#PAGE_PARAMS# #ADS_HEAD_SCRIPTS# #MICRODATA#

Integrative and quantitative view of the CtrA regulatory network in a stalked budding bacterium


Authors: Oliver Leicht aff001;  Muriel C. F. van Teeseling aff001;  Gaël Panis aff002;  Celine Reif aff001;  Heiko Wendt aff001;  Patrick H. Viollier aff002;  Martin Thanbichler aff001
Authors place of work: Department of Biology, University of Marburg, Marburg, Germany aff001;  Department of Microbiology and Molecular Medicine, Institute of Genetics and Genomics in Geneva (iGE3), Faculty of Medicine/CMU, University of Geneva, Geneve, Switzerland aff002;  Center for Synthetic Microbiology (SYNMIKRO), Marburg, Germany aff003;  Max Planck Institute for Terrestrial Microbiology, Marburg, Germany aff004
Published in the journal: Integrative and quantitative view of the CtrA regulatory network in a stalked budding bacterium. PLoS Genet 16(4): e32767. doi:10.1371/journal.pgen.1008724
Category: Research Article
doi: https://doi.org/10.1371/journal.pgen.1008724

Summary

The Alphaproteobacteria show a remarkable diversity of cell cycle-dependent developmental patterns, which are governed by the conserved CtrA pathway. Its central component CtrA is a DNA-binding response regulator that is controlled by a complex two-component signaling network, mediating distinct transcriptional programs in the two offspring. The CtrA pathway has been studied intensively and was shown to consist of an upstream part that reads out the developmental state of the cell and a downstream part that integrates the upstream signals and mediates CtrA phosphorylation. However, the role of this circuitry in bacterial diversification remains incompletely understood. We have therefore investigated CtrA regulation in the morphologically complex stalked budding alphaproteobacterium Hyphomonas neptunium. Compared to relatives dividing by binary fission, H. neptunium shows distinct changes in the role and regulation of various pathway components. Most notably, the response regulator DivK, which normally links the upstream and downstream parts of the CtrA pathway, is dispensable, while downstream components such as the pseudokinase DivL, the histidine kinase CckA, the phosphotransferase ChpT and CtrA are essential. Moreover, CckA is compartmentalized to the nascent bud without forming distinct polar complexes and CtrA is not regulated at the level of protein abundance. We show that the downstream pathway controls critical functions such as replication initiation, cell division and motility. Quantification of the signal flow through different nodes of the regulatory cascade revealed that the CtrA pathway is a leaky pipeline and must involve thus-far unidentified factors. Collectively, the quantitative system-level analysis of CtrA regulation in H. neptunium points to a considerable evolutionary plasticity of cell cycle regulation in alphaproteobacteria and leads to hypotheses that may also hold in well-established model organisms such as Caulobacter crescentus.

Keywords:

Cell cycle and cell division – Gene regulation – Phosphorylation – Regulator genes – Cell polarity – Regulons – Neptunium – Caulobacter crescentus

Introduction

The ability to initiate processes such as DNA replication at the right moment in the cell cycle is crucial for the fitness of all cells. To ensure the correct timing of events, cells need to sense multiple input signals and integrate these signals in order to determine whether or not to start a certain process. A common strategy in both eukaryotes and prokaryotes is to use protein phosphorylation as a switching mechanism in this cellular information switchboard [13]. The integration of different signals often involves complex phosphorylation cascades, in which the reception of a signal leads to the phosphorylation of proteins, which then in turn phosphorylate downstream targets. These systems are called phosphorelays, or two component systems (TCSs) when the context is limited to a sensory histidine kinase and a downstream response regulator [4, 5]. Phosphorelays and TCSs play an important role in the regulation of cellular development of bacteria, controlling processes such as sporulation [6, 7] and cell cycle regulation [8, 9].

Many bacteria grow and divide without major changes in their overall morphology. However, there are also various species, including members of the Alphaproteobacteria, that undergo dramatic morphological changes when they proceed through their (asymmetric) cell cycle [10, 11]. In the alphaproteobacterial model organism Caulobacter crescentus, the two daughter cells have different shapes and fates: the stalked cell is sessile and can produce offspring, whereas the swarmer cell is motile and non-reproductive [9, 12]. This asymmetry is mainly dictated by the differential activation of the DNA-binding response regulator CtrA, whose phosphorylated form (CtrA~P) specifically accumulates in the swarmer but not in the stalked sibling [13, 14]. CtrA~P regulates, both positively and negatively, global gene expression by interacting with ~100 different promoter regions [15]. Apart from that, it binds to sites at the chromosomal origin of replication, thereby preventing the transition of cells into S phase [16]. In concert with other global regulators, such as the replication initiator/transcription factor DnaA [1719] and the σ70 cofactor GcrA [20, 21], CtrA thus ensures the ordered progression of cells through their developmental program, licensing processes such as motility and cell division at the right moment in the cell cycle.

CtrA itself is regulated at the level of activity and abundance, and both of these parameters are affected by phosphorylation. The activation of CtrA proceeds via a phosphorelay including the histidine kinase CckA and the phosphotransferase ChpT [22, 23]. The different fates of the daughter cells result from the differential control of CckA activity at the two cell poles by the pseudokinase DivL and the single-domain response regulator DivK. At the flagellated pole, DivL interacts with CckA and stimulates its kinase activity, thereby mediating CtrA phosphorylation. At the stalked pole, by contrast, DivL is sequestered by DivK~P, which facilitates the transition of CckA to the phosphatase mode and the dephosphorylation of CtrA by reversion of the CckA-ChpT pathway [24, 25] (Fig 1). At the early predivisional stage, this cellular asymmetry is further enhanced by the preferential accumulation of CckA at the new cell pole [26, 27].

Model of the cell-cycle dependent activation and stabilization of CtrA in <i>C</i>. <i>crescentus</i>.
Fig. 1. Model of the cell-cycle dependent activation and stabilization of CtrA in C. crescentus.
(A) At the flagellated pole, PleC dephosphorylates PleD and DivK. DivL activates the kinase activity of CckA, which leads to the phosphorylation of CtrA via ChpT and, thus, to the activation of CtrA. ChpT also phosphorylates and thereby inactivates CpdR. (B) DivJ localizes to the stalked pole where it phosphorylates PleD and DivK. DivK~P sequesters DivL and thereby prevents DivL from activating the kinase activity of CckA. PleD~P synthesizes cyclic-di-GMP (cdG), which activates the phosphatase activity of CckA. As consequence, CckA dephosphorylates CtrA and CpdR. CpdR, together with PopA (activated by cdG made by PleD~P) and RcdA, stimulates the degradation of CtrA by ClpXP. The active forms of some of the proteins are marked with an asterisk (*).

Phosphotransfer also plays an important role in the upstream pathway, which modulates the activity state of DivK. At the stalked pole, the localization factor SpmX [28, 29] recruits the histidine kinase DivJ [30], which in turn phosphorylates DivK and thereby enables the transition of CckA to the phosphatase mode [24, 25]. In parallel, DivJ phosphorylates the guanylate cyclase PleD, which then synthesizes the second messenger cyclic-di-GMP (c-di-GMP) [31] and, in this way, further stimulates the phosphatase activity of CckA [32]. At the flagellated pole, by contrast, the localization factor PodJ recruits the histidine kinase/phosphatase PleC [29, 33, 34], which dephosphorylates both DivK and PleD, thus triggering transition of CckA to the kinase state [24, 25]. Apart from that, PodJ is also involved in the positioning of DivK, DivL and CckA to the flagellated pole in predivisional cells [12].

In addition to its inactivation by dephosphorylation, C. crescentus CtrA is also subject to targeted proteolysis by the ClpXP protease complex [35, 36]. This process is activated by two convergent pathways that are both initiated by the histidine kinase DivJ. On the one hand, the reversal of the CckA-ChpT pathway induced by the phosphorylation of DivK not only dephosphorylates CtrA but also the adapter protein CpdR, which then primes ClpXP for CtrA degradation [37, 38] (Fig 1). On the other hand, c-di-GMP formed upon phosphorylation of PleD binds and activates a second adapter protein, PopA, which then recruits CtrA to the primed ClpXP protease, triggering its degradation [39, 40].

Apart from C. crescentus, the Alphaproteobacteria include various other species with intriguing cell biological features [11]. The extensive body of knowledge on CtrA-dependent cell cycle regulation in C. crescentus has therefore sparked investigations into the conservation of this central pathway, with the aim to understand how its architecture evolved to adapt to different growth modes. These studies have shown that the CtrA pathway is essential in the polarly growing Rhizobiales [4144], but dispensable in related species that divide by symmetric binary fission, such as Rhodobacter capsulatus and Sphingomonas melonis [4548]. However, the CtrA pathway is still uninvestigated in the stalked budding alphaproteobacteria, a group of organisms that also includes the newly established model organism Hyphomonas neptunium [49, 50]. H. neptunium is closely related to C. crescentus [51] and shows a biphasic life cycle in which a motile swarmer cells sheds its flagellum, establishes a stalk and finally produces a daughter cell at the end of this stalk by expansion of the terminal stalk segment (stalk-terminal budding) [10, 52, 53] (Fig 2). It is currently unknown how the cell cycle regulatory circuitry has changed during evolution to give rise to this unusual mode of proliferation. Clarification of this issue may identify conserved features that have so far been neglected in well-studied model organisms such as C. crescentus, because they have become redundant in these species. Additionally, it may shed light on the plasticity of cell cycle regulatory networks and its role in the striking morphological diversification that alphaproteobacteria underwent during their adaptation to distinct environmental niches.

Biphasic life cycle of <i>H</i>. <i>neptunium</i>.
Fig. 2. Biphasic life cycle of H. neptunium.
Shown are a timelapse series (DIC) and a cartoon of a single H. neptunium cell that progresses through its cell cycle. The swarmer cell sheds its flagellum and differentiates into a stalked cell, which then produces a bud at the end of its stalk, thereby ultimately giving rise to a new swarmer cell. The stalked mother cell elongates its stalk and then immediately starts the next division cycle, whereas the swarmer cell first needs to differentiate into a stalked cell to proliferate. Scale bar: 3 μm.

In this study, we show that the entire cell cycle regulatory pathway of C. crescentus, starting from the polarity-determining factors DivJ and PleC down to CtrA, is conserved in H. neptunium. Its downstream part, consisting of DivL, CckA, ChpT and CtrA, is essential and controls various critical functions, including DNA replication, cell division and flagellar synthesis. Strikingly, the upstream part of this pathway appears largely redundant, as DivK has no apparent effect on cell cycle progression and its interactors DivJ and PleC only make minor contributions. A quantification of the information flow through the entire CtrA pathway, using a novel approach that compares the regulons of the different nodes, suggests the existence of multiple so-far unidentified factors. These may include additional histidine kinases that are responsible for almost half of the input into the CtrA pathway as well as a response regulator that acts in parallel to DivK and that feeds signals from DivJ and PleC into a later part of the CtrA pathway. Localization studies reveal that CckA is distributed throughout the nascent daughter cell instead of forming a polar focus, which could have implications for the mechanism that CtrA uses to ensure asymmetry in the fate of the mother and daughter cell. Finally, we show that CtrA is not regulated at the level of protein abundance in H. neptunium. These results provide insight in the differential evolution of the cell cycle regulatory circuitry in C. crescentus, which divides by asymmetric binary fission, and H. neptunium, which divides by stalk-terminal budding. Furthermore, they set the stage for the identification of new factors that play a major role the regulation of the H. neptunium cell cycle and may also contribute to cell cycle progression and cellular differentiation in other alphaproteobacteria.

Results

The upper part of the CtrA pathway is functionally redundant, while the lower part is essential

The H. neptunium genome encodes homologs of all proteins directly involved in the CtrA pathway of C. crescentus (Table 1). DivJ, PleC, DivK, DivL and CckA from H. neptunium could (partially) complement temperature-sensitive mutations in the corresponding factors in C. crescentus (S1A Fig), verifying that they are indeed bona fide homologs. Induction of CtrA from H. neptunium in either wild-type or ctrAts strains of C. crescentus led to cell elongation and eventually to lysis, even at the permissive temperature (S1B–S1D Fig), suggesting that the CtrA homologs of these two organisms have functionally diverged, or respond differently to regulation. To test the importance of DivJ, PleC, DivK, PleD, CckA, ChpT and CtrA in H. neptunium, we set out to generate in-frame deletions in the respective genes. Notably, mutants were obtained for all genes belonging to the upstream part of the pathway (divJ, pleC, divK and pleD). An analysis of these strains for potential cell cycle-related phenotypes showed that most of them exhibited wild-type morphologies and grow rates (Fig 3A and 3B). The ΔdivJ mutant, by contrast, grew slightly more slowly, and the majority of the stalked cells displayed an aberrant morphology, as reflected by elongated stalks and/or swollen cell bodies (Fig 3A and 3B). Apart from that, the ΔdivJ and ΔpleC strains had clear motility defects, whereas the ΔdivK and ΔpleD mutants still showed wild-type swimming behavior (Fig 3C). The motility defects of the ΔdivJ and ΔpleC strains and the growth and morphology defects of the ΔdivJ strain could be complemented by expression of the respective genes from an inducible promoter (Fig 3C and S2 Fig). Collectively, these results show that the DivJ/PleC-DivK module only has a minor effect on cell cycle regulation in H. neptunium. This finding is in stark contrast to the situation in C. crescentus, where DivK is essential [54] and DivJ [29, 55] as well as PleC [29, 56] play prominent roles in the control of cellular asymmetry and development.

DivJ, PleC, DivK and PleD have minor effects on <i>H</i>. <i>neptunium</i> morphology and motility.
Fig. 3. DivJ, PleC, DivK and PleD have minor effects on H. neptunium morphology and motility.
(A) DIC images of the H. neptunium wild type and mutants lacking DivJ (OL94), PleC (OL20), DivK (OL95) and PleD (OL21). The scaling of all images is identical. Scale bar: 3 μm. (B) Quantification of the morphological defects and the generation times of the strains described in (A). The number of cells analyzed to determine the percentage of aberrant stalk cells is 454 (wt), 438 (ΔdivJ), 430 (ΔpleC), 437 (ΔdivK), and 439 (ΔpleD). The generation times given represent the mean values from at least four replicate growth curves. (C) Motility of H. neptunium cells lacking DivJ, PleC, DivK or PleD. A mutant lacking the flagellar protein FliL (HW2) was used as a negative control. Strains complemented for DivJ (OL123) and PleC (OL23) were analyzed to verify that the motility defects of the mutant strains are caused by the lack of DivJ or PleC, respectively. To quantify motility, the indicated strains were spotted on soft agar and incubated for 6 days prior to imaging and quantification of the area of growth. In case of the DivJ and PleC complementation strains, inducer (300 μM CuSO4) was added to the medium. The swarm areas given are normalized to the value obained for the wild-type strain and represent the mean of at least fourteen replicates. Asterisks indicate statistically significant differences between the wild type and the respective mutant strain (one-sided Mann Whitney U-test): * p-value <0.05, ** p-value <0.01 and *** p-value <0.001.
Tab. 1. H. neptunium homologs of cell cycle-related genes from C. crescentus.
<i>H</i>. <i>neptunium</i> homologs of cell cycle-related genes from <i>C</i>. <i>crescentus</i>.
The table summarizes the query coverage, e-value and degree of identity for the best hits in H. neptunium obtained in BLAST searches with the respective proteins from C. crescentus as a query. The e-values of the respective second-best hits are given for comparison. The information in the column ‘Best reciprocal hit?’ indicates whether the indicated C. crescentus protein is the best hit in a BLAST search performed with the respective H. neptunium homolog as a query. NA indicates that no further hits were detected.

Unlike for the upstream part of the CtrA pathway, it was impossible to obtain H. neptunium deletion mutants lacking components of the downstream pathway, including DivL, CckA, ChpT and CtrA. To verify the essentiality of these proteins, we aimed to express their genes under the control of an inducible promoter and then study the effects of their depletion. This approach failed in the case of CtrA, possibly due to the high expression level of ctrA and the limited strength of the inducible promoters established for H. neptunium [49]. However, we did succeed in constructing conditional divL, cckA, and chpT mutants. Under restrictive conditions, all three strains showed a highly aberrant, heterogeneous morphology, with a high incidence of severely enlarged cell bodies and elongated and/or ectopic stalks (Fig 4A). At the same time, they showed a considerable reduction in the growth rate, although the effect was less pronounced for DivL-depleted cells, possibly due to leaky expression of the divL gene (Fig 4B). Flow cytometric analysis revealed that these defects correlated with a considerable increase in the DNA content, with many cells accumulating far more than two chromosome equivalents (Fig 4C). Interestingly, despite their aberrant shapes, the mutant cells were still able to segregate DNA to the nascent daughter cell compartments (S3 Fig), suggesting that their elevated DNA content is caused by overreplication of the chromosome in the absence of cell division. Collectively, these results demonstrate that the downstream part of the CtrA pathway, including DivL, CckA, ChpT as well as CtrA, has an essential role in H. neptunium cell cycle regulation, whereas the upstream part is, in large part, dispensable.

The essential proteins DivL, CckA and ChpT affect cell morphology and chromosome content.
Fig. 4. The essential proteins DivL, CckA and ChpT affect cell morphology and chromosome content.
(A) DIC images of the H. neptunium wild type and conditional mutants depleted of DivL (OL177), CckA (OL161) and ChpT (OL152). Protein depletion was achieved by cultivation of the strains for at least 24 h in the absence of inducer (300 μM CuSO4). The scaling of all images is identical. Scale bar: 5 μm. (B) Growth of conditional divL (OL177), cckA (OL161) and chpT (OL152) mutants under inducing (300 μM CuSO4) and non-inducing conditions. The curves shown represent the mean of three independent experiments. (C) Flow cytometric analysis showing the DNA content of the indicated conditional mutants under inducing and non-inducing conditions (n = 30,000 cells per conditions). Data obtained for the wild-type strain are shown for comparison. The fluorescence intensities corresponding to one (1n) or two (2n) chromosome equivalents are indicated.

DivJ and PleC mark opposite poles, whereas CckA dynamically localizes to entire cell compartments

To further study the functional conservation of the CtrA pathway in H. neptunium, we decided to determine the localization patterns of DivJ, PleC and CckA. To this end, the corresponding endogenous genes were replaced with alleles encoding fluorescent protein fusions. Time-lapse analysis revealed that DivJ is absent in most swarmer cells but then localizes to the old pole of early stalked cells, retaining this position until the end of the division cycle (Fig 5A and 5E). PleC, by contrast, is initially positioned at the old pole of the swarmer cell and then relocates to the tip of the growing stalk, so that it finally ends up at flagellated pole of the bud compartment (Fig 5B and 5E). Similar to their C. crescentus homologs, the two proteins thus mark opposite ends of the predivisional cell. Further analyses showed that DivJ and PleC frequently colocalize with the H. neptunium SpmX and PodJ homologs, respectively (Fig 5C–5E), and depend on their corresponding partner protein for proper localization (S4 Fig). Interestingly, in the absence of PodJ, most cells not only lacked a PleC focus but also became misshapen once they entered the stalked stage (S4 Fig and S5 Fig). As this effect was not observed for a ΔpleC mutant, PodJ may have important additional functions that go beyond its interaction with PleC. Unlike in the case of DivJ and PleC, the localization pattern of CckA clearly differed between H. neptunium and C. crescentus. The H. neptunium homolog, a membrane protein containing two predicted transmembrane domains (like its homolog in C. crescentus), showed an even distribution in swarmer cells, assumed an irregular, patchy pattern in stalked cells, and finally accumulated in the nascent bud without forming a noticeable polar focus (Fig 5F and 5G). Importantly, cells producing fluorescently tagged CckA do not show any morphological or growth defect (Fig 5 and S6 Fig). Thus, CckA appears to exert its function without accumulating at the pole in H. neptunium. Together, our results suggest that DivJ and PleC are involved in establishing cell polarity in H. neptunium, whereas CckA aids in the asymmetric activation of CtrA.

DivJ and PleC mark opposite poles and CckA accumulates in the nascent bud.
Fig. 5. DivJ and PleC mark opposite poles and CckA accumulates in the nascent bud.
(A) and (B) Timelapse experiments following H. neptunium cells producing (A) DivJ-Venus (OL146) or (B) PleC-eYFP (OL151) from their respective native promotors. (C) Colocalization of DivJ-Venus (expressed from its native promotor) and SpmX-mCherry (induced by addition of 300 μM CuSO4) (OL32). (D) Colocalization of PleC-eYFP (expressed from its native promotor) and PodJ-mCherry (induced by addition of 300 μM CuSO4) (OL33). (E) Quantification of the number of fluorescent foci in cells imaged as described in panels (C) and (D). For each strain, 100 cells were analyzed per developmental stage. Note that the intensity of the PodJ-mCherry signal was close to the detection limit. Therefore, the counts given for this fusion protein may be an underestimation, explaining the discrepancy in the number of PodJ-mCherry and PleC-eYFP foci. (F) Timelapse experiment following a representative H. neptunium cell expressing CckA-Venus from its native promotor (OL2). (G) Kymograph of the fluorescence signal in the cell depicted in (F). Shown is the maximum signal on a line connecting the two poles and spanning the entire width of the cell. Scale bars: 2 μm.

The connectivity of the components constituting the CtrA pathway is conserved in H. neptunium

In C. crescentus, the localization of DivJ and PleC to opposite cell poles leads to the differential phosphorylation of their common target DivK and, thus, to the asymmetric activation of DivL and the downstream CckA-ChpT-CtrA phosphorelay in the two daughter cells [22, 2325, 29, 30]. To clarify whether these interactions are conserved in H. neptunium, we aimed to analyze the connectivity of the different proteins using in vivo and in vitro approaches. In doing so, we first focused on the upstream part of the CtrA pathway (Fig 6A). Bacterial two-hybrid analysis confirmed that both DivJ and PleC interact with DivK and that DivK interacts with DivL (Fig 6B). Furthermore, PleC and DivL showed a strong self-interaction, as has been suggested for PleC and shown for DivL in C. crescentus [57, 58]. Since both DivJ and PleC are also able to phosphorylate DivK in vitro (Fig 6C–6E), we conclude that the protein-protein interactions within the upstream part of the CtrA pathway are conserved. Next, we aimed to confirm the functionality of the CckA-ChpT-CtrA phosphorelay (Fig 7A). Since it was not possible to achieve autophosphorylation of H. neptunium CckA in vitro, we resorted to its functionally equivalent C. crescentus homolog (CckACC) (see S1A Fig) as a phosphoryl donor. Our in vitro assays show that ChpT can be phosphorylated by CckACC and subsequently transfer a phosphoryl group to CtrA (Fig 7B and 7C). Moreover, they confirm that the response regulator domain of H. neptunium CckA can phosphorylate ChpT after it has received a phosphoryl group from the autophosphorylated kinase domain of CckACC (Fig 7D and 7E and S7 Fig). These results suggest that the signaling cascade involved in the activation of CtrA is conserved in H. neptunium, although H. neptunium CckA may need a specific trigger to switch to its kinase state.

The connectivity within the DivJ/PleC-DivK module is conserved in <i>H</i>. <i>neptunium</i>.
Fig. 6. The connectivity within the DivJ/PleC-DivK module is conserved in H. neptunium.
(A) Scheme of the interactions within the DivJ/PleC-DivK-DivL module, as demonstrated in this study. Green arrows indicate phosphotransfer. The black arrow indicates a regulatory protein-protein interaction. (B) Bacterial two-hybrid assay testing for interactions between the proteins in the DivJ/PleC-DivK-DivL. The indicated proteins were fused to the T18 (black) or T25 (blue) subunit of adenylate cyclase from Bordetella pertussis, respectively. The fusions were produced in the reporter strain E. coli BTH101, and their interaction was analyzed by spotting of the cells on MacConkey agar. Interactions are indicated by a purple coloration of the colonies. The GCN4 leucin-zipper-region from yeast (Zip) was used as positive control, and empty vectors encoding only the T25 subunit as negative controls. (C) Autophosphorylation of DivJHN and PleCHN upon incubation with [γ32P]-ATP. Samples were taken at the indicated time points and analyzed by SDS-PAGE. Radioactivity was detected by phosphor imaging. (D) and (E) In vitro phosphotransfer from (D) DivJHN and (E) PleCHN to DivKHN. After autophosphorylation of DivJHN for 30 min and of PleCHN for 45 min, the indicated proteins (marked with pluses) were incubated for 90 sec. After addition of SDS sample buffer to stop the reactions, the mixtures were separated by SDS-PAGE, and radioactivity was detected by phosphor imaging.
The connectivity within the CckA-ChpT-CckA phosphorelay is conserved in <i>H</i>. <i>neptunium</i>.
Fig. 7. The connectivity within the CckA-ChpT-CckA phosphorelay is conserved in H. neptunium.
(A) Scheme of the CckA-ChpT-CtrA phosphorelay. Green arrows indicate phosphotransfer. (B) Autophosphorylation of CckACC upon incubation with [γ32P]-ATP. Samples were taken at the indicated time points after the start of the reaction and analyzed by SDS-PAGE prior to detection of radioactivity by phosphor imaging. (C) In vitro phosphotransfer from ChpTHN to CtrAHN with CckACC as a phosphoryl donor. After autophosphorylation of CckACC for 30 min, the indicated proteins (marked with pluses) were mixed and incubated for 5 min. After termination of the reactions by addition of SDS sample buffer, the mixtures were separated by SDS-PAGE and radioactivity was detected by phosphor imaging. (D) Autophosphorylation of CckA-KDCC upon incubation with [γ32P]-ATP. Samples were taken at the indicated time points and analyzed by SDS-PAGE prior to detection of radioactivity by phosphor imaging. (E) In vitro phosphotransfer from CckA-RRHN to ChpTHN with CckA-KDCC as a phosphoryl donor. After autophosphorylation of CckA-KDCC for 45 min, the indicated proteins (marked with pluses) were mixed and incubated for 5 min. The reaction was terminated by addition of SDS sample buffer, and the mixtures were separated by SDS-PAGE prior to detection of radioactivity by phosphor imaging.

Global analyses indicate a central role of CtrA in the regulation of cell cycle-related processes

Having shown the conservation of the pathway mediating CtrA activation, we set out to identify the regulatory targets of CtrA in H. neptunium. For this purpose, we depleted cells of CckA or ChpT to inhibit CtrA phosphorylation (and induce its partial depletion; see S8A Fig and S8B Fig) and then performed global transcriptomic analyses to identify genes that were differentially regulated in these conditions. The expression profiles obtained for CckA- and ChpT-depleted cells were highly similar (Fig 8A), verifying that both proteins act in the same pathway. In total, 381 genes were significantly (at least 2.5-fold; p<0.05) up- or down-regulated in the absence of CckA and/or ChpT (Fig 8A, S9A Fig and S1 Data). We define these 381 genes as the global CtrA regulon, although it cannot be excluded that some of them are controlled by other regulators than CtrA. To pinpoint the genes that are directly regulated by CtrA, we determined the chromosomal binding sites of CtrA by chromatin immunoprecipitation and subsequent high-throughput sequencing of isolated DNA fragments (ChIP-seq) (Fig 8B and S2S5 Data). Out of the 222 sites identified in this analysis, 211 were located in intergenic regions, potentially affecting the expression of 285 genes (S6 Data). A comparison of the sequences bound by CtrA identified the consensus motif TTAA-N7-TTAAC (Fig 8C). Very similar motifs have been reported for CtrA homologs from other species [14, 15, 41, 42, 46, 5961], which confirms the specificity of the ChIP-seq approach. Based on the ChIP-seq data, we then selected genes that carried a CtrA binding site in their coding or promoter region and showed differential regulation upon CckA/ChpT depletion (55 genes) or genes that were part of a putative operon that fulfilled these criteria (39 genes). The resulting 94 genes were defined as the direct CtrA regulon (Fig 8D and 8E, S9C Fig and S7 Data). Notably, several genes that are tightly bound by CtrA (S6 Data) are not part of the CtrA regulon. It is possible that the high-affinity binding sites in their promoters recruit non-phosphorylated CtrA or residual CtrA~P that is left after depletion of CckA or ChpT, leading to a lack of significant changes in expression under the experimental conditions used. Alternatively, since we analyzed mixed cultures, some CtrA-regulated genes may not be included in the global regulon because their mean expression levels (averaged over the cell cycle) were not significantly different from the levels obtained after CckA or ChpT depletion. Notably, the finding that many CtrA-bound genes are not differentially expressed upon inactivation of CtrA is not unique to H. neptunium, as a similar phenomenon has been previously observed for C. crescentus [15].

The direct CtrA regulon contains 94 genes.
Fig. 8. The direct CtrA regulon contains 94 genes.
(A) Transcriptional profiles of cells depleted of CckA and ChpT for 24 h. Shown are the log2-fold changes in the expression levels of all 381 genes that were differentially regulated upon disruption of the CckA-ChpT-CtrA phosphorelay (as compared to the wild-type strain). (B) Distribution of CtrA binding sites on the H. neptunium chromosome as identified by ChIP-seq analysis. The graph reveals a total of 222 CtrA binding sites, 211 of which are in intergenic regions. (C) CtrA consensus binding motif obtained by a comparison of the 80 chromosomal regions that were most enriched in the ChIP-seq data shown in (B). (D) Venn diagram showing the number of genes (single or in an operon) bound by CtrA (in blue) and the number of genes present in the entire CtrA regulon (in red). The intersection of these two gene sets defines the direct CtrA regulon and comprises 94 genes. (E) Direct CtrA regulon. Shown are the 94 genes contained in the direct CtrA regulon and the log2-fold changes in their expression compared to the wild type after depletion (24 h) of ChpT or CckA. Only genes with an RPKM value of >25, a p-value of <0.05 and a log2-fold difference of >2 were taken in account.

A functional categorization of the genes bound and/or regulated by CtrA revealed that many of them have no predicted function (37% of the direct regulon) or a putative role in 'cellular processes and signaling' (43% of the direct regulon) (S9 Fig). Specific functions that are extensively regulated by CtrA are flagellar motility (15 genes; 16% of the direct regulon) and pili-mediated adhesion (14 genes; 15% of the direct regulon) (S7 Data and S8 Data). Consistent with this result, a large majority of the genes and operons in the motility island of H. neptunium are preceded by a CtrA binding site (Fig 9). As in C. crescentus [62], pili genes are generally up-regulated by CtrA in H. neptunium. For flagellar genes, by contrast, the mode of regulation varies, suggesting a more complex regulatory mechanism. CtrA also appears to affect morphogenesis and cell division by directly controlling the expression of mreB and the cell division genes ftsAKQZ and ftsB (HNE_1978), respectively (Fig 8E). Moreover, it may influence cell cycle progression by modulating the levels of the universal second messenger c-di-GMP, as it regulates eight genes that are predicted to be involved in c-di-GMP synthesis or degradation in both a direct (HNE_0906, HNE_1423, HNE_1960, HNE_2067, HNE_2435, HNE_2558) and indirect (HNE_0279, HNE_3504) fashion (S7 Data and S8 Data). Interestingly, as in C. crescentus [21, 63], CtrA also interacts with the promoter region of its own gene (which additionally contains a binding site for the global regulator GcrA), suggesting the existence of an autoregulatory feedback loop (Fig 9B). Moreover, it controls the gene for the CtrA co-regulator SciP, which in C. crescentus represses the transcription of ctrA and other CtrA targets [64, 65]. CtrA may thus also regulate its own transcription in an indirect fashion (S7 Data and S8 Data). However, while affecting its accumulation at the transcriptional level, CtrA does not regulate genes for factors known to mediate CtrA proteolysis in C. crescentus, such as CpdR, RcdA, PopA and ClpXP (S7 Data). Notably, CtrA does not only regulate a functionally diverse set of genes but it also binds to a site next to the chromosomal origin of replication (as identified in [66]) (Fig 9C), suggesting that it could contribute to the control of DNA replication.

CtrA binds near the replication origin, in its own promotor region and in the motility island.
Fig. 9. CtrA binds near the replication origin, in its own promotor region and in the motility island.
(A) Distribution of CtrA binding sites in the motility island of H. neptunium (HNE_0239-HNE_0276). The graph shows the number of sequenced fragments per nucleotide in rpm (reads per million) for a ChIP-seq experiment performed with CtrA as a bait (blue). Sequencing results obtained for total input DNA (in grey) are presented as a negative control. The scheme on top provides an overview of the motility island. Genes predicted to be in the same operon are shown in the same color, orphan genes are indicated by white arrows. Genes or operons not preceded by a CtrA binding site are indicated by a red frame. (B) CtrA binding in the promoter region of the ctrAHN gene. Shown are the normalized number of sequenced fragments (rpm) obtained in ChIP-seq experiments with CtrA (blue) and GcrA (red). Sequencing results obtained for total input DNA (grey) are shown as a negative control. (C) CtrA binding at the chromosomal origin of replication (ori). Colors as in (A).

The DivJ/PleC-DivK module only has a minor effect on the activity of the CckA-ChpT-CtrA phosphorelay

Detailed knowledge of the CtrA targets does not only reveal the regulatory potential of this central cell cycle regulator in H. neptunium, but it also facilitates a quantitative assessment of the signal flow within the CtrA pathway. In particular, it offers the possibility to clarify the contribution of the non-essential DivJ/PleC-DivK module to the activity of the essential CckA-ChpT-CtrA phosphorelay. To this end, we determined the transcriptional profiles of the ΔdivJ, ΔpleC and ΔdivK mutants and compared them to those of the CckA and ChpT depletion strains (Fig 10A and S9 Data). In general, the three mutants showed considerably smaller changes in global gene expression than the two depletion strains, supporting the notion that the upstream part of the CtrA pathway has a smaller effect on the activity of CtrA than its downstream part. To enable a meaningful comparison of the data, we applied less stringent criteria to define the regulons of DivJ, PleC and DivK, using a p-value of <0.25 and a log2-fold difference of >0.5 as thresholds (instead of <0.05 and >1.3, respectively, as used for CckA and ChpT). These lower thresholds could potentially lead to the inclusion of some false positives in the respective regulons. Nevertheless, the DivJ, PleC and DivK regulons identified in this manner comprised only 61, 114 and 40 genes, respectively (S10S12 Data), and were thus considerably smaller than the CtrA regulon (381 genes). Surprisingly, the DivK regulon shows only little overlap with the DivJ (8%) or PleC (25%) regulons (Fig 10B), indicating that the roles of DivJ and PleC may go beyond the regulation of DivK activity.

DivJ, PleC and DivK only have a minor effect on the CckA-ChpT-CtrA pathway.
Fig. 10. DivJ, PleC and DivK only have a minor effect on the CckA-ChpT-CtrA pathway.
(A) Comparison of the transcriptional changes observed for genes in the direct CtrA regulon in the indicated depletion (-) or deletion (Δ) backgrounds. Shown are the log2-fold changes in transcript levels in the mutant strains as compared to the wild type. (B) Venn diagram depicting the overlap between the regulons of DivJ, PleC and DivK. The number of genes in the respective regulons is given in parentheses. Only genes with an RPKM value of >25, a p-value of <0.25 and a log2-fold difference of >0.5 were taken into account. (C) Proportion of genes in the regulons of DivJ, PleC and DivK that are regulated by CtrA. Genes present in the direct CtrA regulon are shown in yellow. Genes that are not included in the direct CtrA regulon but have a CtrA binding site or are in an operon and coregulated with a gene that has a CtrA binding site are shown in red. Genes not regulated by CtrA are shown in white. The same thresholds were applied as in (B) to define the DivJ, PleC and DivK regulons.

While many of the genes regulated by DivJ, PleC and DivK are of unknown function, some are predicted to be involved in known, cell cycle-regulated processes. Consistent with the motility defect of the corresponding mutants (Fig 3C), DivJ and PleC affect the expression of genes related to flagellar assembly and rotation. Moreover, the DivJ and DivK regulons contain genes involved in c-di-GMP metabolism, and all three regulons contain various pili-related genes (S10S12 Data). The major functions regulated by DivJ, PleC and DivK are thus similar to those regulated by CtrA. Nevertheless, there is only a partial overlap between the respective regulons (Fig 10C, S9 Data and S13 Data), with only 87% of the genes in the DivJ regulon and ~ 50% of the genes in the DivK and PleC regulons controlled and/or bound by CtrA (Fig 10A and S10S12 Data). These results suggest that the regulatory effects of DivJ, PleC and DivK are not only mediated through CtrA but, directly or indirectly, also through one or more additional transcription factors. In addition, they show that the CtrA pathway as described here is a leaky pipeline, in which only part of the signal is transmitted to the next protein of the CtrA signaling network, suggesting the existence of additional regulatory components (e.g. response regulators or other transcription factors).

CtrA is not regulated at the level of proteolysis

In C. crescentus and other species investigated, the activity of CtrA is tightly controlled in time and space to ensure proper cell cycle progression, based on a multi-layered regulatory network that acts at the levels of transcription, phosphorylation and protein degradation [9, 13, 63]. Having verified the transcriptional autoregulation (Fig 9B) and CckA-dependent phosphorylation (Fig 7C) of CtrA in H. neptunium, we aimed to determine if this central regulator shows cell cycle-dependent changes in protein abundance. For this purpose, we first tested whether our synchronization protocol for H. neptunium permits the detection of fluctuations in protein levels over the course of the cell cycle, using flagellin as a marker. Previous work has shown that the flagellum of H. neptunium is shed at the onset of stalk formation and then re-synthesized in the nascent bud compartment [50]. Consistent with this observation, the cellular levels of flagellin decreased markedly during the swarmer-to-stalked cell transition and then increased again towards the end of the cell cycle (Fig 11A and 11B), confirming the validity of our analytical approach. In contrast, the levels of CtrA did not show any appreciable cell cycle-dependent changes (Fig 11A and 11B) under the conditions used, suggesting that it may not be subject to extensive targeted proteolysis. To further test this hypothesis, we followed the amount of CtrA in the population after inhibiting protein translation (Fig 11E and 11F). We indeed observed only a moderate (12±4%) decrease in CtrA levels within a period of 7 h (2.5 generation times), which was very different from the rapid degradation of CtrACC in C. crescentus (t½ = 20 min, i.e. ~0.2 generation times) (S10 Fig). To further investigate the apparent lack of CtrA proteolysis, we next focused on the role of RcdA and CpdR, two adapter proteins required to target CtrA to the ClpXP protease in C. crescentus [3538] and other alphaproteobacteria [67, 68]. A mutant lacking CpdR, which is involved in the degradation of multiple ClpXP substrates [38], displayed mild morphological defects, including enlarged cell bodies and stalks. Inactivation of the CtrA-specific adapter RcdA, by contrast, did not have any obvious phenotypic consequences (Fig 11C), supporting the notion that CtrA degradation is not relevant for proper cell cycle progression in H. neptunium. Importantly, neither the ΔcpdR nor ΔrcdA mutant showed any increase in the average CtrA level (Fig 11D). Moreover, we did not observe any obvious change in the stability of CtrA in the two mutant backgrounds after the inhibition of protein synthesis (Fig 11E and 11F). Collectively, these results strongly suggest that unlike in C. crescentus [35, 36] and S. meliloti [67], the activity of CtrA is not regulated at the level of protein abundance in H. neptunium.

CtrA levels are stable throughout the cell cycle.
Fig. 11. CtrA levels are stable throughout the cell cycle.
(A) Immunoblot analysis showing the levels of CtrAHN over the course of the H. neptunium cell cycle. After synchronization, wild-type swarmer cells were transferred into pre-warmed medium and cultivated for 4 h. At the indicated time points, samples were taken and probed with an anti-CtrAHN antibody. To assess the synchrony of the cells, the same samples were additionally analyzed with an anti-flaggelin antibody. Shown are representative immunoblots (IB) and sections of the corresponding membranes stained with Amido black as loading controls (LC). The bands representing the proteins of interest are indicated by arrowheads. (B) Levels of CtrA and flagellin at different stages of the cell cycle. Samples were taken from a synchronized culture and analyzed by immunoblotting with anti-CtrAHN and anti-flagellin antibodies, as described in (A). Specific immunosignals were quantified and normalized to the total protein content of the respective lane, as detected by staining with Amido black. Values represent the mean of 7 (CtrAHN) or 8 (flagellin) replicates. Error bars indicate the standard deviation. (C) DIC images of exponentially growing H. neptunium ΔrcdA (OL44) and ΔcdpR (OL28) cells. Scale bar: 5 μm. (D) Comparison of the levels of CtrA in H. neptunium wild type (WT), ΔrcdA and ΔcdpR cells. Cells from an exponentially growing culture were probed with anti-CtrAHN antibody. Shown is a representative immunoblot (IB) and a band on the corresponding membrane stained with Amido black as a loading control (LC). (E) Immunoblot analysis showing the stability of CtrA in the wild-type (WT), ΔrcdA and ΔcdpR backgrounds. After growth of the cells to exponential phase, translation was inhibited by addition of chloramphenicol. Samples were taken at the indicated timepoints, adjusted to the same OD580 and probed with anti-CtrAHN antiserum. Shown are representative immunoblots (IB) and sections of the corresponding membranes stained with Amido black as a loading control (LC). (F) Levels of CtrA after translation inhibition in the wild-type (WT), ΔrcdA and ΔcdpR backgrounds. Shown are the mean intensities of the CtrA immunosignals obtained in three independent experiments of the type described in (E), corrected for cell growth after the addition of chloramphenicol. Error bars indicate the standard deviation.

Discussion

The members of most bacterial lineages divide by binary fission without prominent morphological or physiological differences between the daughter cells. Alphaproteobacteria, by contrast, have developed a variety of complex, bi- or multiphasic life cycles and thus offer the unique opportunity to study how a conserved cell cycle regulatory network has been rewired during the course of evolution to bring about fundamentally different outputs. Some of the most intricate life cycles within this order are observed in the genera Hyphomonas and Hyphomicrobium, which both divide by stalk-terminal budding. In this study, we established that the CtrA signaling cascade is essential for proper cell cycle regulation in H. neptunium and controls multiple central processes such as cell shape, motility, cell division and replication. Unlike in the closely related, but morphologically distinct species C. crescentus, the upstream part of the cascade (DivJ/PleC-DivK) has only a minor impact on CtrA activity. Moreover, there are characteristic differences in subcellular location of essential CtrA pathway components in H. neptunium, as CckA does not condense into polar foci but spreads throughout entire cell compartments. Finally, the activity of H. neptunium CtrA does not appear to be regulated at the level of protein abundance.

Interestingly, although the direct CtrA regulons of C. crescentus and H. neptunium cover similar functions, such as motility, cell division, and c-di-GMP metabolism, their precise composition differs considerably. Only slightly more than half of the genes in the direct regulon of CtrAHN have a homolog in C. crescentus, and only seven of these homologs are part of the direct regulon of CtrACC (S8 Data). This striking evolutionary plasticity in the target spectrum is in line with previous results showing that CtrA, for instance, affects cell division by controlling minCD in Sinorhizobium meliloti [67] but mipZ, ftsZ and, potentially, ftsQA in C. crescentus [15]. In H. neptunium, by contrast, CtrA regulates the cell division genes ftsAKQZ but none of the known Z-ring placement factors, such as the H. neptunium mipZ homolog (S8 Data).

The CtrA pathway also shows plasticity with respect to the importance of its components, in particular in its upstream part. In S. meliloti, PleC and DivK are essential, and DivJ cannot be deleted in the absence of the histidine kinase CbrA [69, 70]. In C. crescentus, by contrast, only DivK is essential [54], although mutants lacking DivJ or PleC show major cell polarity defects [29]. Our results now demonstrate that DivJ, PleC and DivK are all non-essential in H. neptunium. The observations that DivK is dispensable and its lack has no obvious phenotypic consequences is surprising. Since its interaction partner DivL is essential and critical for proper cell morphology and development, we propose the existence of one or multiple so-far unidentified factors that act in parallel to DivK to control DivL activity. The mild phenotypes of the ΔdivJ and, even more so, the ΔpleC mutant raise the question of how cell polarity is ultimately determined in H. neptunium. A factor that may be involved is the second messenger c-di-GMP. In C. crescentus, the levels of c-di-GMP vary in a cell cycle-dependent manner, mostly due to changes in the activity of the hybrid guanylate cyclase/response regulator PleD, thereby critically contributing to the differential regulation of CckA activity in the two daughter cells [31, 32]. However, we found that the inactivation of PleD does not have any obvious effect in H. neptunium. It thus remains to be determined if c-di-GMP in fact has a role in H. neptunium cell cycle regulation and, if so, what protein takes over the role of PleD in coupling cell polarity to CckA stimulation. Alternatively, additional histidine kinases, similar to CbrA in S. meliloti [70], could be involved in marking the two opposite cell poles to control cellular asymmetry.

Interestingly, a comparison of the transcriptional profiles obtained in this study showed that the DivJ/PleC-DivK module only provides part of the signal feeding into CtrA. To obtain an integrated and quantitative view of the CtrA pathway, we devised a novel way to characterize the flow of information in signaling pathways. To gain insight into the connectivity between the different nodes, we compared all regulons described in this study and determined the number of genes that are shared between pairs of regulons as a measure of the signal that is transferred between nodes. This approach only provides a partial picture of the signal flow through the different nodes, as it does not take into account the amplitude of the transmitted signal (i.e. the difference in gene expression between regulons). Moreover, there is some uncertainty in the connectivity values, because the regulon of DivL is still unknown and the number and nature of the genes contained in the different regulons depends on the thresholds used in the analysis. The results obtained can, nonetheless, provide unprecedented insight into the CtrA regulatory pathway and predict missing factors at specific positions in the signaling cascade. For instance, this quantitative analysis verified that there must be a thus-far unknown factor next to DivK that feeds more signal into the lower part of the CtrA pathway than DivK itself (S13 Data). In addition, there should be a minimum of three additional factors that together provide more than half of the input into the CtrA pathway (Fig 12, brown ovals). These missing components may not only account for the small sizes of the DivJ, PleC and DivK regulons but also explain the low amplitudes of the transcriptional changes within these regulons. Moreover, they may be the reason why a considerable number of genes bound by CtrA were not differentially regulated upon depletion of CckA or ChpT. It will be interesting to perform similar quantitative analyses in other species and test for the completeness of the established CtrA regulatory pathways. As a first attempt, we made use of published microarray data [15, 22, 71, 72] to investigate the signal flow in the CtrA regulatory pathway of C. crescentus (S14 Data). Previous work has shown that DivK is essential in C. crescentus and S. meliloti, suggesting that the putative factor acting in parallel to DivK in H. neptunium is either absent or functionally less important in these species. However, in line with our findings and our preliminary quantitative analysis (S14 Data), there may be additional histidine kinases acting in parallel to DivJ and PleC in C. crescentus. Indeed, previous work has shown that a C. crescentus mutant lacking both pleC and divJ [29] has a milder phenotype than cells depleted of DivK [54]. Moreover, DivK is still partially phosphorylated in the absence of functional DivJ and PleC [29], suggesting that DivKCC receives signals from at least one other, so-far unknown kinase. Further studies will be required to identify the predicted additional components of the CtrA regulatory pathway in H. neptunium and to determine whether these or similar factors are present in other alphaproteobacteria as well.

Model of the cell cycle regulatory cascade controlling CtrA activity in <i>H</i>. <i>neptunium</i>.
Fig. 12. Model of the cell cycle regulatory cascade controlling CtrA activity in H. neptunium.
DivJ, PleC and a thus-far unidentified factor (oval A; 1%) control the phosphorylation state of DivK. The role of DivK is subordinate to that of an additional regulator (oval B; ≥23%), which receives input from DivJ and PleC and potentially other proteins. DivK and the unidentified regulator transfer the majority of the signal to DivL, from where it is passed on via CckA and ChpT to CtrA. Additional unidentified factors feed into DivL and/or CckA (oval C; 44%) and ChpT (oval E; 10%), although some of the signal from DivK could also bypass DivL and CckA via an additional pathway (oval D; 2%). Finally, there may be a pathway that acts in parallel to ChpT to transfer signals from CckA to CtrA (oval F, 29%). The thickness of the lines is proportional to the fraction of the total signal reaching CtrA that is transferred via the corresponding connection (as indicated by the percentages above). Note that the sum of the input signals amounts to more than 100%, because some genes are present in the regulons of multiple input proteins. Essential proteins are marked with asterisks.

The results obtained in this and previous [49] work provide comprehensive insight into the spatiotemporal organization of the CtrA regulatory pathway in H. neptunium. Similar to the situation in C. crescentus, DivJ and PleC localize to opposite poles once the cell has left the swarmer stage (Fig 13). PleD, by contrast, differs from its C. crescentus homolog [73] by localizing to the new pole of the nascent daughter cell [49], and the amount of PleD that accumulates at the pole, as opposed to the diffuse population in the cytoplasm, appears to be higher in H. neptunium. The significance of this difference in PleD localization remains unclear, as the deletion of pleDHN does not produce any noticeable phenotype, while PleDCC function is important for motility and stalk formation in C. crescentus [74, 75]. CckA generally localizes to similar subcellular regions in H. neptunium and C. crescentus, as it is initially distributed within the swarmer cell envelope and later condenses in the nascent swarmer cell [26, 27]. However, in H. neptunium, CckA does not form distinct polar foci but rather spreads throughout the respective compartments. It remains to be determined how CckA can be specifically enriched in the nascent bud without being tethered to the cell pole, especially since PodJ [12], which in C. crescentus helps to recruit CckA to the nascent swarmer pole via MopJ [76] and possibly DivL [25], is still polarly localized in H. neptunium. The involvement of PodJ in the positioning of multiple factors may explain why cells lacking PodJ have a more severe phenotype than those lacking PleC (compare S5 Fig to Fig 3). Notably, the role of PodJHN seems to go beyond that of its C. crescentus homolog, because a C. crescentus ΔpodJ mutant has only a comparatively mild cell cycle defect [12].

Localization patterns of cell cycle regulatory proteins in <i>H</i>. <i>neptunium</i> and <i>C</i>. <i>crescentus</i>.
Fig. 13. Localization patterns of cell cycle regulatory proteins in H. neptunium and C. crescentus.
(A) Localization patterns in H. neptunium. Similar to its C. crescentus homolog, the H. neptunium histidine kinase/phosphatase PleC localizes to the flagellated pole of the swarmer cell and then relocates to the new pole once stalk formation initiates. Concomitant with the relocation of PleC, the histidine kinase DivJ localizes to the old (previously flagellated) pole, where it remains until cell division takes place. The predicted hybrid guanylate cycle/response regulator PleD first colocalizes with PleC at the flagellated pole and then relocates to the nascent bud to colocalize with DivJ. The histidine kinase CckA, by contrast, is initially dispersed throughout the swarmer cell membrane. As the cell cycle progresses, it condenses in the bud compartment but, unlike its C. crescentus homolog, never forms a distinct polar focus. (B) Localization patterns in C. crescentus. PleC localizes to the flagellated pole of the swarmer cell, acting as a phosphatase (P) that dephosphorylates DivK. Once the cell sheds the flagellum and starts to progress through its cell cycle, it switches from the phosphatase to the kinase (K) mode and relocates to the newly flagellated pole. DivJ, by contrast, localizes to the stalked pole, where it cooperates with PleC to maintain DivK in the phosphorylated state. CckA forms clusters that are randomly distributed throughout the swarmer cell envelope. At later stages of the cell cycle, it localizes to the stalked and flagellated poles, where it acts as a phosphatase (P) and kinase (K), respectively.

Another interesting question is why H. neptunium cells can tolerate a more diffuse CckA localization. In C. crescentus, two hypotheses have been put forward to explain the condensation of CckA into a polar focus. First, CckA was shown to require a high enough local concentration to self-interact efficiently and become active as a kinase [26, 27, 77, 78]. In addition, its accumulation at the flagellated pole may help to establish a gradient of CtrA~P with a maximum close to the pole-proximal replication origin, thereby blocking replication initiation in the (nascent) swarmer cell [77, 79, 80]. The absence of a polar CckA focus in the bud compartment of H. neptunium may suggest that CckA activity is not regulated at the level of protein concentration in this species, unless its activity is triggered at considerably lower threshold levels. However, there may be other mechanisms to still establish a CtrA~P gradient in H. neptunium. For instance, the narrow lumen of the stalk [10] may be sufficient to slow down the diffusion of CtrA(~P) from the mother cell to the bud compartments. Another aspect that may contribute to gradient formation could be the specific and/or non-specific DNA-binding activity of CtrA(~P). A recent study in C. crescentus revealed that CtrA shows constrained diffusion, likely because of its interaction with the nucleoid [77]. Previous work showed that its affinity for specific DNA binding sites increases by two orders of magnitude upon phosphorylation [8183], and it is conceivable that phosphorylation also increases, at least to some extent, its affinity for non-specific DNA. Importantly, in H. neptunium, the presence of the stalk leads to an almost complete spatial separation of the two sister chromosomes at all stages of the cell cycle. Unlike in C. crescentus, where the two sister nucleoids form a continuum, the majority of CtrA~P molecules generated by CckA in the nascent bud may thus be trapped on the local nucleoid in the bud compartment. In contrast, dephosphorylated CtrA that is generated in the mother cell may be preferentially retained on the mother cell nucleoid, or equilibrate rapidly within the cell in case its DNA binding affinity is too low to affect its diffusional behavior. The enrichment of CtrA~P in the nascent bud may obviate the need to degrade CtrA in the mother cell to re-initiate the next cell cycle once cell division has occurred. Indeed, the observation that H. neptunium CtrA is not degraded in a cell cycle-dependent manner represents a striking difference between this species and other alphaproteobacteria that have been studied so far [22, 37, 67]. Its unusual morphology thus renders H. neptunium an excellent model system to develop and study new hypotheses on the mechanisms that underlie the control of cell differentiation by CtrA and to clarify how gradients of DNA-binding proteins such as CtrA(~P) could be established within bacterial cells.

Collectively, the comparison of H. neptunium and C. crescentus reveals significant differences in the CtrA pathway with respect to (i) the precise nature of the genes that it regulates, (ii) the importance of the individual proteins involved the pathway and, thus, the degree of redundancy at the nodes in the pathway and (iii) the localization pattern of the major polarity determinant CckA. These findings indicate a striking plasticity in the connectivity and output of the pathway among even closely related alphaproteobacterial species that likely results from their adaptation to distinct lifestyles and cell morphologies. The comprehensive and quantitative analysis of the CtrA regulatory network presented and the hypotheses developed in this work can inspire research in more well-studied model organisms and thereby contribute to a deeper understanding of the regulatory pathways mediating bacterial cell cycle regulation.

Material and methods

Growth conditions

The H. neptunium wild type LE670 (ATCC 15444) [50] and its derivatives were cultivated at 28°C in Difco Marine Broth 2216 (BD Biosciences), supplemented with kanamycin (100/200) or triclosan (0.25/0.25) when required (μg ml-1 in liquid/solid medium). The expression of genes placed under the control of the copper-responsive promoter PCu [49] was induced with 300 μM CuSO4. The C. crescentus wild type (NA1000) [84] and its derivatives were grown in peptone yeast extract (PYE) medium at 28°C. When appropriate, media were supplemented with kanamycin (30/50). Gene expression from the xylose-inducible Pxyl promoter [85] was induced with 0.3% xylose. E. coli strains were cultivated in LB at 37°C, supplemented with the following antibiotics where appropriate (μg ml-1 in liquid/solid medium): ampicillin (200/200), kanamycin (30/50), triclosan (10/20) and chloramphenicol (30/20). E. coli strain WM3064 was grown in the presence of 300 μM 2,6-diaminopimelic acid (DAP). To induce the expression of genes from the T7 promoter, media were supplemented with 0.5 mM isopropyl-β-D-thiogalactopyranoside (IPTG). All strains were cultivated aerobically while shaking at 210 rpm.

Construction of plasmids and strains

The strains, plasmids and oligonucleotides used in this study are listed in S1S6 Tables. All plasmids were verified by DNA sequencing. E. coli TOP10 (Invitrogen) and XL1-Blue (Agilent Technologies) were used for cloning purposes. The heterologous overproduction of proteins was achieved in Rosetta 2(DE3)pLysS (Merck), whereas bacterial two-hybrid experiments were performed in E. coli BTH101 (EuroMedex). H. neptunium cells were transformed by conjugation with E. coli WM3064 as described previously [49]. Genomic in-frame deletions in H. neptunium were generated by double-homologous recombination using the counter-selectable sacB marker [10, 86]. Briefly, H. neptunium cells were transformed with a non-replicating plasmid harbouring 600 bp long up- and downstream flanking regions and the first and last 18 bp of the genomic region to be deleted. Correct integration of the plasmid was confirmed by colony PCR and transformants were grown to stationary phase in non-selective media, after which they were plated (dilution 1:200) on MB agar plates supplemented with 3% sucrose. After 5–8 days of incubation at 28°C, single colonies were re-streaked in parallel on plates containing either sucrose or the antibiotic, whose corresponding antibiotic cassette was carried on the plasmid used for construction. Individual clones growing in the presence of sucrose were subsequently tested for successful deletions by colony PCR with oligonucleotides priming outside the flanking regions used for recombination. For the construction of conditional cckA and chpT mutants, a copy of the respective gene was placed under control of PCu on a non-replicating plasmid and integrated at the PCu locus as previously described [49]. The native copy of the gene was then deleted as described above while expression of the ectopic copy was induced. C. crescentus cells were transformed by electroporation as previously described [87].

Growth curves

Cells were grown to mid-exponential growth phase and diluted to an optical density at 580 nm (OD580) of 0.05. Measurements were performed in a volume of 1 ml in 24-well polystyrene microtiter plates (BD Bioscience), using an Epoch2 microplate reader (BioTek). Growth was monitored over a period of 30 h at 28°C or 30°C with double-orbital shaking, with data points acquired at 30 min intervals. To follow growth after inhibition of translation, media were supplemented with 5 μg/ml chloramphenicol. The division time of each strain was calculated by fitting at least four replicate growth curves to a suitable growth model [88].

Microscopy

Exponentially growing cells were immobilized on pads consisting of 1% agar and imaged with an Axio Observer.Z1 microscope (Zeiss, Germany) equipped with a Zeiss Plan-Apochromat 100x/1.46 Oil DIC objective (Zeiss, Germany). An X-Cite 120PC metal halide light source (EXFO, Canada) and ET-CFP, ET-YFP or ET-TexasRed filter cubes (Chroma, USA) were used for fluorescence detection. Images were acquired with a pco.edge sCMOS camera (PCO, Germany), recorded using VisiView (Visitron Systems, Germany) and processed using Metamorph 7.7.5 (Molecular Devices, USA) and Adobe Illustrator CS5 (Adobe Systems, USA). To stain chromosomal DNA, exponentially growing cells were mixed with 4′,6-diamidino-2-phenylindole (DAPI) at a final concentration of 4 mg/ml and then incubated for 20 min at 28°C in the dark prior to analysis. Phenotypes were assessed by manually determining the fraction of aberrantly shaped cells in a representative population of stalked cells. Cell shape was classified as aberrant if the stalk was unusually wide or reached more than twice the normal length, if a cell had more than one stalk or if the area of the mother cell or bud exceeded the normal value by more than 1.5-fold.

Motility assays

Exponential H. neptunium cultures were diluted to an OD580 of 0.2 and dropped onto soft-agar plates consisting of 0.25% agar in 30% MB medium. The area of growth covered by each strain was measured after 6 days of incubation at 28°C using ImageJ v1.47 [89] and compared to that obtained for the wild type and a non-motile ΔfliL mutant. The data shown represent the mean of at least fourteen replicates. A one-sided Mann-Whitney U-test was performed to test for significant differences in the motility of wild-type and mutant cells. Raw data are shown in S15 Data.

Flow cytometry

Cells from an exponential culture were mixed with 10 μM Vybrant® DyeCycle Orange Stain (Molecular Probes) and incubated for 45 min at 28°C in the dark. Samples were analysed with a LSRFortessa analyzer (BD Biosciences). Fluorescence was excited using an Argon-green laser (514 nm) at 100 mW in combination with a 542/27 bandpass filter. Cell size was assessed based on the forward scatter, using a laser (488 nm) at 100 mW. Each strain was analyzed at least in duplicate (n = 30,000 cells per run). Cytometry data was acquired using FACSDiva Software 8.0 (BD Biosciences) and analyzed using FlowJo 5.4+ (FlowJo LLC, USA).

Bacterial two-hybrid assays

The genes under study were cloned into the vectors pUT18(-C) and pK(N)T25, which carried genes encoding the T18 and T25 fragments of Bordetella pertussis adenylate cyclase (obtained as part of the BACTH kit from EuroMedex). After transfer of the resulting plasmids into reporter strain E. coli BTH101, transformants were plated on MacConkey agar (Carl Roth) containing 1% maltose. Cells from single colonies were inoculated into LB medium, grown to stationary phase, and dropped onto MacConkey agar containing 1% maltose. Interactions were indicated by a red to purple coloration of the colonies after 48 h of growth at 37°C. A strain producing fusions of the yeast GCN4 leucin zipper regions (Zip) to the T18 and T25 fragments were included as a positive control in all assays [90]. Furthermore, for each interaction to be analyzed, cells carrying the respective pUT18(-C) derivative and empty pK(N)T25 served as a negative control.

Protein purification

To purify His6-DivJHN, His6-PleCHN, His6-DivKHN, His6-CckAHN-RR, His6-ChpTHN and His6-CtrAHN, E. coli Rosetta 2(DE3)pLysS was transformed with plasmids pOL135 (DivJHN), pOL208 (PleCHN), pOL134 (DivKHN), pJR75 (CckAHN-RR), pOL207 (ChpTHN) or pOL145 (CtrAHN), respectively, and grown at 37°C (or at 18°C in the case of PleCHN) in LB medium supplemented with 0.5% glucose. When the culture had reached an OD580 of 0.5–0.7, expression of the respective protein was induced with 0.5 mM IPTG. Once the culture had reached an OD580 of 2.5, the cells were harvested by centrifugation for 10 min at 4,500 g, washed twice in buffer A (50 mM Na2HPO4, 300 mM NaCl, 10 mM imidazole, adjusted to pH 8.0 with HCl) and stored at -80°C until further use. To purify the proteins, the cells were thawed on ice, resuspended in buffer A containing 10 μg/ml DNase I and 100 μg/ml PMSF, and disrupted by two to four passages through a French press at 16,000 psi. After centrifugation of the lysate at 30,000 g for 30–60 min at 4°C to remove cell debris, the supernatant was applied to a 5 ml HisTrap HP column (GE Healthcare) equilibrated with buffer A. The column was washed with 5 column volumes (CV) of buffer A, and protein was eluted with a linear imidazole gradient obtained by mixing buffer A with buffer B (50 mM Na2HPO4, 300 mM NaCl, 500 mM imidazole, adjusted to pH 8.0 with HCl) at a flow rate of 1 ml/min using an ÄKTA purifier 10 FPLC system (GE Healthcare). After analysis by SDS-PAGE, elution fractions containing the protein of interest were pooled and subjected to a second purification step (except in the case of His6-CckA-RRHN). To this end, the pooled fractions were loaded onto a 120 ml Superdex 200 16/60 (GE-Healthcare, USA) size exclusion column equilibrated with buffer C (50 mM Na2HPO4, 300 mM NaCl, adjusted to pH 8.0 with HCl), and protein was eluted with 1 CV of buffer C at a flow rate of 0.5 ml/min. Fractions containing the desired protein at high concentrations and purity (as visualized by SDS-PAGE) were pooled and dialyzed against buffer D (50 mM Na2HPO4, 300 mM NaCl, 5% glycerol, adjusted to pH 8.0 with HCl). The protein solution was then aliquotted, snap-frozen in liquid nitrogen and stored at -80°C until further use.

His6-CckACC and His6-CckA-KDCC were purified as described previously [91]. E. coli Rosetta 2(DE3)pLysS containing plasmid pMSB3 (CckACC) or pMSB2 (CckA-KDCC) was treated as described above with the following changes. After harvest, the cells were resuspended in lysis buffer (20 mM Tris-HCl, 0.5 M NaCl, 10% glycerol, 20 mM imidazole, 0.1% Triton X-100, pH 7.9) containing 10 μg/ml DNase I and 100 μg/ml PMSF. The HisTrap HP column (GE Healthcare) was equilibrated with wash buffer (20 mM HEPES-KOH, 0.5 M NaCl, 10% glycerol, 20 mM imidazole, 0.1% Triton X-100, pH 8.0), and protein was eluted with a linear gradient obtained by mixing wash buffer and elution buffer (20 mM HEPES-KOH, 0.5 M NaCl, 10% glycerol, 250 mM imidazole, pH 8.0) at a flow rate of 2 ml/min. Fractions containing the desired proteins at high concentrations and purity were pooled, dialyzed against storage buffer (10 mM HEPES-KOH, 50 mM KCl, 10% glycerol, 0.1 mM EDTA, 1 mM dithiothreitol, pH 8.0), snap-frozen and stored at -80°C until further use.

To obtain HNE_0264 (Flagellin) for antibody generation, a His6-SUMO-tagged derivative of the protein was overproduced in E. coli Rosetta 2(DE3)pLysS transformed with plasmid pJR74 and purified as described for His6-CckACC and His6-CckA-KDCC with the following changes. The lysis buffer contained 0.3 M KCl instead of 0.5 M NaCl, and protein was eluted with 10 CV of a linear gradient of 30–250 mM imidazole at a flow rate of 1 ml/min. After the first purification step on a 5 ml HisTrap HP column, the His6-SUMO-tag was cleaved off in buffer containing 1 mM DTT by treatment with His6-Ulp1 protease for 2 h at 4°C. The protein solution was then applied again to a 5 ml HisTrap HP column as described above. Untagged HNE_0264 was recovered from the flow through, dialyzed against storage buffer (50 mM Tris-HCl, 150 mM KCl, 10% glycerol, pH 8.0), snap-frozen and stored at -80°C until further use.

Phosphotransfer profiling

Phosphotransfer was analyzed essentially as described previously [91]. To promote the autophosphorylation of His6-CckACC and His6-CckACC -KD, the proteins were incubated at 30°C in storage buffer (10 mM HEPES-KOH, 50 mM KCl, 10% glycerol, 0.1 mM EDTA, pH 8.0) containing 2 mM DTT, 5 mM MgCl2, 500 μM ATP and 5 μCi [γ32P]-ATP (∼3,000 Ci/mmol, Hartmann Analytic). The autophosphorylation of DivJ and PleC was achieved by incubation of the proteins at 28°C in buffer R1 (50 mM Na2HPO4, 300 mM NaCl, pH 8.0) or buffer R2 (50 mM Tris/HCl, 50 mM NaCl, pH 7.6) containing 5 mM MgCl2 and 5 μCi [γ32P]-ATP. The phosphotransfer reactions were started by mixing phosphorylated His6-PleCHN, His6-CckAHN, His6-CckAHN-RR (autophosphorylated for 45 min) or His6-DivJHN (autophosphorylated for 30 min) with the acceptor protein(s) to be tested, which were diluted to a concentration of 3–5 μM in the respective storage buffer supplemented with 5 mM MgCl2. The mixtures were incubated for 5 min at 30°C (for reactions including CckA, ChpT and/or CtrA) or for 90 sec at 28°C (for reactions including DivJ, PleC and/or DivK). After addition of the same volume of 2x sample buffer (500 mM Tris/HCl pH 6.8, 8% SDS, 40% glycerol, 400 mM β-mercaptoethanol, 0.01% bromophenol blue), the proteins were loaded, without boiling, onto an 8–16% TGX Precast Protein Gel (Bio-Rad) and separated for 50 min at a constant voltage of 150V at room temperature. After removal of unincorporated ATP in the dye front with a scalpel, the gel was transferred to a plastic bag and exposed to a phosphor screen overnight. The phosphor screen was scanned with a Storm 840 or 860 system (Molecular Dynamics) at a resolution of 50 dots/cm. Each phosphotransfer assay was repeated three times.

Chromatin immunoprecipitation coupled to deep sequencig (ChIP-seq)

A culture of exponentially growing H. neptunium wild type cells at an OD580 of 0.5 was treated with paraformaldehyde (1% final concentration) in 10 mM sodium phosphate buffer (pH 7.6) at RT for 10 min to achieve crosslinking. Subsequently, the cultures were incubated for an additional 30 min on ice and washed three times in phosphate-buffered saline (PBS, pH 7.4). The resulting cell pellets were frozen in liquid nitrogen and stored at -80°C. After resuspension in TES buffer (10 mM Tris-HCl pH 7.5, 1 mM EDTA, 100 mM NaCl) containing 10 mM of DTT, the cells were incubated in the presence of Ready-Lyse lysozyme solution (Epicentre, Madison, WI) for 10 minutes at 37°C according to the manufacturer's instructions. The lysates were sonicated (Bioruptor Pico) at 4°C using 15 bursts of 30 sec to shear DNA fragments to an average length of 0.3–0.5 kbp and cleared by centrifugation at 14,000 rpm for 2 min at 4°C. After adjustment of the volumes (relative to the protein concentration) to 1 ml using ChIP buffer (0.01% SDS, 1.1% Triton X-84 100, 1.2 mM EDTA, 16.7 mM Tris-HCl pH 8.1, 167 mM NaCl,) containing protease inhibitors (Roche), the lysates were pre-cleared with 80 μl of protein-A agarose (Roche) and 100 μg BSA. 2% of each pre-cleared lysate was kept as total input (negative control) samples. The rest of the pre-cleared lysates was incubated overnight at 4°C with polyclonal antibodies targeting CtrAHN or GcrACC [21] (1:1,000 dilution). The immuno-complexes were captured after incubation with Protein-A agarose (pre-saturated with BSA) during a 2 h incubation at 4°C and subsequently washed with low-salt washing buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-HCl pH 8.1, 150 mM NaCl), with high-salt washing buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-HCl pH 8.1, 500 mM NaCl), with LiCl washing buffer (0.25 M LiCl, 1% NP-40, 1% deoxycholate, 1 mM EDTA, 10 mM Tris-HCl pH 8.1) and finally twice with TE buffer (10 mM Tris-HCl pH 8.1, 1 mM EDTA). Immunocomplexes were eluted from the Protein-A agarose beads with two times 250 μl elution buffer (SDS 1%, 0.1 M NaHCO3, freshly prepared) and then, in parallel to the total input samples, incubated overnight with 300 mM NaCl at 65°C to reverse the crosslinks. The samples were then treated with 2 μg of Proteinase K for 2 h at 45°C in 40 mM EDTA and 40 mM Tris-HCl (pH 6.5). DNA was extracted using phenol:chloroform:isoamyl alcohol (25:24:1), ethanol-precipitated using 20 μg of glycogen as a carrier and resuspended in 50 μl of DNAse/RNAse-free water.

Immunoprecipitated chromatin was used to prepare sample libraries for deep-sequencing at Fasteris SA (Geneva, Switzerland). ChIP-seq libraries were prepared using the DNA Sample Prep Kit (Illumina) according to the manufacturer's instructions. Single-end runs (50 cycles) were performed on an Illumina Genome Analyzer IIx or HiSeq2000, yielding several million reads. The single-end sequence reads (stored as FastQ files) were mapped against the genome of Hyphomonas neptunium ATCC 15444 (CP000158) using Bowtie version 0.12.9 (-qS best parameters, http://bowtie-bio.sourceforge.net/). ChIP-seq read sequencing and alignment statistics are summarized in S2 Data. After conversion of the output into standard genomic position format files (BAM format; using Samtools; http://samtools.sourceforge.net), the data were imported into SeqMonk version 0.34.1 (Babraham Bioinformatics; http://www.bioinformatics.babraham.ac.uk/projects/seqmonk/) to build ChIP-seq normalized sequence read profiles. Briefly, the genome was subdivided into 50 bp probes, and for every probe, we calculated the number of reads per probe as a function of the total number of reads (per million, using the Read Count Quantitation option). Analyzed data as shown in Figs 7 and 8 are provided in S3 Data (50 bp resolution) and S4 Data (10 bp resolution). Using the web-based analysis platform Galaxy (https://usegalaxy.org), CtrA and GcrA ChIP-seq peaks were called with MACS2 [92] relative to the total input DNA samples. The q-value (false discovery rate, FDR) cut-off for called peaks was 0.05. Peaks were rank-ordered according to fold-enrichment (S5 Data), and peaks with a fold-enrichment values >2 were retained for further analysis. Consensus sequences common to the top 80 enriched CtrAHN-associated loci were identified by scanning peak sequences (+/- 75 bp relative to the coordinates of the peak summit) for conserved motifs using MEME (http://meme-suite.org/) [93]. Binding sites of CtrA close to the origin of replication were identified using the previously identified origin region [66]. To locate CtrA binding sites upstream of transcriptional units in the motility islands, operons were predicted with DOOR2 [94].

Expression profiling of mutants by RNA-Seq and analysis of regulons

To assess the global expression changes in the ΔdivJ, ΔpleC and ΔdivK mutants and the conditional cckA and chpT mutants as compared to wild-type cells, two biological replicates of each strain (including the wild type) were grown to mid-exponential growth phase. After dilution of each culture to an OD580 of 0.15–0.2, the cells were washed two times in MB medium, harvested by centrifugation, snap-frozen in liquid nitrogen and stored at -80°C. For the conditional mutants and a wild-type replicate, two aliquots of cells were harvested: one that was grown in the presence of CuSO4 for 24 h, and one that was depleted of the respective protein by growth in the absence of CuSO4 for 24 h. RNA extraction, cDNA library preparation, sequencing as well as the normalization of the expression data were performed by Vertis Biotechnologie AG (Germany). In order to identify genes that were differentially expressed in the ΔdivJ, ΔpleC or ΔdivK deletion mutants or the cckA or chpT conditional mutants, the log2-fold changes of the normalized expression values (RPKM) were compared between the wild-type and mutant strains. For the cckA and chpT conditional mutants, genes were considered as differentially expressed when they showed at least a 1.3 log2-fold change in their expression with a p-value of at least 0.05 in a paired T-test. For the ΔdivJ, ΔpleC and ΔdivK mutants, a 0.5 log2-fold change and a p-value of at least 0.25 were used as thresholds. Genes represented by only a few transcripts (RPKM values below 25) were excluded from further analyses. The evaluation as well as the statistical analysis of the expression data was performed in Excel 2016 (Microsoft). For visual representation of the relevant expression data, genes were grouped with Cluster 3.0 [95] using the City-block distance option and visualized in Java TreeView 1.1.6r4 [96].

Genes that were differentially expressed in the cckA and/or chpT conditional mutants and featured a CtrA binding site in their immediate upstream region, in the gene itself or in a gene in the same operon, as predicted by DOOR2 [94], were defined as the CtrA direct regulon. To quantify the information flow in the CtrA pathway, the overlap between the different regulons was determined. When a gene was differentially regulated such that it passed the log2-fold thresholds described above, it was counted as contributing to the information flow. In case of the cckA and chpT conditional mutants, genes located in an operon with a regulated gene were also counted as part of the regulons. As the ΔdivJ, ΔpleC and ΔdivK mutants showed only very moderate transcriptional changes, we also considered genes in the information flux analysis that were left out of the regulons of DivJ, PleC and DivK because of a too high p-value. The 89 genes constituting the CtrA direct regulon were normalized to 100% and from the number of overlapping genes, the percentage of the signal leading up to CtrA was calculated at every arrow. Missing histidine kinases were postulated based on a difference between the overlap between the respective regulon and the CtrA direct regulon and that of the regulon of the preceding node, as well as from evaluating the signal input and output in total and at every node. These two approaches gave highly similar results for all missing histidine kinases.

Synchronization of H. neptunium

Cells were synchronized by filtration as described previously [97]. In brief, H. neptunium was grown to late exponential phase (OD580 of 0.6), pelleted (3,000 g at 4°C), resupended in ice-cold PBS and filtered through 1.2 μm pore-sized nitrocellulose filters (Merck-Millipore). The flow-through was collected on ice, filtered again (pore size 0.8 μm, Merck-Millipore). The flow-through of the second filtration step, which was highly enriched in swarmer cells, was centrifuged and the resulting pellet was resuspended in pre-warmed media to an OD580 of 0.4. The culture was then incubated at 28°C and samples were withdrawn at the indicated intervals. The synchrony of the culture was assessed using microscopy.

Antibody generation and immunoblot analysis

Polyclonal antibodies against His6-CtrA and flagellin (purified as described above) were raised in rabbits by Eurogentec (Belgium) and Davids Antibodies (Germany), respectively. Immunoblot analysis was performed on cultures that had been harvested by centrifugation and resuspended in 2x SDS sample buffer such that each sample was normalized to the OD580 of the culture. Cells were lyzed by incubation at 95°C for 10 min and proteins were separated on 11–15% SDS-polyacrylamide gels and transferred to a polyvinylidene fluoride (PVDF) membrane (Millipore, Germany). Immunodetection was performed with polyclonal rabbit sera targeting CtrAHN, CtrACC [13], GFP (Sigma Aldrich) (all 1:10,000) and flagellin (1:2,000) according to standard procedures, using goat-anti-rabbit conjugated to horseradish peroxidase (PerkinElmer, USA) as a secondary antibody. Immunocomplexes were visualized using the Western Lightning plus-ECL chemiluminescent reagent (PerkinElmer, USA) in a ChemiDoc MP imaging system (Bio-Rad, USA). ImageLab 5.0 (Bio-Rad, USA) and Adobe Illustrator CS5 (Adobe Systems, USA) were used for acquisition, quantification and further image processing.

To visualize fluctuations in the levels of CtrA and flagellin over the course of the cell cycle, samples were taken from the same culture every 30 min from the onset of synchrony. Samples were loaded on separate gels for CtrA and flagellin immunodetection. Both experiments were repeated 7 or 8 times in total, with samples stemming from the same synchronization event. The amounts of CtrA and flagellin were normalized for t0 and all bands were normalized for the total amount of protein present in the entire lane (as measured by Amido black staining) in order to enable a comparison between the different timepoints. Raw data are shown in S15 Data.

Determination of CtrA stability in vivo

In order to probe the stability of CtrA in vivo, H. neptunium or C. crescentus cells were grown to mid-exponential growth phase and protein synthesis was blocked by addition of 5–10 μg/ml chloramphenicol for H. neptunium [49] and 100 μg/ml chloramphenicol for C. crescentus. At the indicated intervals, 1 or 2 ml samples were withdrawn for immunoblot analysis and loaded on a gel after normalization to the OD580 of the sample (as described above). CtrA was detected using the α-CtrA antibody (1:10000) raised in this study. Signals were quantified with the ImageLab 5.0 software (Bio-Rad, USA). For each strain, immunoblots of two clones from three independent experiments were analysed. As the focus was on the total amount of CtrA present in the culture, growth curves recorded in the presence of 5 μg/ml chloramphenicol (as described above) were used to correct for the dilution of CtrA due to cell growth after inhibition of translation. Raw data are shown in S15 Data.

Bioinformatics

All DNA sequences were retrieved from KEGG [98] and protein sequences were obtained from KEGG or UniProt [99]. The SMART database [100] was used to identify (transmembrane) domains in histidine kinases and response regulators. Homologs of proteins from C. crescentus were identified via BLAST [101]. The annotation of genes in the different regulons was performed manually, using the annotations in UniProt. COG categories were obtained via the Microbial Genomic context Viewer [102] and COG function classes using the table provided on the NCBI COG website (ftp://ftp.ncbi.nih.gov/pub/wolf/COGs/COG0303/cogs.csv).

Supporting information

S1 Fig [a]
Cross-complementation shows the (partial) conservation of function of cell cycle-relevant genes between . and . .

S2 Fig [ol123]
Expression of can complement the phenotype of an . Δ mutant

S3 Fig [ol177]
. cells still segregate chromosomal DNA after depletion of DivL, CckA or ChpT.

S4 Fig [ol36]
Polar localization of DivJ and PleC depends on SpmX and PodJ respectively.

S5 Fig [ol34]
Lack of , but not of , leads to morphological defects.

S6 Fig [a]
CckA-Venus supports normal growth and is stably expressed.

S7 Fig [tif]
CckA-KD can phosphorylate CtrA directly when CckA-RRHN is absent.

S8 Fig [a]
The CtrA level decreases upon depletion of CckA and ChpT.

S9 Fig [a]
CtrA regulation predominantly affects genes of unknown function as well as genes involved in cellular processes and signaling.

S10 Fig [a]
The half-life of CtrACC in . is considerably shorter than the generation time.

S1 Table [pdf]
. strains used in this study.

S2 Table [pdf]
. strains used in this study.

S3 Table [pdf]
. strains used in this study.

S4 Table [pdf]
General plasmids used in this study.

S5 Table [pdf]
Plasmids generated in this study.

S6 Table [pdf]
Oligonucleotides in this study.

S1 Data [xlsx]
Total CtrA regulon, comprising all genes that are differentially expressed upon depletion of CckA and/or ChpT.

S2 Data [xlsx]
Summary of the ChIP-seq sequencing and alignment statistics.

S3 Data [xlsx]
Distribution of the CtrA and GcrA ChIP-seq reads over the . chromsome (50 bp resolution).

S4 Data [xlsx]
Distribution of CtrA and GcrA ChIP-seq reads within specific regions of the . chromosome (10 bp resolution).

S5 Data [xlsx]
MACS2 peak detection analysis of the CtrA and GcrA ChIP-seq experiments.

S6 Data [xlsx]
CtrA binding sites on the . binding sites as determined by ChIP-seq analysis.

S7 Data [xlsx]
Direct CtrA regulon.

S8 Data [xlsx]
Comparison between the direct CtrA regulons of . and . .

S9 Data [xlsx]
Comparison of the genes in the direct CtrA regulon in different backgrounds.

S10 Data [xlsx]
DivJ regulon.

S11 Data [xlsx]
PleC regulon.

S12 Data [xlsx]
DivK regulon.

S13 Data [xlsx]
Quantification of overlap between the direct CtrA, ChpT, CckA, DivK, PleC and DivJ regulons enables a reconstruction of the signal flow through the CtrA pathway.

S14 Data [xlsx]
Analysis of the signal flow in the CtrA regulatory pathway of . .

S15 Data [xlsx]
Raw data.


Zdroje

1. Esser D, Hoffmann L, Pham TK, Brasen C, Qiu W, Wright PC, et al. Protein phosphorylation and its role in archaeal signal transduction. FEMS Microbiol Rev. 2016;40(5):625–647. doi: 10.1093/femsre/fuw020 27476079.

2. Kobir A, Shi L, Boskovic A, Grangeasse C, Franjevic D, Mijakovic I. Protein phosphorylation in bacterial signal transduction. Biochim Biophys Acta. 2011;1810(10):989–994. doi: 10.1016/j.bbagen.2011.01.006 21266190.

3. Kyriakis JM. In the beginning, there was protein phosphorylation. J Biol Chem. 2014;289(14):9460–9462. doi: 10.1074/jbc.R114.557926 24554697.

4. Hoch JA. Two-component and phosphorelay signal transduction. Curr Opin Microbiol. 2000;3(2):165–170. doi: 10.1016/s1369-5274(00)00070-9 10745001.

5. Stock AM, Robinson VL, Goudreau PN. Two-component signal transduction. Annu Rev Biochem. 2000;69:183–215. doi: 10.1146/annurev.biochem.69.1.183 10966457.

6. Higgins D, Dworkin J. Recent progress in Bacillus subtilis sporulation. FEMS Microbiol Rev. 2012;36(1):131–148. doi: 10.1111/j.1574-6976.2011.00310.x 22091839.

7. Sarwar Z, Garza AG. Two-component signal transduction systems that regulate the temporal and spatial expression of Myxococcus xanthus sporulation genes. J Bacteriol. 2016;198(3):377–385. doi: 10.1128/JB.00474-15 26369581.

8. Lasker K, Mann TH, Shapiro L. An intracellular compass spatially coordinates cell cycle modules in Caulobacter crescentus. Curr Opin Microbiol. 2016;33:131–139. doi: 10.1016/j.mib.2016.06.007 27517351.

9. Tsokos CG, Laub MT. Polarity and cell fate asymmetry in Caulobacter crescentus. Curr Opin Microbiol. 2012;15(6):744–750. doi: 10.1016/j.mib.2012.10.011 23146566.

10. Cserti E, Rosskopf S, Chang YW, Eisheuer S, Selter L, Shi J, et al. Dynamics of the peptidoglycan biosynthetic machinery in the stalked budding bacterium Hyphomonas neptunium. Mol Microbiol. 2017;103(5):875–895. doi: 10.1111/mmi.13593 27997718.

11. Randich AM, Brun YV. Molecular mechanisms for the evolution of bacterial morphologies and growth modes. Front Microbiol. 2015;6:580. doi: 10.3389/fmicb.2015.00580 26106381.

12. Curtis PD, Quardokus EM, Lawler ML, Guo X, Klein D, Chen JC, et al. The scaffolding and signalling functions of a localization factor impact polar development. Mol Microbiol. 2012;84(4):712–735. doi: 10.1111/j.1365-2958.2012.08055.x 22512778.

13. Domian IJ, Quon KC, Shapiro L. Cell type-specific phosphorylation and proteolysis of a transcriptional regulator controls the G1-to-S transition in a bacterial cell cycle. Cell. 1997;90(3):415–424. doi: 10.1016/s0092-8674(00)80502-4 9267022.

14. Quon KC, Marczynski GT, Shapiro L. Cell cycle control by an essential bacterial two-component signal transduction protein. Cell. 1996;84(1):83–93. doi: 10.1016/s0092-8674(00)80995-2 8548829.

15. Laub MT, Chen SL, Shapiro L, McAdams HH. Genes directly controlled by CtrA, a master regulator of the Caulobacter cell cycle. Proc Natl Acad Sci U S A. 2002;99(7):4632–4637. doi: 10.1073/pnas.062065699 11930012.

16. Quon KC, Yang B, Domian IJ, Shapiro L, Marczynski GT. Negative control of bacterial DNA replication by a cell cycle regulatory protein that binds at the chromosome origin. Proc Natl Acad Sci U S A. 1998;95(1):120–125. doi: 10.1073/pnas.95.1.120 9419339.

17. Collier J. Cell cycle control in Alphaproteobacteria. Curr Opin Microbiol. 2016;30:107–113. doi: 10.1016/j.mib.2016.01.010 26871482.

18. Collier J, Murray SR, Shapiro L. DnaA couples DNA replication and the expression of two cell cycle master regulators. EMBO J. 2006;25(2):346–356. doi: 10.1038/sj.emboj.7600927 16395331.

19. Hottes AK, Shapiro L, McAdams HH. DnaA coordinates replication initiation and cell cycle transcription in Caulobacter crescentus. Mol Microbiol. 2005;58(5):1340–1353. doi: 10.1111/j.1365-2958.2005.04912.x 16313620.

20. Haakonsen DL, Yuan AH, Laub MT. The bacterial cell cycle regulator GcrA is a σ70 cofactor that drives gene expression from a subset of methylated promoters. Genes Dev. 2015;29(21):2272–2286. doi: 10.1101/gad.270660.115 26545812.

21. Holtzendorff J, Hung D, Brende P, Reisenauer A, Viollier PH, McAdams HH, et al. Oscillating global regulators control the genetic circuit driving a bacterial cell cycle. Science. 2004;304(5673):983–987. doi: 10.1126/science.1095191 15087506.

22. Biondi EG, Reisinger SJ, Skerker JM, Arif M, Perchuk BS, Ryan KR, et al. Regulation of the bacterial cell cycle by an integrated genetic circuit. Nature. 2006;444(7121):899–904. doi: 10.1038/nature05321 17136100.

23. Jacobs C, Domian IJ, Maddock JR, Shapiro L. Cell cycle-dependent polar localization of an essential bacterial histidine kinase that controls DNA replication and cell division. Cell. 1999;97(1):111–120. doi: 10.1016/s0092-8674(00)80719-9 10199407.

24. Chen YE, Tsokos CG, Biondi EG, Perchuk BS, Laub MT. Dynamics of two phosphorelays controlling cell cycle progression in Caulobacter crescentus. J Bacteriol. 2009;191(24):7417–7429. doi: 10.1128/JB.00992-09 19783630.

25. Tsokos CG, Perchuk BS, Laub MT. A dynamic complex of signaling proteins uses polar localization to regulate cell-fate asymmetry in Caulobacter crescentus. Dev Cell. 2011;20(3):329–341. doi: 10.1016/j.devcel.2011.01.007 21397844.

26. Angelastro PS, Sliusarenko O, Jacobs-Wagner C. Polar localization of the CckA histidine kinase and cell cycle periodicity of the essential master regulator CtrA in Caulobacter crescentus. J Bacteriol. 2010;192(2):539–552. doi: 10.1128/JB.00985-09 19897656.

27. Iniesta AA, Hillson NJ, Shapiro L. Cell pole-specific activation of a critical bacterial cell cycle kinase. Proc Natl Acad Sci U S A. 2010;107(15):7012–7017. doi: 10.1073/pnas.1001767107 20351295.

28. Radhakrishnan SK, Thanbichler M, Viollier PH. The dynamic interplay between a cell fate determinant and a lysozyme homolog drives the asymmetric division cycle of Caulobacter crescentus. Genes Dev. 2008;22(2):212–225. doi: 10.1101/gad.1601808 18198338.

29. Wheeler RT, Shapiro L. Differential localization of two histidine kinases controlling bacterial cell differentiation. Mol Cell. 1999;4(5):683–694. doi: 10.1016/s1097-2765(00)80379-2 10619016.

30. Ohta N, Lane T, Ninfa EG, Sommer JM, Newton A. A histidine protein kinase homologue required for regulation of bacterial cell division and differentiation. Proc Natl Acad Sci U S A. 1992;89(21):10297–10301. doi: 10.1073/pnas.89.21.10297 1438215.

31. Abel S, Bucher T, Nicollier M, Hug I, Kaever V, Abel Zur Wiesch P, et al. Bi-modal distribution of the second messenger c-di-GMP controls cell fate and asymmetry during the caulobacter cell cycle. PLoS Genet. 2013;9(9):e1003744. doi: 10.1371/journal.pgen.1003744 24039597.

32. Lori C, Ozaki S, Steiner S, Bohm R, Abel S, Dubey BN, et al. Cyclic di-GMP acts as a cell cycle oscillator to drive chromosome replication. Nature. 2015;523(7559):236–239. doi: 10.1038/nature14473 25945741.

33. Hinz AJ, Larson DE, Smith CS, Brun YV. The Caulobacter crescentus polar organelle development protein PodJ is differentially localized and is required for polar targeting of the PleC development regulator. Mol Microbiol. 2003;47(4):929–941. doi: 10.1046/j.1365-2958.2003.03349.x 12581350.

34. Viollier PH, Sternheim N, Shapiro L. Identification of a localization factor for the polar positioning of bacterial structural and regulatory proteins. Proc Natl Acad Sci U S A. 2002;99(21):13831–13836. doi: 10.1073/pnas.182411999 12370432.

35. Jenal U, Fuchs T. An essential protease involved in bacterial cell-cycle control. EMBO J. 1998;17(19):5658–5669. doi: 10.1093/emboj/17.19.5658 9755166.

36. McGrath PT, Iniesta AA, Ryan KR, Shapiro L, McAdams HH. A dynamically localized protease complex and a polar specificity factor control a cell cycle master regulator. Cell. 2006;124(3):535–547. doi: 10.1016/j.cell.2005.12.033 16469700.

37. Iniesta AA, McGrath PT, Reisenauer A, McAdams HH, Shapiro L. A phospho-signaling pathway controls the localization and activity of a protease complex critical for bacterial cell cycle progression. Proc Natl Acad Sci U S A. 2006;103(29):10935–10940. doi: 10.1073/pnas.0604554103 16829582.

38. Joshi KK, Berge M, Radhakrishnan SK, Viollier PH, Chien P. An adaptor hierarchy regulates proteolysis during a bacterial cell cycle. Cell. 2015;163(2):419–431. doi: 10.1016/j.cell.2015.09.030 26451486.

39. Duerig A, Abel S, Folcher M, Nicollier M, Schwede T, Amiot N, et al. Second messenger-mediated spatiotemporal control of protein degradation regulates bacterial cell cycle progression. Genes Dev. 2009;23(1):93–104. doi: 10.1101/gad.502409 19136627.

40. Ozaki S, Schalch-Moser A, Zumthor L, Manfredi P, Ebbensgaard A, Schirmer T, et al. Activation and polar sequestration of PopA, a c-di-GMP effector protein involved in Caulobacter crescentus cell cycle control. Mol Microbiol. 2014;94(3):580–594. doi: 10.1111/mmi.12777 25171231.

41. Barnett MJ, Hung DY, Reisenauer A, Shapiro L, Long SR. A homolog of the CtrA cell cycle regulator is present and essential in Sinorhizobium meliloti. J Bacteriol. 2001;183(10):3204–3210. doi: 10.1128/JB.183.10.3204-3210.2001 11325950.

42. Bellefontaine AF, Pierreux CE, Mertens P, Vandenhaute J, Letesson JJ, de Bolle X. Plasticity of a transcriptional regulation network among alpha-proteobacteria is supported by the identification of CtrA targets in Brucella abortus. Mol Microbiol. 2002;43(4):945–960. doi: 10.1046/j.1365-2958.2002.02777.x 11929544.

43. Brown PJ, de Pedro MA, Kysela DT, Van der Henst C, Kim J, de Bolle X, et al. Polar growth in the Alphaproteobacterial order Rhizobiales. Proc Natl Acad Sci U S A. 2012;109(5):1697–1701. doi: 10.1073/pnas.1114476109 22307633.

44. Kim J, Heindl JE, Fuqua C. Coordination of division and development influences complex multicellular behavior in Agrobacterium tumefaciens. PLoS One. 2013;8(2):e56682. doi: 10.1371/journal.pone.0056682 23437210.

45. Bird TH, MacKrell A. A CtrA homolog affects swarming motility and encystment in Rhodospirillum centenum. Arch Microbiol. 2011;193(6):451–459. doi: 10.1007/s00203-011-0676-y 21243338.

46. Francez-Charlot A, Kaczmarczyk A, Vorholt JA. The branched CcsA/CckA-ChpT-CtrA phosphorelay of Sphingomonas melonis controls motility and biofilm formation. Mol Microbiol. 2015;97(1):47–63. doi: 10.1111/mmi.13011 25825287.

47. Greene SE, Brilli M, Biondi EG, Komeili A. Analysis of the CtrA pathway in Magnetospirillum reveals an ancestral role in motility in Alphaproteobacteria. J Bacteriol. 2012;194(11):2973–2986. doi: 10.1128/JB.00170-12 22467786.

48. Lang AS, Beatty JT. Genetic analysis of a bacterial genetic exchange element: the gene transfer agent of Rhodobacter capsulatus. Proc Natl Acad Sci U S A. 2000;97(2):859–864. doi: 10.1073/pnas.97.2.859 10639170.

49. Jung A, Eisheuer S, Cserti E, Leicht O, Strobel W, Möll A, et al. Molecular toolbox for genetic manipulation of the stalked budding bacterium Hyphomonas neptunium. Appl Environ Microbiol. 2015;81(2):736–744. doi: 10.1128/AEM.03104-14 25398860.

50. Leifson E. Hyphomicrobium neptunium sp. n. Antonie Van Leeuwenhoek. 1964;30:249–256. doi: 10.1007/bf02046730 14221685.

51. Badger JH, Eisen JA, Ward NL. Genomic analysis of Hyphomonas neptunium contradicts 16S rRNA gene-based phylogenetic analysis: implications for the taxonomy of the orders ' Rhodobacterales ' and Caulobacterales . Int J Syst Evol Microbiol. 2005;55(Pt 3):1021–1026. doi: 10.1099/ijs.0.63510–0 15879228.

52. Hirsch P. Budding bacteria. Annu Rev Microbiol. 1974;28(0):391–444. doi: 10.1146/annurev.mi.28.100174.002135 4611332.

53. Moore RL. The biology of Hyphomicrobium and other prosthecate, budding bacteria. Annu Rev Microbiol. 1981;35:567–594. doi: 10.1146/annurev.mi.35.100181.003031 6170249.

54. Hecht GB, Lane T, Ohta N, Sommer JM, Newton A. An essential single domain response regulator required for normal cell division and differentiation in Caulobacter crescentus. EMBO J. 1995;14(16):3915–3924. 7664732.

55. Sommer JM, Newton A. Pseudoreversion analysis indicates a direct role of cell division genes in polar morphogenesis and differentiation in Caulobacter crescentus. Genetics. 1991;129(3):623–630. 1752411.

56. Ely B, Croft RH, Gerardot CJ. Genetic mapping of genes required for motility in Caulobacter crescentus. Genetics. 1984;108(3):523–532. 6437899.

57. Childers WS, Xu Q, Mann TH, Mathews II, Blair JA, Deacon AM, et al. Cell fate regulation governed by a repurposed bacterial histidine kinase. PLoS Biol. 2014;12(10):e1001979. doi: 10.1371/journal.pbio.1001979 25349992.

58. Deich J, Judd EM, McAdams HH, Moerner WE. Visualization of the movement of single histidine kinase molecules in live Caulobacter cells. Proc Natl Acad Sci U S A. 2004;101(45):15921–15926. doi: 10.1073/pnas.0404200101 15522969.

59. Brassinga AK, Siam R, McSween W, Winkler H, Wood D, Marczynski GT. Conserved response regulator CtrA and IHF binding sites in the alpha-proteobacteria Caulobacter crescentus and Rickettsia prowazekii chromosomal replication origins. J Bacteriol. 2002;184(20):5789–5799. doi: 10.1128/JB.184.20.5789-5799.2002 12270838.

60. Brilli M, Fondi M, Fani R, Mengoni A, Ferri L, Bazzicalupo M, et al. The diversity and evolution of cell cycle regulation in alpha-proteobacteria: a comparative genomic analysis. BMC Syst Biol. 2010;4:52. doi: 10.1186/1752-0509-4-52 20426835.

61. Francis N, Poncin K, Fioravanti A, Vassen V, Willemart K, Ong TA, et al. CtrA controls cell division and outer membrane composition of the pathogen Brucella abortus. Mol Microbiol. 2017;103(5):780–797. doi: 10.1111/mmi.13589 27893179.

62. Mignolet J, Panis G, Viollier PH. More than a Tad: spatiotemporal control of Caulobacter pili. Curr Opin Microbiol. 2018;42:79–86. doi: 10.1016/j.mib.2017.10.017 29161615.

63. Domian IJ, Reisenauer A, Shapiro L. Feedback control of a master bacterial cell-cycle regulator. Proc Natl Acad Sci U S A. 1999;96(12):6648–6653. doi: 10.1073/pnas.96.12.6648 10359766.

64. Gora KG, Tsokos CG, Chen YE, Srinivasan BS, Perchuk BS, Laub MT. A cell-type-specific protein-protein interaction modulates transcriptional activity of a master regulator in Caulobacter crescentus. Mol Cell. 2010;39(3):455–467. doi: 10.1016/j.molcel.2010.06.024 20598601.

65. Tan MH, Kozdon JB, Shen X, Shapiro L, McAdams HH. An essential transcription factor, SciP, enhances robustness of Caulobacter cell cycle regulation. Proc Natl Acad Sci U S A. 2010;107(44):18985–18990. doi: 10.1073/pnas.1014395107 20956288.

66. Jung A, Rassbach A, Pulpetta RL, van Teeseling MCF, Heinrich K, Sobetzko P, et al. Two-step chromosome segregation in the stalked budding bacterium Hyphomonas neptunium. Nat Commun. 2019;10(1):3290. doi: 10.1038/s41467-019-11242-5 31337764.

67. Pini F, De Nisco NJ, Ferri L, Penterman J, Fioravanti A, Brilli M, et al. Cell cycle control by the master regulator CtrA in Sinorhizobium meliloti. PLoS Genet. 2015;11(5):e1005232. doi: 10.1371/journal.pgen.1005232 25978424.

68. Willett JW, Herrou J, Briegel A, Rotskoff G, Crosson S. Structural asymmetry in a conserved signaling system that regulates division, replication, and virulence of an intracellular pathogen. Proc Natl Acad Sci U S A. 2015;112(28):E3709–3718. doi: 10.1073/pnas.1503118112 26124143.

69. Fields AT, Navarrete CS, Zare AZ, Huang Z, Mostafavi M, Lewis JC, et al. The conserved polarity factor podJ1 impacts multiple cell envelope-associated functions in Sinorhizobium meliloti. Mol Microbiol. 2012;84(5):892–920. doi: 10.1111/j.1365-2958.2012.08064.x 22553970.

70. Pini F, Frage B, Ferri L, De Nisco NJ, Mohapatra SS, Taddei L, et al. The DivJ, CbrA and PleC system controls DivK phosphorylation and symbiosis in Sinorhizobium meliloti. Mol Microbiol. 2013;90(1):54–71. doi: 10.1111/mmi.12347 23909720.

71. Chen JC, Hottes AK, McAdams HH, McGrath PT, Viollier PH, Shapiro L. Cytokinesis signals truncation of the PodJ polarity factor by a cell cycle-regulated protease. EMBO J. 2006;25(2):377–386. doi: 10.1038/sj.emboj.7600935 16395329.

72. Jacobs C, Ausmees N, Cordwell SJ, Shapiro L, Laub MT. Functions of the CckA histidine kinase in Caulobacter cell cycle control. Mol Microbiol. 2003;47(5):1279–1290. Epub 2003/02/27. doi: 10.1046/j.1365-2958.2003.03379.x 12603734.

73. Paul R, Weiser S, Amiot NC, Chan C, Schirmer T, Giese B, et al. Cell cycle-dependent dynamic localization of a bacterial response regulator with a novel di-guanylate cyclase output domain. Genes Dev. 2004;18(6):715–727. doi: 10.1101/gad.289504 15075296.

74. Aldridge P, Jenal U. Cell cycle-dependent degradation of a flagellar motor component requires a novel-type response regulator. Mol Microbiol. 1999;32(2):379–391. doi: 10.1046/j.1365-2958.1999.01358.x 10231493.

75. Hecht GB, Newton A. Identification of a novel response regulator required for the swarmer-to-stalked-cell transition in Caulobacter crescentus. J Bacteriol. 1995;177(21):6223–6229. doi: 10.1128/jb.177.21.6223-6229.1995 7592388.

76. Sanselicio S, Berge M, Theraulaz L, Radhakrishnan SK, Viollier PH. Topological control of the Caulobacter cell cycle circuitry by a polarized single-domain PAS protein. Nat Commun. 2015;6:7005. doi: 10.1038/ncomms8005 25952018.

77. Lasker K, von Diezmann L, Zhou X, Ahrens DG, Mann TH, Moerner WE, et al. Selective sequestration of signalling proteins in a membraneless organelle reinforces the spatial regulation of asymmetry in Caulobacter crescentus. Nat Microbiol. 2020; doi: 10.1038/s41564-019-0647-7 31959967.

78. Mann TH, Childers WS, Blair JA, Eckart MR, Shapiro L. A cell cycle kinase with tandem sensory PAS domains integrates cell fate cues. Nat Commun. 2016;7:11454. doi: 10.1038/ncomms11454 27117914.

79. Chen YE, Tropini C, Jonas K, Tsokos CG, Huang KC, Laub MT. Spatial gradient of protein phosphorylation underlies replicative asymmetry in a bacterium. Proc Natl Acad Sci U S A. 2011;108(3):1052–1057. doi: 10.1073/pnas.1015397108 21191097.

80. Tropini C, Rabbani N, Huang KC. Physical constraints on the establishment of intracellular spatial gradients in bacteria. BMC Biophys. 2012;5:17. doi: 10.1186/2046-1682-5-17 22931750.

81. Reisenauer A, Quon K, Shapiro L. The CtrA response regulator mediates temporal control of gene expression during the Caulobacter cell cycle. J Bacteriol. 1999;181(8):2430–2439. 10198005.

82. Siam R, Marczynski GT. Cell cycle regulator phosphorylation stimulates two distinct modes of binding at a chromosome replication origin. EMBO J. 2000;19(5):1138–1147. doi: 10.1093/emboj/19.5.1138 10698954.

83. Siam R, Marczynski GT. Glutamate at the phosphorylation site of response regulator CtrA provides essential activities without increasing DNA binding. Nucleic Acids Res. 2003;31(6):1775–1779. doi: 10.1093/nar/gkg271 12626719.

84. Evinger M, Agabian N. Envelope-associated nucleoid from Caulobacter crescentus stalked and swarmer cells. J Bacteriol. 1977;132(1):294–301. 334726.

85. Meisenzahl AC, Shapiro L, Jenal U. Isolation and characterization of a xylose-dependent promoter from Caulobacter crescentus. J Bacteriol. 1997;179(3):592–600. doi: 10.1128/jb.179.3.592-600.1997 9006009.

86. Thanbichler M, Shapiro L. MipZ, a spatial regulator coordinating chromosome segregation with cell division in Caulobacter. Cell. 2006;126(1):147–162. doi: 10.1016/j.cell.2006.05.038 16839883.

87. Gilchrist A, Smit J. Transformation of freshwater and marine caulobacters by electroporation. J Bacteriol. 1991;173(2):921–925. doi: 10.1128/jb.173.2.921-925.1991 1987172.

88. Huang L. A new mechanistic growth model for simultaneous determination of lag phase duration and exponential growth rate and a new Bĕlehdrádek-type model for evaluating the effect of temperature on growth rate. Food Microbiol. 2011;28(4):770–776. doi: 10.1016/j.fm.2010.05.019 21511137.

89. Schneider CA, Rasband WS, Eliceiri KW. NIH Image to ImageJ: 25 years of image analysis. Nat Methods. 2012;9(7):671–675. doi: 10.1038/nmeth.2089 22930834.

90. Karimova G, Pidoux J, Ullmann A, Ladant D. A bacterial two-hybrid system based on a reconstituted signal transduction pathway. Proc Natl Acad Sci U S A. 1998;95(10):5752–5756. doi: 10.1073/pnas.95.10.5752 9576956.

91. Skerker JM, Prasol MS, Perchuk BS, Biondi EG, Laub MT. Two-component signal transduction pathways regulating growth and cell cycle progression in a bacterium: a system-level analysis. PLoS Biol. 2005;3(10):e334. doi: 10.1371/journal.pbio.0030334 16176121.

92. Zhang Y, Liu T, Meyer CA, Eeckhoute J, Johnson DS, Bernstein BE, et al. Model-based analysis of ChIP-Seq (MACS). Genome Biology. 2008;9:R137. doi: 10.1186/gb-2008-9-9-r137 18798982

93. Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, et al. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 2009;37:W202–208. doi: 10.1093/nar/gkp335 19458158.

94. Mao F, Dam P, Chou J, Olman V, Xu Y. DOOR: a database for prokaryotic operons. Nucleic Acids Res. 2009;37(Database issue):D459–463. doi: 10.1093/nar/gkn757 18988623.

95. de Hoon MJ, Imoto S, Nolan J, Miyano S. Open source clustering software. Bioinformatics. 2004;20(9):1453–1454. doi: 10.1093/bioinformatics/bth078 14871861.

96. Saldanha AJ. Java Treeview–extensible visualization of microarray data. Bioinformatics. 2004;20(17):3246–3248. doi: 10.1093/bioinformatics/bth349 15180930

97. Wali TM, Hudson GR, Danald DA, Weiner RM. Timing of swarmer cell cycle morphogenesis and macromolecular synthesis by Hyphomicrobium neptunium in synchronous culture. J Bacteriol. 1980;144(1):406–412. 6158509.

98. Kanehisa M, Sato Y, Kawashima M, Furumichi M, Tanabe M. KEGG as a reference resource for gene and protein annotation. Nucleic Acids Res. 2016;44(D1):D457–462. doi: 10.1093/nar/gkv1070 26476454.

99. Consortium TU. UniProt: the universal protein knowledgebase. Nucleic Acids Res. 2017;45(D1):D158–169. doi: 10.1093/nar/gkw1099 27899622.

100. Letunic I, Bork P. 20 years of the SMART protein domain annotation resource. Nucleic Acids Res. 2018;46(D1):D493–496. doi: 10.1093/nar/gkx922 29040681.

101. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. Basic local alignment search tool. J Mol Biol. 1990;215(3):403–410. doi: 10.1016/S0022-2836(05)80360-2 2231712.

102. Overmars L, Kerkhoven R, Siezen RJ, Francke C. MGcV: the microbial genomic context viewer for comparative genome analysis. BMC Genomics. 2013;14:209. doi: 10.1186/1471-2164-14-209 23547764.


Článek vyšel v časopise

PLOS Genetics


2020 Číslo 4
Nejčtenější tento týden
Nejčtenější v tomto čísle
Kurzy

Zvyšte si kvalifikaci online z pohodlí domova

Hypertenze a hypercholesterolémie – synergický efekt léčby
nový kurz
Autoři: prof. MUDr. Hana Rosolová, DrSc.

Multidisciplinární zkušenosti u pacientů s diabetem
Autoři: Prof. MUDr. Martin Haluzík, DrSc., prof. MUDr. Vojtěch Melenovský, CSc., prof. MUDr. Vladimír Tesař, DrSc.

Úloha kombinovaných preparátů v léčbě arteriální hypertenze
Autoři: prof. MUDr. Martin Haluzík, DrSc.

Halitóza
Autoři: MUDr. Ladislav Korábek, CSc., MBA

Terapie roztroušené sklerózy v kostce
Autoři: MUDr. Dominika Šťastná, Ph.D.

Všechny kurzy
Přihlášení
Zapomenuté heslo

Zadejte e-mailovou adresu, se kterou jste vytvářel(a) účet, budou Vám na ni zaslány informace k nastavení nového hesla.

Přihlášení

Nemáte účet?  Registrujte se

#ADS_BOTTOM_SCRIPTS#