for ppGpp-DksA synergism during transcription
.
Cell
.
2004
;
118
(
3
):
297
309
.
129.
Shin
Y
,
Qayyum
MZ
,
Pupov
D
,
Esyunina
D
,
Kulbachinskiy
A
,
Murakami
KS
.
Structural basis of ribosomal RNA transcription regulation
.
Nat Commun
.
2021
;
12
(
1
):
528
.
130.
Lennon
CW
,
Ross
W
,
Martin-Tumasz
S
,
Toulokhonov
I
,
Vrentas
CE
,
Rutherford
ST
, et al
.
Direct interactions between the coiled-coil tip of DksA and the trigger loop of RNA polymerase mediate transcriptional regulation
.
Genes Dev
.
2012
;
26
(
23
):
2634
46
.
131.
O’Leary
NA
,
Wright
MW
,
Brister
JR
,
Ciufo
S
,
Haddad
D
,
McVeigh
R
, et al
.
Reference sequence (RefSeq) database at NCBI: current status, taxonomic expansion, and functional annotation
.
Nucleic Acids Res
.
2016
;
44
(
D1
):
D733
45
.
132.
Sievers
F
,
Wilm
A
,
Dineen
D
,
Gibson
TJ
,
Karplus
K
,
Li
W
, et al
.
Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega
.
Mol Syst Biol
.
2011
;
7
:
539
.
133.
Abramson
J
,
Adler
J
,
Dunger
J
,
Evans
R
,
Green
T
,
Pritzel
A
, et al
.
Accurate structure prediction of biomolecular interactions with AlphaFold3
.
Nature
.
2024
;
630
(
8016
):
493
500
.
134.
Meng
EC
,
Pettersen
EF
,
Couch
GS
,
Huang
CC
,
Ferrin
TE
.
Tools for integrated sequence-structure analysis with UCSF Chimera
.
BMC Bioinformatics
.
2006
;
7
:
339
.
135.
Meng
EC
,
Goddard
TD
,
Pettersen
EF
,
Couch
GS
,
Pearson
ZJ
,
Morris
JH
, et al
.
UCSF ChimeraX: tools for structure building and analysis
.
Protein Sci
.
2023
;
32
(
11
):
e4792
.