Skip to main content

ORIGINAL RESEARCH article

Front. Microbiol., 02 November 2018
Sec. Evolutionary and Genomic Microbiology
This article is part of the Research Topic DNA Replication Origins in Microbial Genomes, Volume II View all 12 articles

HipA-Mediated Phosphorylation of SeqA Does not Affect Replication Initiation in Escherichia coli

  • 1Section for Functional Genomics, Department of Biology, Center for Bacterial Stress Response and Persistence, University of Copenhagen, Copenhagen, Denmark
  • 2Laboratoire de Chimie Bactérienne, Université Aix-Marseille, CNRS, Marseille, France

The SeqA protein of Escherichia coli is required to prevent immediate re-initiation of chromosome replication from oriC. The SeqA protein is phosphorylated at the serine-36 (Ser36) residue by the HipA kinase. The role of phosphorylation was addressed by mutating the Ser36 residue to alanine, which cannot be phosphorylated and to aspartic acid, which mimics a phosphorylated serine residue. Both mutant strains were similar to wild-type with respect to origin concentration and initiation synchrony. The minimal time between successive initiations was also unchanged. We therefore suggest that SeqA phosphorylation at the Ser36 residue is silent, at least with respect to SeqA's role in replication initiation.

Introduction

In Escherichia coli the DnaA initiator protein binds ATP and ADP with equal affinity (Sekimizu et al., 1987). DnaA binds three high-affinity sites in the origin, oriC, throughout the cell cycle irrespective of the bound nucleotide. The relative amounts of DnaAATP and DnaAADP, respectively fluctuate during the cell cycle with the DnaAATP/DnaAADP ratio peaking at initiation (Kurokawa et al., 1999). This results in binding of a number of additional DnaA binding sites of low affinity and with a preference for DnaAATP (Skarstad and Katayama, 2013; Leonard and Grimwade, 2015; Katayama et al., 2017). This induces origin opening, allows for helicase loading and replisome assembly.

Immediate re-initiation of new and hemimethylated origins is prevented by SeqA-binding to 11 GATC sites located within the minimal oriC (Campbell and Kleckner, 1990; Lu et al., 1994; Boye et al., 2000). The binding of SeqA to the origin prolongs the duration of the DNA hemi-methylated phase; a process called sequestration. Sequestration lasts approximately one-third of a cell cycle where re-initiation is prevented by SeqA denying DnaAATP access to GATC-containing low affinity DnaA boxes in oriC (Nievera et al., 2006). The sequestration period allows the cells to distinguish between “old” and “new” origins, and provides a time window where the DnaAATP level is lowered by RIDA (Kato and Katayama, 2001) and DDAH (Kasho and Katayama, 2013). Sequestration is finally terminated when GATC sequences within oriC become fully methylated by Dam methyltransferase.

In seqA mutant cells the sequestration period is shortened or absent (von Freiesleben et al., 2000), re-initiations occur frequently leading to over-initiation, and replication initiation becomes highly asynchronous (Lu et al., 1994). Conversely, excess SeqA protein prolongs the sequestration period, delays initiation, but does not affect initiation synchrony (Fossum et al., 2003; Charbon et al., 2011).

The SeqA protein contains two functional domains, an N-terminal oligomerization domain (SeqA-N; residues 1–33) and a C-terminal DNA-binding domain (SeqA-C; residues 65–181), which are joined by a flexible linker (residues 34–64; Chung et al., 2009). The interaction of SeqA with DNA occurs mainly in the major groove of the hemimethylated GATC sequences (Guarné et al., 2002), and data have suggested that two adjacent GATC sequences, up to 31 bp apart, interacting with the SeqA dimer are sufficient for strong binding (Guarné et al., 2005).

Recently, a stable isotope labeling by amino acids in cell culture (SILAC)-based quantitative phosphoproteomic approach combined with high-resolution mass spectrometry identified residue serine-36 (Ser36) in SeqA as a direct phosphorylation target for the kinase activity of the high persister protein A, HipA (Semanjski et al., 2018). HipA is an eukaryotic-like serine-threonine protein kinase that induces the stringent response, inhibits cell growth and confers cellular persistence through phosphorylation and inactivation of the glutamyl-tRNA-synthetase, GltX (Germain et al., 2013; Kaspy et al., 2013; Semanjski et al., 2018). The hipA gene constitutes a type II TA module with the adjacent upstream hipB gene, encoding the HipB antitoxin. HipB interacts directly with HipA to form a protein complex that represses the hipBA operon through binding to operators in the hipBA promoter region (Black et al., 1994), thereby counteracting the negative effect on cell growth caused by even low amounts of wild-type HipA (Korch and Hill, 2006).

It is not known whether phosphorylation at residue Ser36 of SeqA affects the activity and function of SeqA. Adding a phosphate group with negative charge to a protein, can promote changes in the structural conformation by altering the interactions with nearby amino acids. This might activate or inhibit the activity of the protein (Chao et al., 2014) or result in function modifications (Johnson and Barford, 1993).

Here, we tested the effect of Ser36 phosphorylation of SeqA on chromosome replication initiation. Two variants of SeqA were constructed, in which the Ser36 residue was either mutated to alanine (S36A) or aspartic acid (S36D). The S36A mutation impairs Ser36 phosphorylation, whereas the S36D mutation mimics the conformation of Ser36 phosphorylated SeqA (i.e., phospho-mimetic; (Arany et al., 2013). As both seqA mutants were similar to the wild-type with respect to synchrony and length of the sequestration period, our data suggest that HipA-mediated Ser36 phosphorylation of SeqA constitutes a neutral effect on the role of SeqA in E. coli replication initiation.

Materials and Methods

Media and Growth Conditions

Cells were grown in AB minimal medium (Clark and Maaløe, 1967) supplemented with 1 μg/ml thiamine, 0.2% glucose and 0.5% casamino acids (glucose-CAA medium). When necessary, antibiotic selection was maintained at the following final concentrations: kanamycin, 50 μg/ml; chloramphenicol, 20 μg/ml; tetracycline, 10 μg/ml; ampicillin, 150 μg/ml. All cells were cultured at 37°C, except when otherwise indicated. Cell growth was monitored by measuring optical density at 450 nm (OD450).

Bacterial Strains

All strains used were derived from E. coli K-12 MG1655 (F, λ, rph-1; Guyer et al., 1981) and are listed in Table 1. The ΔhipBA::frt::kan::frt (Germain et al., 2013) and dnaA46 tnaA600::Tn10 (Kogoma and von Meyenburg, 1983) alleles were moved by P1-phage-mediated transduction (Miller, 1972). To construct the chromosomal seqA mutant strains (seqAS36A and seqAS36D, respectively), base substitutions were made in the codon for Ser36 (5′-TCC-3′ to 5′-GCC-3′; Ser(S) to Ala(A), and 5′-TCC-3′ to 5′-GAC-3′; Ser(S) to Asp(D), respectively) using splicing by overlap extension (SOEing) polymerase chain reaction (PCR) (Horton et al., 1989). All primers are listed in Table 2. For each seqA variant two initial PCR products of the MG1655 chromosome were generated. 1) The lower region of the seqA gene, spanning residues 27–181, was amplified using primers “SeqA_down_bw_XmaI” and either “SeqA_pos36_SA_fw” or “SeqA_pos36_SD_fw”. 2) The upper region of the seqA gene was amplified using primers “SeqA_up_fw_SacI” and “SeqA_intern_bw” that generates a fragment with an overlap of 21 bp with the seqA downstream PCR product. A secondary amplification was performed using equimolar ratios of the two PCR products as template, and the oligonucleotides, SeqA_down_bw_XmaI, and SeqA_up_fw_SacI, as primers. The resulting PCR fragments were digested with XmaI and SacI, and cloned into the same sites of the 3.9 kb suicide vector, pRUC1437, a derivative of pSW29T (Demarre et al., 2005), carrying the aph gene encoding kanaymicn resistance, and the sacB gene. The resulting plasmids were transformed into strain S17-1 (recA thi pro hsdRM+ RP4-2 Tc::Mu-Km::Tn7 λ-pir lysogen TpR SmR; (Simon et al., 1983) before being transferred into ALO 2956 cells by conjugation. Selection of exconjugants carrying the chromosomally integrated recombinant suicide plasmids as well as subsequent sucrose-mediated selection for loss of the sacB gene (i.e., loss of suicide vector sequences; (Donnenberg and Kaper, 1991), leaving either a wild-type or a mutant variant of the seqA gene on the MG1655 lacIZYA::cat chromosome, was performed as described previously (Riber et al., 2009). Chromosomal seqA mutant strains were verified by DNA sequencing of PCR fragments amplified from the seqA region using DNA oligonucleotides, “SeqA_chr_fw” and “SeqA_chr_bw,” as primers.

TABLE 1
www.frontiersin.org

Table 1. Bacterial strains.

TABLE 2
www.frontiersin.org

Table 2. Primers.

Plasmids

All plasmids used are listed in Table 3. Plasmids pLR77 and pLR75 were constructed by PCR amplifying the seqA variant genes (including the native seqA ribosome binding site) from MG1655 lacIZYA::cat cells carrying either the seqAS36A or seqAS36D chromosomal genes (see above), respectively, using DNA oligonucleotides, SeqA_up_fw_EcoRI and SeqA_down_bw_HindIII, as primers. The resultant PCR fragments were digested with EcoRI and HindIII and inserted downstream the IPTG inducible lacPA1−04/03 promoter (Lanzer and Bujard, 1988) of plasmid pFH2102 (von Freiesleben et al., 2000), cut with the same enzymes. The inserted seqA mutant genes were later verified by DNA sequencing.

TABLE 3
www.frontiersin.org

Table 3. Plasmids.

Flow Cytometry and Cell Cycle Analysis

Exponentially growing cells (OD450 = 0.15–0.30) were treated with rifampicin (300 μg/ml; SERVA Electrophoresis GmbH) and cephalexin (36 μg/ml; Sigma-Aldrich) to inhibit initiation of DNA replication and cell division, respectively (Løbner-Olesen et al., 1989). Incubation continued for a minimum of 4 h at 37°C to allow completion of ongoing rounds of replication. Cells were fixed in 70% ethanol and stained with 90 μg/ml mithramycin (SERVA Electrophoresis GmbH) and 20 μg/ml ethidium bromide (Sigma-Aldrich) as described (Løbner-Olesen et al., 1989). Flow cytometry was performed as previously described (Løbner-Olesen et al., 1989) using an Apogee A10 instrument (Apogee, Inc.). For all samples a minimum of 50.000 cells were analyzed. Numbers of origins per cell and relative cell mass were determined as previously described (Løbner-Olesen et al., 1989).

Immunoblot Procedure

Samples of 2 ml of exponentially growing cells (OD450 = 0.3–0.4) were harvested. Proteins were separated by SDS-PAGE and SeqA protein detected by Western blot using rabbit antiserum raised against SeqA protein (Torheim et al., 2000) as previously described (Riber and Lobner-Olesen, 2005). The membrane was scanned using a 230 V GenoView imaging system equipped with a UV transilluminator (VWR). Quantification was done using the ImageJ software.

Multiple Sequence Alignment Analysis

Multiple alignment analysis of SeqA amino acids sequences was performed in the MEGA version 7.0.26 software (Kumar et al., 2016) using the default settings of the integrated ClustalW algorithm (Larkin et al., 2007). Selected species including SeqA protein accession numbers were: Escherichia coli K-12 (accession: AAA19855.1), Vibrio cholerae (accession: AOY47782.1), Pasteurella multocida PM70 (accession: AAK02440.1), Haemophilus influenzae Rd (accession: NP_438362.1), Yersinia enterocolitica (accession: CNB62546.1), Serratia marcescens (accession: KFL03527.1), Actinobacillus pleuropneumoniae (accession: SQF64393.1), and Glaesserella parasuis (accession: STO80764.1).

Results

Changing the SeqA Ser36 Residue Mainly Affects the Linker Region of SeqA

In order to determine any putative role of SeqA phosphorylation, we generated two mutations at the chromosomal codon 36 of seqA. In one strain, the codon for Ser36 was replaced with that of an aspartic acid (seqAS36D). The SeqAS36D mimics the conformation of Ser36 phosphorylated SeqA (Arany et al., 2013). In a second strain the codon for Ser36 was replaced with that of an alanine (seqAS36A). The resulting protein, SeqAS36A, is phosphorylation impaired at position 36 (Arany et al., 2013).

We used the RaptorX web server (Källberg et al., 2012) to predict the tertiary structures (Peng and Xu, 2011; Ma et al., 2012) of the wild-type SeqA and the SeqAS36A and SeqAS36D proteins. This revealed a significant level of resemblance (Figure 1A). By pairwise and multiple structural alignments of the SeqA protein variants, TMScore values above 0.9 were obtained, illustrating a significantly increased likelihood (>90% of chance) that the proteins pairwise and all together share similar folds, RaptorX Structure Alignment Server; (Wang et al., 2011, 2013), with SeqA and SeqAS36D being structurally most alike [TMScore (WT vs. S36D) = 0.96]. The structural differences caused by changing the Ser36 residue seem to affect only the flexible linker region between SeqA-N and SeqA-C (Figure 1A; Chung et al., 2009).

FIGURE 1
www.frontiersin.org

Figure 1. Replication initiation is not affected by the seqAS36A and seqAS36D mutations. (A) Prediction of tertiary structures of SeqA, SeqAS36A and SeqAS36D proteins using the RaptorX Structure Prediction web server (Källberg et al., 2012). (B) Wild-type, seqAS36A, seqAS36D, and ΔseqA cells were grown at 37°C in AB minimal medium supplemented with glucose and casamino acids. Cells were treated with rifampicin and cephalexin prior to flow cytometric analysis. Cell cycle parameters are shown in the insert. “Ori/cell” represents the average number of origins per cell, whereas “Ori/mass” represents the origin concentration. “Mass” and “Ori/mass” measures are relative to wild-type cells. (C) SeqA protein content determined by Western blot analysis. All quantifications are relative to wild-type cells. The relevant seqA genotype is indicated on the figure. (D) HipBA deficient cells were grown and subjected to flow cytometric analysis as described in (B) above.

Replication Initiation Is Not Affected by seqAS36A and seqAS36D Mutations

We used flow-cytometry to determine cell cycle parameters of wild-type and seqA mutants. The two seqA mutants grew with similar doubling times as wild-type cells in minimal medium supplemented with glucose and casamino acids, whereas cells deficient in SeqA grew with an ~30% increased doubling time relative to that of wild-type cells (Figure 1B). Following treatment with rifampicin and cephalexin, wild-type, seqAS36A and seqAS36D cells were similar and contained mainly 2, 4, or 8 fully replicated chromosomes, indicative of initiation synchrony (Skarstad et al., 1986). As the average cell mass and numbers of origins per cell were similar, so was the origin concentration between these three cell types. SeqA deficient cells showed an asynchronous initiation phenotype with an increased average number of origins, which illustrates a lost ability to negatively regulate replication initiation. The average cell mass was similar to that of wild-type cells resulting in an increased origin concentration (Figure 1B).

Because the SeqAS36A and SeqAS36D protein levels were comparable or slightly elevated relative to that of wild-type SeqA protein (Figure 1C), these data altogether suggest that phosphorylation of SeqA at position 36 has little influence on its activity in replication initiation control. This was further corroborated by analyzing cells deficient in the HipA kinase, i.e., with a knock-out of the hipA gene. Here we found that ΔhipBA::kan mutant cells displayed similar cell cycle parameters as wild-type cells (Figure 1D).

Overproduction of SeqA, SeqAS36A, or SeqAS36D Proteins All Restore Initiation Synchrony in ΔSeqA Mutant Cells

We proceeded to examine whether overexpression of wild-type and mutant SeqA proteins could reveal any difference in activity among the phospho-impaired (S36A), phospho-mimetic (S36D) and wild-type SeqA proteins.

We expressed the seqA, seqAS36A, and seqAS36D genes from the IPTG-inducible lacPA1/04−03 promoter in SeqA deficient cells. Exponentially growing cells were induced with 1 mM IPTG at time 0 min (T = 0 min). Immunoblot analysis of cells sampled at 120 min following the addition of IPTG indicated that all SeqA proteins were expressed to comparable levels corresponding to an ~12- to 14-fold increase in SeqA level relative to wild-type cells (Figure 2A). Both wild-type and mutant SeqA proteins complemented ΔseqA cells to the same extent when produced from a plasmid. Cells containing mainly two or four origins, indicative of initiation synchrony, dominated the population already after 30 min induction of the seqA variant genes (Figure 2B). A larger increase in mutant SeqA proteins (T = 120 min) resulted in no significant asynchrony relative to wild-type (Figure 2B). This is in agreement with earlier data on SeqA overproduction (Fossum et al., 2003).

FIGURE 2
www.frontiersin.org

Figure 2. Similar effects of SeqAWT, SeqAS36A, or SeqAS36D protein overproduction on replication initiation. SeqA deficient (ΔseqA) cells carrying the SeqA expression plasmids, pMAK7 (pLac-seqA), pLR77 (pLac-seqAS36A), pLR75 (pLac-seqAS36D), and pFH2102 (vector) were grown exponentially at 37°C in AB minimal medium supplemented with glucose and casamino acids. At time, T = 0 min (top panel), IPTG was added to a final concentration of 1 mM, and samples were subsequently removed at the indicated time points. (A) SeqA immunoblot sampled at 120 min. A sample of wild-type cells (without plasmid) is included to allow for relative quantification of SeqA levels. (B) Samples were taken at 0, 30, 60, and 120 min following IPTG induction and treated with rifampicin and cephalexin prior to flow cytometric analysis.

The Minimal Time Between Successive Initiations Is Not Altered by seqAS36A and seqAS36D Mutations

Changes in the duration of sequestration by increasing or decreasing the level of Dam methylase (von Freiesleben et al., 2000) or by increasing the SeqA level (Charbon et al., 2011) were previously found to have relatively modest effects on the cell cycle relative to complete loss of sequestration. We therefore proceeded to determine whether the SeqA mutant proteins affected the length of the sequestration period, defined as the minimal time between successive initiations (von Freiesleben et al., 2000).

We introduced the dnaA46 allele into seqAS36A and seqAS36D cells by P1-transduction. The resultant strains are initiation proficient at 30°C (permissive temperature), but not at 42°C (non-permissive temperature) due to a reversible defect in nucleotide binding (Carr and Kaguni, 1996). Wild-type, seqAS36A and seqAS36D cells carrying the dnaA46 allele were grown exponentially at 30°C. The average number of origins per cell for all three strains was close to 2 (Figures 3A–C) and the SeqA proteins were produced in similar amounts (Figure 3D). When cells were shifted to 42°C, initiations ceased whereas cells continued to grow and divide, resulting in most cells ending up having one fully replicated chromosome after 90 min (Figures 3A–C). Upon a shift back to 30°C, where the DnaA46 protein was reactivated, all cells initiated replication, i.e., doubled their origin content, within a short period of time. This round of initiation was followed by a period of ~20 min where all newly formed origins were inert to further initiation, after which replication initiation resumed (Figures 3A–C). This 20-min period represents the minimal time between successive initiations (von Freiesleben et al., 2000), and it did not differ between wild-type and seqA mutant cells (Figures 3A–C). SeqA deficient cells were previously shown to reinitiate frequently without this 20-min delay (von Freiesleben et al., 2000).

FIGURE 3
www.frontiersin.org

Figure 3. The seqAS36A and seqAS36D mutations do not change the minimal time between successive initiations. dnaA46 (A), dnaA46 seqAS36A (B), and dnaA46 seqAS36D (C) cells were grown exponentially at 30°C in AB minimal medium supplemented with glucose and casamino acids. At time T = −90 min the cultures were shifted to the non-permissive temperature (42°C) and at time T = 0 min (illustrated by the gray vertical lines) shifted back to 30°C. At the times indicated samples were removed for treatment with rifampicin and cephalexin prior to flow cytometric analysis. The median (the value above and below which 50% of the distribution can be found) was used as a robust measure of the central tendency of individual cells (von Freiesleben et al., 2000) and is plotted as origins per cell. Replication resumes by 2 min (1 to 2 ori/cell). The red vertical indicates a second roundof firing giving rise to 4 ori/cell (A–C). The panels on the right-hand side of the figure show selected DNA histograms for rifampicin-cephalexin treated cultures (D). SeqA protein content determined by Western blot analysis for wild-type, seqAS36A, seqAS36D, or ΔseqA cells carrying the dnaA46 allele. All quantifications are relative to SeqA+ cells.

Serine36 of SeqA Is Not Phylogenetically Conserved

We aligned SeqA amino acid sequences from the Vibrio cholerae, Pasteurella multocida PM70, Haemophilus influenzae Rd, Yersinia enterocolitica, Serratia marcescens, Actinobacillus pleuropneumoniae and Glaesserella parasuis with that of E. coli K12. All of these bacteria are known to carry hipBA genes. We looked for conservation of Ser36 along with the two flanking amino acids Phe35 and Ala37 (Table 4). None of these amino acids were conserved among the species with Ser36 showing the least degree of conservation. On the other hand Thr18, Ile21, and Ala25 which are instrumental in oligomerization of SeqA (Guarné et al., 2005), were completely conserved. For Arg116, Thr117, Arg118, Asn150, and Asn152 that make contact with the GATC sequence in DNA (Fujikawa et al., 2004) we also observed a high degree of conservation between species (Table 4). This may indicate a limited role of Ser36 for SeqA function.

TABLE 4
www.frontiersin.org

Table 4. Serine 36 of E. coli SeqA is not conserved between bacterial species.

Discussion

Recently, it was shown that residue Ser36 in the SeqA protein is a target for phosphorylation by the serine-threonine kinase, HipA (Semanjski et al., 2018). HipA is mostly known for its role in bacterial persister formation through phosphorylation of a conserved serine, Ser239, residue in the GltX aminoacyl-tRNA synthetase, which inactivates the enzyme to arrest cell growth (Germain et al., 2013; Kaspy et al., 2013). Here, we wanted to determine whether Ser36 phosphorylation could alter SeqA activity. It was tempting to speculate that the Ser36 phosphorylation would activate SeqA, thereby enhancing its inhibition of replication initiation, which would contribute to shut down chromosomal replication in persister cells. SeqA was found to be endogenous phosphorylated in wild-type E. coli cells, and was revealed as a direct phosphorylation target of HipA in vitro. When the hipA gene was expressed from a p15A based plasmid, the fraction of wild-type SeqA found to be phosphorylated at residue Ser36 was ~7% following 95 min induction (Semanjski et al., 2018). It could be argued that this is a relative small fraction of the total SeqA protein. However, one should be aware that the actual phosphorylation status of SeqA may depend on the specific conditions provided. In the Semanjski study HipA expression was countered by the antitoxin HipB produced from the chromosome. The fraction of phosphorylated SeqA may therefore not reflect the fraction of SeqA being phosphorylated during an actual stress-induced situation where HipA becomes fully induced without HipB-mediated neutralization, and where the overall protein synthesis is affected. Also, it remains unknown whether all SeqA molecules present in the cell are actually available to HipA-mediated phosphorylation. The oligomerization domain of SeqA (residue 1–33) is located close to the HipA phosphorylation domain at residue Ser36 (see below), and hence it is not clear whether SeqA oligomers are available to phosphorylation, or whether only SeqA monomers become phosphorylated.

The Ser36 residue is located in the flexible linker between the N-terminal oligomerization domain and the C-terminal DNA binding domain (Chung et al., 2009). Neither of the phospho-impaired (S36A) nor the phospho-mimetic (S36D) SeqA proteins have any change in linker length nor are they affected in prolin or other amino acid residues suggested as most preferred in linker regions (George and Heringa, 2002), suggesting that changes in flexibility and hydrophobicity are non-significant upon phosphorylation of Ser36. This agrees well with the tertiary structural predictions of the SeqA, SeqAS36A, and SeqAS36D proteins that indicated the mutations to cause minor structural changes to the linker region only, leaving the N- and C-terminal domains unaffected. This might explain our observations that function and activity of the SeqA mutant proteins seemed unaffected by the Ser36 mutations with respect to replication initiation control.

Although we have assumed that substituting a serine residue with a negatively charged amino acid, such as aspartic acid, imparts the negative charge associated with serine phosphorylation, caution should be taken as this is not always the case. The phospho-mimetic proteins may fail to recapitulate the true steric and charge-based nature of phosphorylation (Paleologou et al., 2008). Also, the “phosphorylation status” mimicked by phospho-mimetics is non-reversible, and hence cannot reflect the true state of phosphorylation-mediated protein modification. Therefore, the SeqAS36D protein may deviate in activity from the phosphorylated wild-type SeqA protein.

However, because removal of the HipA kinase in wild-type cells revealed no replication phenotype, we find it unlikely that HipA-mediated Ser36 phosphorylation affects the activity of SeqA, at least with respect to its function in replication initiation, and at least under the conditions provided in this study. SeqA phosphorylation may therefore be an example of a silent phosphorylation. This has previously been observed for pepsin and ovalbumin, where serine phosphorylation did not affect protein activity, and the function of the phosphate group remained unknown (Johnson and Barford, 1993). The proposal that SeqA phosphorylation is silent is reinforced by the low degree of Ser36 conservation between hipBA carrying bacterial species compared to highly conserved amino acids crucial for oligomerization and DNA binding activity.

Author Contributions

LR, EG, and AL-O planned the experiments. LR, BK, and LK performed the experiments. LR, BK, LK, and AL-O analyzed data. LR, BK, and AL-O wrote the manuscript.

Funding

This work was supported by the Center for Bacterial Stress Response and Persistence (BASP) by a grant from the Danish National Research Foundation (DNRF120).

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Acknowledgments

We thank Maja Semanjski and Prof. Boris Macek from the Proteome Center Tuebingen, Germany, for sharing their data on SeqA phosphorylation by HipA prior to publication.

References

Arany, I., Clark, J., Reed, D. K., and Juncos, L. A. (2013). Chronic nicotine exposure augments renal oxidative stress and injury through transcriptional activation of p66shc. Nephrol. Dial Trans. 28, 1417–1425. doi: 10.1093/ndt/gfs596

PubMed Abstract | CrossRef Full Text | Google Scholar

Black, D. S., Irwin, B., and Moyed, H. S. (1994). Autoregulation of hip, an operon that affects lethality due to inhibition of peptidoglycan or DNA synthesis. J. Bacteriol. 176, 4081–4091.

Google Scholar

Bolivar, F., Rodriguez, R. L., Greene, P. J., Betlach, M. C., Heyneker, H. L., Boyer, H. W., et al. (1977). Construction and characterization of new cloning vehicles. II. a multipurpose cloning system. Gene 2, 95–113.

PubMed Abstract | Google Scholar

Boye, E., Lobner-Olesen, A., and Skarstad, K. (2000). Limiting DNA replication to once and only once. EMBO Rep. 1, 479–483. doi: 10.1093/embo-reports/kvd116

PubMed Abstract | CrossRef Full Text | Google Scholar

Campbell, J. L., and Kleckner, N. (1990). E. coli oriC and the dnaA gene promoter are sequestered from dam methyltransferase following the passage of the chromosomal replication fork. Cell 62, 967–979.

PubMed Abstract | Google Scholar

Carr, K. M., and Kaguni, J. M. (1996). The A184V missense mutation of the dnaA5 and dnaA46 alleles confers a defect in ATP binding and thermolability in initiation of Escherichia coli DNA replication. Mol. Microbiol. 20, 1307–1318.

PubMed Abstract | Google Scholar

Chao, J. D., Wong, D., and Av-Gay, Y. (2014). Microbial protein-tyrosine kinases. J. Biol. Chem. 289, 9463–9472. doi: 10.1074/jbc.R113.520015

PubMed Abstract | CrossRef Full Text | Google Scholar

Charbon, G., Riber, L., Cohen, M., Skovgaard, O., Fujimitsu, K., Katayama, T., et al. (2011). Suppressors of DnaA(ATP) imposed overinitiation in Escherichia coli. Mol. Microbiol. 79, 914–928. doi: 10.1111/j.1365-2958.2010.07493.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Chung, Y. S., Brendler, T., Austin, S., and Guarne, A. (2009). Structural insights into the cooperative binding of SeqA to a tandem GATC repeat. Nucleic Acids Res. 37, 3143–3152. doi: 10.1093/nar/gkp151

PubMed Abstract | CrossRef Full Text | Google Scholar

Clark, D. J., and Maaløe, O. (1967). DNA replication and the division cycle in Escherichia coli. J. Mol. Biol. 23, 99–112.

Google Scholar

Demarre, G., Guerout, A. M., Matsumoto-Mashimo, C., Rowe-Magnus, D. A., Marliere, P., and Mazel, D. (2005). A new family of mobilizable suicide plasmids based on broad host range R388 plasmid (IncW) and RP4 plasmid (IncPalpha) conjugative machineries and their cognate Escherichia coli host strains. Res. Microbiol. 156, 245–255. doi: 10.1016/j.resmic.2004.09.007

PubMed Abstract | CrossRef Full Text | Google Scholar

Donnenberg, M. S., and Kaper, J. B. (1991). Construction of an eae deletion mutant of enteropathogenic Escherichia coli by using a positive-selection suicide vector. Infect. Immun. 59, 4310–4317.

PubMed Abstract | Google Scholar

Fossum, S., Søreide, S., and Skarstad, K. (2003). Lack of SeqA focus formation, specific DNA binding and proper protein multimerization in the Escherichia coli sequestration mutant seqA2. Mol. Microbiol. 47, 619–632. doi: 10.1046/j.1365-2958.2003.t01-1-03329.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Fujikawa, N., Kurumizaka, H., Nureki, O., Tanaka, Y., Yamazoe, M., Hiraga, S., et al. (2004). Structural and biochemical analyses of hemimethylated DNA binding by the SeqA protein. Nucleic Acids Res. 32, 82–92. doi: 10.1093/nar/gkh173

PubMed Abstract | CrossRef Full Text | Google Scholar

George, R. A., and Heringa, J. (2002). An analysis of protein domain linkers: their classification and role in protein folding. Protein Eng. 15, 871–879. doi: 10.1093/protein/15.11.871

PubMed Abstract | CrossRef Full Text | Google Scholar

Germain, E., Castro-Roa, D., Zenkin, N., and Gerdes, K. (2013). Molecular mechanism of bacterial persistence by HipA. Mol. Cell 52, 248–254. doi: 10.1016/j.molcel.2013.08.045

PubMed Abstract | CrossRef Full Text | Google Scholar

Guarné, A., Brendler, T., Zhao, Q., Ghirlando, R., Austin, S., and Yang, W. (2005). Crystal structure of a SeqA-N filament: implications for DNA replication and chromosome organization. EMBO J. 24, 1502–1511. doi: 10.1038/sj.emboj.7600634

PubMed Abstract | CrossRef Full Text | Google Scholar

Guarné, A., Zhao, Q., Ghirlando, R., and Yang, W. (2002). Insights into negative modulation of E. coli replication initiation from the structure of SeqA-hemimethylated DNA complex. Nat. Struct. Biol. 9, 839–843. doi: 10.1038/nsb857

PubMed Abstract | CrossRef Full Text | Google Scholar

Guyer, M. S., Reed, R. R., Steitz, J. A., and Low, K. B. (1981). Identification of a sex-factor-affinity site in E. coli as gamma delta. Cold Spring Harb. Symp. Quant. Biol. 45, 135–140.

PubMed Abstract | Google Scholar

Horton, R. M., Hunt, H. D., Ho, S. N., Pullen, J. K., and Pease, L. R. (1989). Engineering hybrid genes without the use of restriction enzymes: gene splicing by overlap extension. Gene 77, 61–68.

PubMed Abstract | Google Scholar

Johnson, L. N., and Barford, D. (1993). The effects of phosphorylation on the structure and function of proteins. Annu. Rev. Biophys. Biomol. Struct. 22, 199–232. doi: 10.1146/annurev.bb.22.060193.001215

PubMed Abstract | CrossRef Full Text | Google Scholar

Källberg, M., Wang, H., Wang, S., Peng, J., Wang, Z., Lu, H., et al. (2012). Template-based protein structure modeling using the RaptorX web server. Nat. Protoc. 7, 1511–1522. doi: 10.1038/nprot.2012.085

PubMed Abstract | CrossRef Full Text | Google Scholar

Kasho, K., and Katayama, T. (2013). DnaA binding locus datA promotes DnaA-ATP hydrolysis to enable cell cycle-coordinated replication initiation. Proc. Natl. Acad. Sci. U.S.A. 110, 936–941. doi: 10.1073/pnas.121207011

PubMed Abstract | CrossRef Full Text | Google Scholar

Kaspy, I., Rotem, E., Weiss, N., Ronin, I., Balaban, N. Q., and Glaser, G. (2013). HipA-mediated antibiotic persistence via phosphorylation of the glutamyl-tRNA-synthetase. Nat. Commun. 4:3001. doi: 10.1038/ncomms4001

PubMed Abstract | CrossRef Full Text | Google Scholar

Katayama, T., Kasho, K., and Kawakami, H. (2017). The DnaA Cycle in Escherichia coli: activation, function and inactivation of the initiator protein. Front. Microbiol. 8:2496. doi: 10.3389/fmicb.2017.02496

PubMed Abstract | CrossRef Full Text | Google Scholar

Kato, J., and Katayama, T. (2001). Hda, a novel DnaA-related protein, regulates the replication cycle in Escherichia coli. EMBO J. 20, 4253–4262. doi: 10.1093/emboj/20.15.4253

PubMed Abstract | CrossRef Full Text | Google Scholar

Kogoma, T., and von Meyenburg, K. (1983). The origin of replication, oriC, and the dnaA protein are dispensable in stable DNA replication (sdrA) mutants of Escherichia coli K-12. EMBO J. 2, 463–468.

PubMed Abstract | Google Scholar

Korch, S. B., and Hill, T. M. (2006). Ectopic overexpression of wild-type and mutant hipA genes in Escherichia coli: effects on macromolecular synthesis and persister formation. J. Bacteriol. 188, 3826–3836. doi: 10.1128/JB.01740-05

PubMed Abstract | CrossRef Full Text | Google Scholar

Kumar, S., Stecher, G., and Tamura, K. (2016). MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol. Biol. Evol. 33, 1870–1874. doi: 10.1093/molbev/msw054

PubMed Abstract | CrossRef Full Text | Google Scholar

Kurokawa, K., Nishida, S., Emoto, A., Sekimizu, K., and Katayama, T. (1999). Replication cycle-coordinated change of the adenine nucleotide-bound forms of DnaA protein in Escherichia coli. EMBO J. 18, 6642–6652.

PubMed Abstract | Google Scholar

Lanzer, M., and Bujard, H. (1988). Promoters largely determine the efficiency of repressor action. Proc. Natl. Acad. Sci. U.S.A. 85, 8973–8977.

PubMed Abstract | Google Scholar

Larkin, M. A., Blackshields, G., Brown, N. P., Chenna, R., McGettigan, P. A., McWilliam, H., et al. (2007). Clustal W and Clustal X version 2.0. Bioinformatics 23, 2947–2948. doi: 10.1093/bioinformatics/btm404

PubMed Abstract | CrossRef Full Text | Google Scholar

Leonard, A. C., and Grimwade, J. E. (2015). The orisome: structure and function. Front. Microbiol. 6:545. doi: 10.3389/fmicb.2015.00545

PubMed Abstract | CrossRef Full Text | Google Scholar

Løbner-Olesen, A., Skarstad, K., Hansen, F. G., von Meyenburg, K., and Boye, E. (1989). The DnaA protein determines the initiation mass of Escherichia coli K-12. Cell 57, 881–889.

PubMed Abstract | Google Scholar

Lu, M., Campbell, J. L., Boye, E., and Kleckner, N. (1994). SeqA: a negative modulator of replication initiation in E. coli. Cell 77, 413–426.

PubMed Abstract | Google Scholar

Ma, J., Peng, J., Wang, S., and Xu, J. (2012). A conditional neural fields model for protein threading. Bioinformatics 28, i59–66. doi: 10.1093/bioinformatics/bts213.

PubMed Abstract | CrossRef Full Text | Google Scholar

Miller, J. H. (1972). Experiments in Molecular Genetics. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory.

Nievera, C., Torgue, J. J., Grimwade, J. E., and Leonard, A. C. (2006). SeqA blocking of DnaA-oriC interactions ensures staged assembly of the E. coli pre-RC. Mol. Cell 24, 581–592. doi: 10.1016/j.molcel.2006.09.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Paleologou, K. E., Schmid, A. W., Rospigliosi, C. C., Kim, H. Y., Lamberto, G. R., Fredenburg, R. A., et al. (2008). Phosphorylation at Ser-129 but not the phosphomimics S129E/D inhibits the fibrillation of alpha-synuclein. J. Biol. Chem. 283, 16895–16905. doi: 10.1074/jbc.M800747200

PubMed Abstract | CrossRef Full Text | Google Scholar

Peng, J., and Xu, J. (2011). RaptorX: exploiting structure information for protein alignment by statistical inference. Proteins 10(Suppl. 79), 161–171. doi: 10.1002/prot.23175

CrossRef Full Text | Google Scholar

Riber, L., Fujimitsu, K., Katayama, T., and Lobner-Olesen, A. (2009). Loss of Hda activity stimulates replication initiation from I-box, but not R4 mutant origins in Escherichia coli. Mol. Microbiol. 71, 107–122. doi: 10.1111/j.1365-2958.2008.06516.x

CrossRef Full Text | Google Scholar

Riber, L., and Lobner-Olesen, A. (2005). Coordinated replication and sequestration of oriC and dnaA are required for maintaining controlled once-per-cell-cycle initiation in Escherichia coli. J. Bacteriol. 187, 5605–5613. doi: 10.1128/JB.187.16.5605-5613.2005

PubMed Abstract | CrossRef Full Text | Google Scholar

Sekimizu, K., Bramhill, D., and Kornberg, A. (1987). ATP activates dnaA protein in initiating replication of plasmids bearing the origin of the E. coli chromosome. Cell 50, 259–265.

PubMed Abstract | Google Scholar

Semanjski, M., Germain, E., Bratl, K., Kiessling, A., Gerdes, K., and Macek, B. (2018). The kinases HipA and HipA7 phosphorylate different substrate pools in Escherichia coli to promote multidrug tolerance. Sci. Signal. 11:eat5750. doi: 10.1126/scisignal.aat5750

PubMed Abstract | CrossRef Full Text | Google Scholar

Simon, R., Priefer, U., and Puhler, A. (1983). A broad host range mobilization system for in vivo genetic-engineering–transposon mutagenesis in gram-negative bacteria. Bio-Technology 1, 784–791. doi: 10.1038/nbt1183-784

CrossRef Full Text | Google Scholar

Skarstad, K., Boye, E., and Steen, H. B. (1986). Timing of initiation of chromosome replication in individual Escherichia coli cells. EMBO J. 5, 1711–1717.

PubMed Abstract | Google Scholar

Skarstad, K., and Katayama, T. (2013). Regulating DNA replication in bacteria. Cold Spring Harb.Perspect. Biol. 5:a012922. doi: 10.1101/cshperspect.a012922

PubMed Abstract | CrossRef Full Text | Google Scholar

Torheim, N. K., Boye, E., Lobner-Olesen, A., Stokke, T., and Skarstad, K. (2000). The Escherichia coli SeqA protein destabilizes mutant DnaA204 protein. Mol. Microbiol. 37, 629–638. doi: 10.1046/j.1365-2958.2000.02031.x

PubMed Abstract | CrossRef Full Text | Google Scholar

von Freiesleben, U., Krekling, M. A., Hansen, F. G., and Lobner-Olesen, A. (2000). The eclipse period of Escherichia coli. EMBO J. 19, 6240–6248. doi: 10.1093/emboj/19.22.6240

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, S., Ma, J., Peng, J., and Xu, J. (2013). Protein structure alignment beyond spatial proximity. Sci. Rep. 3:1448. doi: 10.1038/srep01448

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, S., Peng, J., and Xu, J. (2011). Alignment of distantly related protein structures: algorithm, bound and implications to homology modeling. Bioinformatics 27, 2537–2545. doi: 10.1093/bioinformatics/btr432

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: E. coli, SeqA protein, phosphorylation, HipA kinase, initiation synchrony, minimal inter-initiation time

Citation: Riber L, Koch BM, Kruse LR, Germain E and Løbner-Olesen A (2018) HipA-Mediated Phosphorylation of SeqA Does not Affect Replication Initiation in Escherichia coli. Front. Microbiol. 9:2637. doi: 10.3389/fmicb.2018.02637

Received: 18 July 2018; Accepted: 16 October 2018;
Published: 02 November 2018.

Edited by:

Feng Gao, Tianjin University, China

Reviewed by:

Didier Mazel, Institut Pasteur, France
Jolanta Zakrzewska-Czerwinska, University of Wrocław, Poland
Dhruba Chattoraj, National Institutes of Health (NIH), United States

Copyright © 2018 Riber, Koch, Kruse, Germain and Løbner-Olesen. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Anders Løbner-Olesen, lobner@bio.ku.dk
Leise Riber, lriber@bio.ku.dk

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.