ABSTRACT
Conjugative plasmids are widespread among prokaryotes, highlighting their evolutionary success. Conjugation systems on most natural plasmids are repressed by default. The negative regulation of F-plasmid conjugation is partially mediated by the chromosomal nucleoid-structuring protein (H-NS). Recent bioinformatic analyses have revealed that plasmid-encoded H-NS homologs are widespread and exhibit high sequence diversity. However, the functional roles of most of these homologs and the selective forces driving their phylogenetic diversification remain unclear. In this study, we characterized the functionality and evolution of Sfx, a H-NS homolog encoded by the model IncX2 plasmid R6K. We demonstrate that Sfx, but not chromosomal H-NS, can repress R6K conjugation. Notably, we find evidence of positive selection acting on the ancestral Sfx lineage. Positively selected sites are located in the dimerization, oligomerization, and DNA-binding interfaces, many of which contribute to R6K repression activity—indicating that adaptive evolution drove the functional divergence of Sfx. We additionally show that Sfx can physically interact with various chromosomally encoded proteins, including H-NS, StpA, and Hha. Hha enhances the ability of Sfx to regulate R6K conjugation, suggesting that Sfx retained functionally important interactions with chromosomal silencing proteins. Surprisingly, the loss of Sfx does not negatively affect the stability or dissemination of R6K in laboratory conditions, reflecting the complexity of selective pressures favoring conjugation repression. Overall, our study sheds light on the functional and evolutionary divergence of a plasmid-borne H-NS-like protein, highlighting how these loosely specific DNA-binding proteins evolved to specifically regulate different plasmid functions.
IMPORTANCE
Conjugative plasmids play a crucial role in spreading antimicrobial resistance and virulence genes. Most natural conjugative plasmids conjugate only under specific conditions. Therefore, studying the molecular mechanisms underlying conjugation regulation is essential for understanding antimicrobial resistance and pathogen evolution. In this study, we characterized the conjugation regulation of the model IncX plasmid R6K. We discovered that Sfx, a H-NS homolog carried by the plasmid, represses conjugation. Molecular evolutionary analyses combined with gain-of-function experiments indicate that positive selection underlies the conjugation repression activity of Sfx. Additionally, we demonstrate that the loss of Sfx does not adversely affect R6K maintenance under laboratory conditions, suggesting additional selective forces favoring Sfx carriage. Overall, this work underscores the impact of protein diversification on plasmid biology, enhancing our understanding of how molecular evolution affects broader plasmid ecology.
KEYWORDS: plasmids, xenogeneic silencing, gene regulation, gene expression, conjugation, evolution
INTRODUCTION
Plasmids are extrachromosomal elements found across all domains of life (1–3) and can be transferred between even phylogenetically distant organisms (4). Conjugation is a major vehicle for plasmid spread, with roughly half or more fully sequenced plasmids being conjugative or mobilizable (5, 6). However, most natural plasmids are conjugatively repressed (7), conjugating only with specific environmental cues or in the presence of viable recipient cells (8). Previous studies have suggested that limiting conjugation can (i) offset the fitness cost of plasmid carriage by reducing energy consumption (9), (ii) minimize the stress response from the expression of conjugation-related genes (10, 11), and (iii) evade bacteriophages that attach to the conjugation pilus (12–15). Therefore, a central question in plasmid biology has been to understand the molecular mechanisms that govern conjugation repression.
Plasmid conjugation in Gram-negative bacteria depends on genes encoding for components involved in mating pair formation (Mpf) and DNA transfer and replication (Dtr) (16). The Mpf genes encode the type-IV secretion system (T4SS) responsible for plasmid transfer. At the same time, the Dtr genes encode proteins that facilitate this process by nicking and coupling the plasmid origin of transfer (oriT) to the T4SS (16). Conjugation repression can occur by regulating the expression and/or activity of Dtr and Mpf genes (17–23), and much of these mechanisms are gleaned from studying the model F plasmid. Repression of F plasmid conjugation involves chromosomal and plasmid-encoded factors that control gene transcription (24), translation (25), and protein stability (25, 26). One such chromosomal factor is the histone-like nucleoid structuring protein (H-NS). This DNA-binding protein preferably oligomerizes along curved, AT-rich DNA, causing DNA condensation and transcriptional silencing (27, 28). Within the F plasmid, H-NS binds to promoters of conjugation-related genes (PY, PM, and PJ) to repress conjugation (24, 29).
Although chromosomal H-NS regulates F plasmid conjugation, many conjugative plasmids encode their own H-NS homologs. These homologs are found across different plasmid incompatibility groups, ranging from IncA/C (30), IncH (31, 32), to IncX (33, 34), indicating a selective advantage for this alternative mode of H-NS carriage. Only a few plasmid-encoded H-NS homologs have been experimentally characterized. Sfh carried by the IncH plasmid pSfR27 is functionally redundant with chromosomal H-NS and is thought to serve as a “molecular H-NS backup (31).” Other homologs are functionally unique. Acr2 carried by the IncA/C plasmid pAR060302 specifically regulates plasmid conjugation genes without affecting most chromosomal gene expression (30). H-NSR27 encoded by the IncH R27 plasmid also represses conjugation and is uniquely resistant to negative regulation that typically antagonizes chromosomal H-NS (32). Across plasmid-encoded H-NS homologs, those carried by IncX plasmids (i.e., the HppX clade) are notable for their sequence diversity, especially within the typically conserved DNA-binding domain (35). Although some homologs within the HppX clade are known to repress conjugation (33, 34), little is known about the processes that drove their phylogenetic diversification, the molecular mechanism of conjugation repression, their interaction with chromosomal-encoded H-NS homologs, and their broader impact on plasmid ecology.
Much of our knowledge of how H-NS regulates plasmid gene expression is gleaned from studying the F plasmid and, more recently, Sfh. However, it is unclear whether these insights are broadly applicable, especially for the phylogenetically divergent HppX clade. To address this, we characterized the evolution and function of Sfx, a H-NS homolog encoded by the model IncX2 plasmid R6K. Here, we report that the evolutionary divergence of Sfx is associated with its functional divergence from chromosomal H-NS. We find that Sfx homologs are often carried by plasmids with the MPFT-type T4SS and exhibit marked amino acid differences relative to chromosomal H-NS, many of which are driven by positive selection. These evolutionary signatures are accompanied by functional divergence, whereby only Sfx, but not H-NS and StpA, can repress R6K conjugation. We find that the C-terminal DNA-binding domain of Sfx is critical for its unique conjugation repression activity. We further show that Sfx can physically interact with chromosomal silencing factors, but only its interaction with Hha contributes to conjugation repression. Additionally, we find that Sfx loss minimally impacts plasmid carrier growth rate and plasmid maintenance in laboratory culture. Collectively, our work characterizes the unique evolutionary and functional divergence of a plasmid-encoded H-NS homolog, highlighting the molecular innovation afforded by protein evolution across different genomic contexts.
RESULTS
Sfx is associated with mobile genetic elements and diverges from other H-NS homologs
IncX plasmids are characterized by a highly syntenic backbone, which includes a module where sfx is encoded immediately downstream of topB (DNA topoisomerase III) (36, 37). We first sought to further characterize the genomic context of Sfx homologs and their evolutionary relationships with other H-NS homologs. We queried the NCBI nr/nt database using three seed templates: Sfx, H-NS of Escherichia coli K12, and StpA of E. coli K12. This search identified 33, 86, and 21 representative sequences using the Sfx, H-NS, and StpA seed templates, respectively (Table S1). These sequences were then aligned and used for maximum likelihood phylogenetic inference. The resultant phylogeny is divided into three main clades. Two clades consist primarily of sequences obtained using H-NS and StpA as seed templates. The third clade is composed exclusively of sequences obtained using the Sfx seed template (Fig. 1A) and forms the sister group of the clade containing E. coli K12 StpA, consistent with previous phylogenetic analyses (35). For clarity, we designated the cluster of Sfx homologs as the Sfx clade, the group containing E. coli K12 H-NS as the H-NS clade, and the group containing E. coli K12 StpA as the StpA clade.
Fig 1.
Sfx diverges phylogenetically from chromosomal and other plasmid-encoded H-NS homologs. (A) Annotated maximum likelihood phylogeny of Sfx, H-NS, and StpA homologs. Nucleotide coding sequences of the homologs are retrieved using tblastn and an in-house script. The sequences are aligned using codon-based alignment via Guidance2, and IQ-TREE2 is used to construct a maximum likelihood phylogeny. The Ler protein carried by E. coli O157:H7 str. Sakai, a distantly related H-NS-like protein, is used as the outgroup to root the phylogeny. Each sequence representative is annotated with associated characteristics (Seed = template used in tblastn, Subfamily = taxonomic classification of the bacteria encoding for the protein, Source = whether the protein is located on a chromosome or plasmid, and Motif = amino acid sequence of the AT-hook motif). Genera that begin with “*” do not belong to the Enterobacteriaceae family. The locations of Sfx, E. coli K12 H-NS, and E. coli K12 StpA are indicated by purple, blue, or yellow diamonds, respectively. The choice of foregrounds used in the clade models is indicated and colored (StpA = yellow, Sfx = purple, and H-NS = blue). The ancestral Sfx branch set as the foreground in the Branch-site model is highlighted in green. The displayed node values represent the ultrafast bootstrap approximates (%). (B) Sfx differs from other H-NS homologs in key residue positions. Amino acid sequences of Sfx (NCBI accession ID: WP_001282381.1), Acr2 (NCBI accession ID: WP_000651490.1), StpA (Uniprot accession: P0ACG1), Sfh (Uniprot accession: Q8GKU0), and H-NS (Uniprot accession: P0ACF8) are aligned using MAFFT G-INS-i. The secondary structure of Sfx, derived using POLYVIEW-2D of a ColabFold predicted structure, is indicated above the alignment, with α representing alpha-helices, β representing beta-sheets, and empty spaces representing coils. The degree of conservation at each residue position amongst Sfx homologs is calculated based on ConSurf analysis using default parameters. For each protein, the position of the linker, a disordered region after the fourth alpha-helix, is outlined with a black border.
Sequences from the H-NS and StpA clades are all chromosomally encoded and restricted to the Enterobacteriaceae family (Fig. 1A). In contrast, sequences within the Sfx clade are either encoded by a plasmid or are found within chromosomal transposons (NCBI accession and range: CP013970.1:2715190–2715642 and CP028735.1:4153062–4153523). Consistent with its association with mobile genetic elements, Sfx homologs are found across a more extensive taxonomic distribution compared with proteins within the H-NS and StpA clades, including species from the Erwinia, Pantoea, Serratia, and Sodalis genera (Fig. 1A).
Despite its phylogenetic clustering, the Sfx and StpA clades differ in their AT-hook motif, a region that inserts into the minor groove of AT-rich DNA and is critical for DNA binding (38). The H-NS and StpA clades and previously characterized plasmid-encoded H-NS homologs, such as Acr2 and Sfh, carry a QGR AT-hook motif (Fig. 1A and B). In contrast, most Sfx homologs carry an RGR motif that is characteristic of xenogeneic silencing proteins encoded by bacteria with a higher genome GC content, including Lsr2 from Mycobacteria (39) and Bv3F from Burkholderia vietnamiensis (40). Another distinct property of Sfx is its longer and more proline-rich predicted linker (four prolines in Sfx linker vs. 1–3 prolines in other H-NS-like proteins; Fig. 1B), which could increase coil rigidity, limit the formation of larger nucleoprotein complexes, and increase target selectivity (41). Collectively, our findings indicate that Sfx homologs form a phylogenetically distinct clade associated with mobile genetic elements. This evolutionary divergence is partly driven by differences in residues that could impact protein oligomerization and DNA binding.
Sfx homologs are often carried by plasmids with the MPFT T4SS
Given the association between Sfx homologs and plasmids, we next explored its distribution across 55 representative IncX plasmid lineages (Table S2). Despite significant sequence divergence among the plasmid lineages—with no identifiable core ORFs according to Roary (42)—a little more than half of the plasmids (31/55; 56.4%) encode for H-NS homologs. Most of these plasmids carry a H-NS homolog with the RGR AT-hook motif (25/31; 80.6%) characteristic of the Sfx clade (Fig. S1). We next assessed whether plasmids carrying Sfx and H-NS homologs differ in the encoded MPF system. The T4SSs of Proteobacterial conjugative plasmids are broadly classified into four MPF classes based on protein homology: MPFF, MPFI, MPFG, and MPFT (43). IncX plasmids often carry MPFT T4SS (44), which tends to confer only the capacity to mate on solid surfaces (43). We find that IncX plasmids encoding for a H-NS homolog tend to also carry the MPFT T4SS (Fig. S1). Conversely, plasmids lacking a MPF system or possessing a MPFF T4SS (i.e., the type of T4SS possessed by the F plasmid) tend not to carry Sfx or any other H-NS homolog (Fig. S1). Taken together, our findings reaffirm the conservation of Sfx across IncX plasmids and reveal a close association of Sfx with MPFT T4SS, suggesting that Sfx may act as a regulator uniquely specific for MPFT-type conjugation systems.
Positive selection drove the evolutionary divergence of Sfx
Next, we used molecular evolution analysis to investigate the evolutionary processes that drove the divergence of the Sfx clade. We first employed clade models to determine the selective forces affecting extant sequences. Our best-performing model (CmD with Sfx and StpA clades set as separate foregrounds) suggests that most protein sites of extant H-NS, StpA, and Sfx clades show signatures of purifying selection (Table S3; Fig. S2), indicating their functional importance. Notably, the StpA clade exhibits the least degree of selective constraints (Fig. S2), which may reflect its recessive phenotype and indicate its potential for adaptive evolution.
We next used the branch-site test to determine the evolutionary processes that drove the divergence of the ancestral Sfx lineage from the StpA clade (ancestral branch highlighted in green in Fig. 1A). The branch-site test detects signatures of positive selection from a single branch selected a priori (the foreground) within a phylogeny (the background). Adaptive evolution/positive selection can be inferred if fitting an alternative branch-site model incorporating positive selection leads to a statistically significant improvement in model fit relative to fitting a null model, where only neutral and purifying evolution is considered (45). We find that fitting the alternative branch-site model leads to a statistically significant (P-value = 0.015) improvement in model fit relative to the null model (Table S3), suggesting that positive selection occurred after the Sfx clade diverged from ancestral StpA.
We next analyzed the potential functionality of positively selected residues indicated by the naive empirical Bayes (NEB) and Bayes empirical Bayes (BEB) analyses (see Table 1 for details). Although NEB outputs are more prone to false positives than BEB for small data sets (46), both methods produced similar results (Fig. 2A; Table 1). Most positively selected sites are conserved within the Sfx, H-NS, and StpA clades (Fig. S3), indicating their functional importance. Overall, positive selection altered the dimerization and oligomerization interfaces of Sfx relative to H-NS (Fig. 2B), which may change the protein’s nucleation profile and heteromeric interactions. For instance, the N9L mutation (corresponding to the L15 (Sfx)-N9 (H-NS) transition) reduces H-NS’ capacity to interact with Hha and repress the hemolysin (hly) operon (47), indicating that Sfx may display altered interaction with Hha. Although not directly part of the dimerization interface, the T27-C21 transition is biochemically similar to the C21S mutation in H-NS, also known to decrease DNA-binding affinity (48). Changes to the oligomerization domain may also impact protein regulation. H-NS is temperature-sensitive, with higher temperature destabilizing the oligomerization interface (e.g., by disrupting the ionic bond D71:R54') (49, 50), causing H-NS to adopt an inhibited configuration (49) and relieve repression (51–53). In Sfx, positive selection led to divergence in many of the sites involved in H-NS dimer-dimer interaction and H-NS auto-inhibition (e.g., S77-D71; Fig. 2B; Table 1) (49, 50).
TABLE 1.
Characteristics of positively selected residues defined based on NEB and BEB analysesa
Domain | Sfx residue | H-NS residue | NEB | BEB | Functional roles |
---|---|---|---|---|---|
N-terminal dimerization | R11 | L5 | Positively selected | Positively selected |
|
N-terminal dimerization | L15 | N9 | Positively selected | Positively selected | |
N-terminal dimerization | S19 | T13 | Positively selected | Positively selected |
|
N-terminal dimerization | F23 | Q17 | Positively selected | Positively selected |
|
N-terminal dimerization | T27* | C21 | Positively selected | Positively selected |
|
N-terminal dimerization | H28 | T22 | Positively selected | Positively selected |
|
N-terminal dimerization | Q31 | T25 | Positively selected | Negatively selected |
|
N-terminal dimerization | L33 | E27 | Positively selected | Positively selected |
|
N-terminal dimerization | Q36 | L30 | Positively selected | Negatively selected |
|
Central oligomerization | A50 | E44 | Positively selected | Negatively selected | |
Central oligomerization | A54 | A48 | Positively selected | Positively selected |
|
Central oligomerization | R57 | V51 | Positively selected | Positively selected |
|
Central oligomerization | R64 | L58 | Positively selected | Positively selected |
|
Central oligomerization | Q65 | Q59 | Positively selected | Positively selected |
|
Central oligomerization | E66 | Q60 | Positively selected | Positively selected |
|
Central oligomerization | Q72 | I66 | Positively selected | Positively selected |
|
Central oligomerization | S77* | D71 | Positively selected | Positively selected |
|
Linker | L96* | R90 | Positively selected | Positively selected |
|
Linker | K98 | Q92 | Positively selected | Positively selected |
|
Linker | A99 | R93 | Positively selected | Positively selected |
|
C-terminal DNA-binding | Y113** | T108 | Positively selected | Positively selected |
|
C-terminal DNA-binding | R117 | Q112 | Positively selected | Negatively selected | |
C-terminal DNA-binding | A120 | T115 | Positively selected | Negatively selected | |
C-terminal DNA-binding | D125 | K120 | Positively selected | Positively selected | |
C-terminal DNA-binding | K140 | K136 | Positively selected | Positively selected |
This table lists all residue positions that show evidence of positive selection within the ancestral Sfx branch, but not among the background sequences (Background ω = 0.105 and Sfx ω = 11.423) using BEB or NEB. Residue positions with a > 95% posterior probability of being positively selected via NEB are indicated by * or ** if both NEB and BEB indicate a posterior probability of being positively selected >95%. Each position in the protein sequence alignment is assigned to domains according to Zhao et al. (50), with the corresponding residues in Sfx and H-NS listed in columns 2 and 3, respectively. Functional annotations are derived from a comprehensive literature review H-NS of E. coli and S. enterica.
Fig 2.
Positively selected sites map to Sfx dimerization, oligomerization, and DNA-binding interfaces. (A) Positions of positively selected sites mapped onto an alignment of Sfx (R6K) and E. coli K12 H-NS. The identity and position of each positively selected site are displayed above (Sfx)/below (H-NS) the alignment. Each site is colored based on the posterior probability that the site has experienced positive selection following the divergence of the ancestral Sfx lineage (naive empirical Bayes results). The linker region of each protein is outlined with a black box. (B) Positively selected sites mapped onto the predicted structures of Sfx and H-NS. The top panel displays the ColabFold-predicted tetrameric structure of Sfx. Residues that have undergone positive selection are colored in magenta. The interaction interfaces of the dimerization, oligomerization, and DNA-binding domains are displayed in the inset boxes, with the top panel containing the predicted Sfx structure and the bottom panel containing the predicted H-NS structure. The N-terminal dimerization domain inset displays the interaction between a dimer (one as a grey cartoon structure and the other as a surface representation structure). The hydrophobicity of the surface representation structure is indicated by its color, and positively selected residues that form the hydrophobic interaction interface of H-NS are labeled. The central oligomerization domain inset displays the interaction between two dimers (one colored in gray and the other colored in black). The locations and charge (blue = positively charged, red = negatively charged) of positively selected residues participating in H-NS oligomerization are labeled. The apostrophe indicates that the residue is located on another protein. Although not positively selected, R60/R54 (Sfx/H-NS) is highlighted due to its ionic interaction with D71 in H-NS. The DNA-binding domain inset displays a surface representation of the linker region and the AT-hook motif (highlighted by a dashed box) colored based on its electrostatic charge (blue = positively charged, red = negatively charged). The locations of positively selected residues that participate in DNA binding in H-NS are labeled.
Among H-NS homologs, only the HppX clade significantly deviates from chromosomal H-NS at the DNA-binding domain (35). Our data suggest that positive selection drove the transitions of key residues that directly interact with DNA in H-NS (Fig. 2B; Table 1), many of which are positively charged. Interestingly, we find that positive selection underlies the unique RGR AT-hook motif of Sfx (Fig. 2B; only under positive selection by NEB but not BEB), further supporting its potential functional significance. Together, our results suggest that the ancestral Sfx lineage experienced positive selection at numerous sites, many of which correspond to residues implicated in protein-protein interaction, environmental regulation, and DNA binding.
Sfx, but not H-NS and StpA, represses R6K conjugation
We next examined whether the sequence divergence displayed by Sfx is of functional importance. To test this, we performed filter paper conjugation experiments involving E. coli carrying wild-type R6K plasmid or one lacking sfx (R6K∆sfx). Deletion of sfx increased R6K conjugation efficiency on filter paper by >1,000-fold (Fig. 3A). Considering that the filter paper is an artificial setting for conjugation, we examined whether Sfx loss would also increase R6K spread in a more naturalized biofilm setting (72). We find that R6K∆sfx conjugates at a higher efficiency than R6K in biofilms (~5-fold; Fig. 3B), affirming that Sfx is a negative regulator of plasmid transmission.
Fig 3.
Sfx represses R6K conjugation on filter paper and within biofilms. (A) sfx deletion increases R6K conjugation on filter-paper. Two OD600U of donor (E. coli K12 carrying R6K or R6K∆sfx) and recipient cells (NaN3r EcoR25) are incubated on a filter-paper disc at 30˚C for 3 h. Conjugation efficiency is measured by the number of transconjugants divided by the number of donors at the end of the conjugation period. (B) sfx deletion increases R6K conjugation in biofilm. Donor (E. coli K12 carrying R6K or R6K∆sfx) and recipient cells (NaN3r EcoR25) are seeded in M9 +0.2% glucose media to a final OD600 of 0.03 and grown at 30˚C for 24 h. Conjugation efficiency is assessed by dividing the number of transconjugant cells by the product of donor and recipient populations (to account for potential differences in recipient population growth). Conjugation efficiency is further normalized by scaling the mean conjugation efficiency of R6K∆sfx carriers to 1. (C) Sfx, but not H-NS, complementation decreases R6K∆sfx conjugation efficiency. The y-axis displays the conjugation efficiency of different donor cells (NaN3r EcoR25 recipient cells) on filter paper. The x-axis indicates the plasmids carried by the donor cells. (D) Chromosomal H-NS and StpA do not repress R6K conjugation. The y-axis displays the conjugation efficiency of different donor cells on filter paper. The x-axis indicates the donor’s genotype (top row) and plasmid (bottom row). (A–D) The data displayed are pooled results from at least two experimental replicates, and each point represents one biological replicate (averaged value across two technical replicates). The Student’s t-test is used to assess statistical significance, and the Benjamini-Hochberg correction is applied when more than one t-test is conducted. Abbreviations: ns (P-value ≥ 0.05); – (placeholder); EV (pHSG576 empty vector, low-copy); p-hns (pHSG576-hns, native hns promoter); p-sfx (pHSG576-sfx, native sfx promoter).
To ensure that the impact of Sfx loss is not polar, we transformed a separate low-copy-number plasmid expressing Sfx from its native promoter into donor cells. We find that complementation reduced the conjugation efficiency of R6K∆sfx donors to levels comparable with that of wild-type R6K donors (Fig. 3C). Interestingly, expressing chromosomal H-NS from the same low-copy plasmid backbone does not affect the conjugation efficiency of R6K∆sfx (Fig. 3C). To test whether chromosomal H-NS and its paralog StpA can also regulate R6K conjugation, we repeated the filter paper conjugation experiments using ∆hns and ∆stpA donors. We find that neither of these chromosomal deletions affects the conjugation efficiency of R6K and R6K∆sfx (Fig. 3D), indicating that chromosomal H-NS homologs do not regulate R6K conjugation. Collectively, our data suggest that Sfx represses R6K conjugation and has functionally diverged from chromosomal H-NS and StpA.
Sfx can partially compensate for H-NS loss
Plasmid-encoded H-NS homologs H-NSR27 and Sfh can partially compensate for ∆hns phenotypes (73, 74). Considering H-NS’s inability to repress R6K conjugation, we wondered whether Sfx, in contrast, can rescue the null phenotypes of E. coli∆hns. The presence of R6K or a low-copy plasmid expressing Sfx from its native promoter compensates for the reduced motility and growth lag of hns mutants (75) (Fig. 4A and C). Conversely, although R6K∆sfx carriage does not significantly impact the growth dynamics of wild-type E. coli (Fig. 4B), it reduces the growth rate of hns mutants (Fig. 4D), suggesting that at least one H-NS homolog (chromosomal H-NS or Sfx) is needed to mitigate the fitness burden of R6K carriage. Accordingly, expression of H-NS or Sfx from its native promoter of a low-copy plasmid rescued the growth defect exhibited by the hns mutant carrying R6K∆sfx (Fig. 4E). Overall, although H-NS cannot repress R6K conjugation, Sfx can partially compensate for H-NS function, suggesting a partial, albeit asymmetrical, overlap between the regulons of H-NS and Sfx.
Fig 4.
Sfx can partially compensate for ∆hns phenotypes. (A) Expressing Sfx increases the motility of hns mutants. Motility assays are performed by inoculating 6 µL of E. coli Keio Collection strains (genotype displayed below the plates) subcultured to a mid-log phase (OD600, ~0.6) into the center of a soft-agar plate (0.3% agar). The plates are incubated at 30˚C for 24 h. The images are representative of two independent experimental replicates, each with two biological replicates. (B–E) Time series growth curve data of various E. coli strains grown at 37˚C in LB or M9 + 0.2% glucose are displayed. When required, 20 µg/mL of chloramphenicol (“Cm”) is added to maintain stable inheritance of the pHSG576 low-copy plasmids. Each data point represents the mean ± standard deviation across two biological replicates, each with two technical replicates. The data are representative of three independent experimental replicates. Abbreviations: wt (wildtype E. coli BW25113); EV (pHSG576 empty vector, low-copy); p-hns (pHSG576-hns, native hns promoter); p-sfx (pHSG576-sfx, native sfx promoter).
The DNA-binding domain is necessary for Sfx-mediated conjugation repression
Given that Sfx functionally diverges from H-NS, we performed a series of gain-of-function experiments to explore which structural elements confer Sfx’s unique ability to repress R6K conjugation. We constructed chimeric proteins with different combinations of H-NS and Sfx domains (indicated in Fig. 5A), expressed them from a low-copy plasmid off the native sfx promoter, and assessed their impact on the conjugation efficiency of R6K∆sfx.
Fig 5.
Sfx C-terminus is necessary for R6K conjugation repression. (A) Schematic of domain assignment for Sfx and H-NS. Secondary structure assignments are based on the ColabFold-predicted structure of Sfx and H-NS. Coil 1 (linker) and coil 2 (contains the AT-hook motif) are two unstructured regions within the DNA-binding domain (DBD). The precise amino acid position at which the domains are split (and where the chimeras are fused) is displayed below each protein schematic. (B and C) The results from filter paper conjugation experiments performed with various donors (indicated on the x-axis) and NaN3r EcoR25 recipients are displayed. The identity of the chimeric proteins expressed from a low-copy plasmid (sfx promoter) is represented by p-NNN where each N represents the domain identities (X = Sfx, H = H NS) and a mini graphic schematic (blue = Sfx, yellow = H NS). (B) The Sfx DNA-binding domain is needed for R6K conjugation repression. The y-axis displays the conjugation efficiency (transconjugant population divided by donor population) of donors carrying R6K or R6K∆sfx and low-copy plasmids encoding for chimeric proteins. Only R6K∆sfx carriers that express chimeric proteins with the Sfx DNA-binding domain (p-NNX) exhibit statistically significantly reduced (P-value < 0.05) conjugation efficiency relative to no plasmid control (R6K∆sfx only). (C) The linker and the AT-hook motif contribute to Sfx’s conjugation repression activity. The same conjugation assay is performed as in panel B, albeit with donors expressing chimeric proteins with point or segmental mutations. Coil 1/2 indicates chimeric proteins carrying Sfx’s coil 1/2 (in either a p-XXH or p-HHH background), whereas p-sfx_QGR represents plasmid encoding for a Sfx protein with the QGR AT-hook motif. For panels B and C, statistical significance is assessed using the Student’s t-test (reference group is E. coli carrying only R6K∆sfx) with Benjamini-Hochberg correction. The data are representative results from two independent experimental replicates. Abbreviations: **** (adjusted P-value < 0.0001); *** (adjusted P-value < 0.001); ns (adjusted P-value ≥ 0.05).
All chimeras containing Sfx’s DNA-binding domain can repress conjugation of R6K∆sfx to varying degrees (Fig. 5B), suggesting that the DNA-binding domain is important for Sfx’s repression activity. The addition of Sfx’s central oligomerization domain improves chimeric protein’s repression activity (p-HXX vs. p-HHX; Fig. 5B), which could suggest that the oligomerization domain contributes to Sfx repression only when the C-terminal domain is present and/or that the presence of the domain improves protein expression/folding.
To narrow down the regions within Sfx’s DNA-binding domain that contribute to conjugation repression, we swapped smaller segments of the H-NS DNA-binding domain with corresponding segments from Sfx. Exchanging either the linker region (coil 1) or the coiled region surrounding the AT-hook motif (coil 2) conferred partial conjugation repression to constructs containing the H-NS DNA binding domain (Fig. 5B and C), suggesting that Sfx’s linker and its unique RGR AT-hook motif are major contributors to its unique repression activity. To further confirm that the functional importance of coil 2 is derived from the RGR AT-hook motif (and not the surrounding residues), we constructed Sfx mutants that carried the QGR AT hook motif and H-NS mutants that carried the RGR AT hook motif. The RGR to QGR mutation decreased but did not completely abolish Sfx’s conjugation repression activity (p-sfx_QGR; Fig. 5C), suggesting that the RGR AT-hook motif and other motifs (e.g., linker) function independently to repress R6K conjugation. Strikingly, the H-NS mutant bearing a single QGR to RGR mutation can partially repress R6K∆sfx conjugation with an efficiency similar to that of H-NS chimeras carrying the entire Sfx coil 2 (~10-fold repression, Fig. 5C). These findings indicate that the RGR AT-hook motif alone accounts for the functional importance of coil 2. Altogether, our results provide evidence that the sequence variations displayed within the DNA-binding domain of Sfx are central to its unique conjugation repression activity.
Sfx interacts with Hha to repress R6K conjugation
H-NS regulatory activity is modulated by its interactions with StpA (76–78), Hha (51, 78, 79), and Cnu/YdgT (80, 81). StpA is a paralog of H-NS that can interact with H-NS to form more thermally stable heterodimers (82) and bridged filaments (H-NS:StpA filaments that connect two DNA duplexes) (78, 83). These H-NS:StpA filaments promote greater RNA polymerase pausing (78) and could be important for gene regulation at high temperatures where H-NS repression is relieved (51, 52). H-NS also interacts with Hha and its paralog Cnu via its dimerization domain (84, 85). H-NS-Hha interaction promotes the formation of bridged filaments that stimulate greater RNA polymerase pausing (78) and is needed to repress expression of various virulence-related genes (e.g., hlyCABD operon (51, 79), LEE pathogenicity island (86, 87), Salmonella pathogenicity island 2 (88)). Although R6K does not carry a Hha homolog, certain conjugative plasmids do (89). For the IncH conjugative plasmid R27, plasmid-encoded H-NS and Hha participate in temperature-dependent conjugation repression (73). The role of Cnu is less understood (81), although recent in vitro work suggests that Cnu may be involved in osmolarity response (90).
Given the importance of protein-protein interactions for H-NS activity and the fact that plasmid-encoded H-NS, such as Sfh, is known to interact with chromosomal H-NS and StpA (91), we used a bacterial two-hybrid system to examine the physical interactions between Sfx, H-NS, StpA, Hha, and Cnu. Due to the toxicity of H-NS overexpression in the presence of ampicillin (92), we used a truncated H-NS lacking its DNA-binding domain (M1-G85) for assays involving high-copy vectors (pUT18; AmpR) encoding for H-NS.
We first validated our assay by recovering previously characterized interactions, including those between H-NS and StpA (93), H-NS and Hha (94), H-NS and Cnu (80, 95), and StpA and Cnu (95) (Fig. 6A). We find that Sfx interacts with H-NS, StpA, and Hha, whereas Sfx-Cnu interaction is only detected when Cnu is expressed from a high-copy vector (pUT18/pUT18C, Fig. 6A and B). Interestingly, a fusion of the reporter fragment to the N- or C-terminus of Hha did not affect Hha-Sfx interaction, unlike Hha-H-NS interaction, which is weakened by Hha C-terminal fusion (Fig. 6A) (94). This discrepancy in Hha sensitivity to terminal fusions indicates potential differences between Sfx-Hha and H-NS-Hha interactions, which corroborates with the fact that key residues involved in H-NS-Hha interactions underwent positive selection in the ancestral Sfx lineage (e.g., L15-N9 transition).
Fig 6.
Sfx-Hha interaction contributes to conjugation repression. (A) Sfx interacts with H-NS, StpA, and Hha. Bacterial two-hybrid assay is used to assess binary protein-protein interactions between H-NS, Sfx, StpA, Hha, and Cnu. A blue spot indicates that the expressed fusion proteins interact with sufficient strength to reconstitute the split adenylate cyclase, leading to the expression of a cAMP-CAP-regulated lacI. A white spot indicates either insufficient interaction between the two Fusion proteins or insufficient expression. Fusion proteins are either made by fusing the adenylate cyclase fragment to the C-terminus (CT) or N-terminus (NT). We expressed the fusion proteins either from a high-copy (pUT18/pUT18C) or a low-copy (pKT25/pKNT25) plasmid. H-NS (M1-G85) lacking the DNA-binding domain is used in lieu of the full-length protein due to the toxicity of H-NS expression from a high-copy vector in the presence of ampicillin. Sfx (∆NT) represents an N-terminally truncated variant (∆D3-E8). Each assay is performed with a positive control (PC; E. coli carrying pKNT25-hns and pUT18-stpA, two proteins that are known to interact) and a negative control (E. coli carrying the plasmid indicated on the top row and a pKNT25/pKT25 empty vector). The assay is performed with three biological replicates, each with two technical replicates. The presented images are representative of two independent experiments. (B) Network depiction of bacterial two-hybrid results. The solid lines suggest that protein interaction is detected for both protein construct combinations (e.g., pUT18-hns and pKNT25-stpA, pUT18-stpA, and pKNT25-hns). The dotted lines suggest that protein interaction is detected only with one of the protein construct combinations. (C) N-terminally truncated Sfx cannot repress R6K conjugation. Various constructs (“–” = no plasmid, EV = empty vector, p-sfx = pHSG576 sfx, p-sfx (∆NT) = pHSG576 sfx (∆E3-D8)) are transformed into E. coli carrying R6K∆sfx and tested for their ability to repress conjugation efficiency on filter paper. (D) Hha co-represses R6K conjugation with Sfx. Filter paper conjugation is performed with various E. coli donors from the Keio collection. The genotypes of the donors are indicated by the top row in the x-axis, and the plasmids they carry are displayed in the bottom row. Complementation of hha deletion is performed by transforming R6K carriers with EV (pHSG576 empty vector) or p-hha (pHSG576-hha with native hha promoter, low-copy). For panels C and D, the data shown are representative of two independent experiments. When indicated, the Student’s t-test with Benjamini-Hochberg correction is used to assess statistical significance. Abbreviations: *** (adjusted P-value < 0.001); ** (adjusted P-value < 0.01); ns (adjusted P-value ≥ 0.05).
Interestingly, Sfx that lacks its N-terminal extension (∆D3-E8/∆NT) cannot form heteromeric interactions but can still interact with wild-type Sfx and N-terminally truncated Sfx (Fig. 6A). N-terminally truncated Sfx cannot repress R6K∆sfx conjugation when expressed from a low-copy plasmid (Fig. 6C), suggesting that heteromeric interactions are needed for Sfx-mediated conjugation repression and/or that the N-terminal extension is necessary for proper Sfx-Sfx interaction. To explore the former possibility, we assessed whether Sfx’s interactions with H-NS, StpA, Hha, and Cnu are necessary for its conjugation repression activity. To this end, we performed filter paper conjugation using R6K/R6K∆sfx-carrying donors with hns, stpA, hha, or cnu deletions (gene deletions verified by lack of PCR amplification). We find that hns, stpA, and cnu deletions do not affect the conjugation efficiency of R6K and R6K∆sfx (Fig. 6D). In contrast, hha deletion increases the rate of R6K conjugation by ~10-fold, and this elevated conjugation efficiency is decreased by the complementation of Hha encoded on a low-copy vector (Fig. 6D). Moreover, hha deletion does not elevate the conjugation rate of R6K∆sfx, indicating that Hha repression of R6K conjugation is dependent on Sfx (Fig. 6D). This result and our bacterial two-hybrid findings indicate that Sfx may interact with Hha to repress R6K conjugation.
Sfx loss does not affect R6K fitness in laboratory settings
Conjugation is an energetically expensive process that can sometimes decrease the fitness of plasmid carriers (96, 97). We have so far been unable to detect a fitness cost (i.e., growth defect in wild-type E. coli) associated with carrying R6K∆sfx in liquid culture (Fig. 4B), which is inconsistent with the expected tradeoff between conjugation and growth (98) and the widespread prevalence of Sfx homologs across MPFT-type IncX plasmids (Fig. S1). Therefore, we next investigated whether this lack of fitness cost of Sfx loss also applies in biofilms, a conjugative-permissive environment where plasmid fitness is dictated by donor (vertical inheritance) and transconjugant (horizontal inheritance) population dynamics. When co-cultured with EcoR25 recipient cells, we find that R6K and R6K∆sfx donors reached a similar population density in the biofilm and planktonic states (Fig. S4A). This result is affirmed by an R6K/R6K∆sfx competition assay, where both plasmid carriers reached a similar population density in the biofilm and planktonic states (Fig. S4B). Overall, we cannot detect a fitness cost to Sfx loss in a laboratory biofilm setting.
We next investigated whether Sfx loss would incur a fitness cost on a longer timescale. To test this, we serially passaged six independent strains of E. coli BW25113 carrying R6K or R6K∆sfx in LB or minimal M9 + 0.2% glucose media for 20 days (~200 generations). Under both media conditions, all lineages stably maintained R6K and R6K∆sfx (Fig. 7A), suggesting that Sfx loss does not impact plasmid stability within the study period, consistent with previous studies of an IncX3 plasmid (34). To investigate if compensatory mutations on the plasmids could account for this stability, we performed whole plasmid sequencing of the ancestor and two independent evolved lineages for each plasmid and passaging condition (a total of four evolved R6K and four R6K∆sfx sequenced). No mutations were detected on the plasmids after 20 days of serial passaging. To investigate whether the passaging selected for chromosomal mutations that affected plasmid conjugation efficiency, we selected two colonies (R6K∆sfx carriers) from each lineage and passaging condition and compared their conjugation rate relative with their ancestors. We find that passaging in LB and M9 media did not impact the conjugation efficiency of R6K∆sfx carriers (one lineage passaged in LB displayed slightly higher conjugation efficiency; Fig. 7B), suggesting that short-term passaging did not select for lower conjugation efficiency. This data is consistent with our inability to detect a fitness cost associated with Sfx loss.
Fig 7.
Sfx loss does not affect R6K stability. (A) R6K and R6K∆sfx are stably maintained throughout serial passaging in LB and M9 media. For each passaging experiment, six independent lineages are diluted 1:1,000 in fresh media without antibiotics every 24 h. The proportion of plasmid carriers is assessed by dividing the number of streptomycin-resistant colonies by the total colony count. Each point represents the mean proportion of plasmid carriers, and the lines delineate the mean ± standard deviation. (B) Passaging in LB or M9 does not decrease donor conjugation efficiency. The filter paper conjugation experiment is performed using the ancestral (day 0) and evolved (day 20) strains from the passaging experiment shown in panel A. The relative conjugation efficiency of the evolved-to-ancestor strain is plotted on the y-axis. The different shapes indicate the lineage of each sample, and each point represents the value of a biological replicate (averaged across two technical replicates).
DISCUSSION
Plasmid conjugation is ubiquitous in natural bacterial communities and is crucial in driving bacterial evolution (1, 99–101). Interestingly, conjugation is repressed by default across most natural conjugative plasmids (7), indicating widespread selective pressures that drove the convergent evolution of repression systems. Understanding the molecular mechanism and evolution of conjugation repression is of major interest in plasmid biology. Here, we report that the evolutionary divergence of Sfx, a plasmid-encoded H-NS homolog, contributes to the unique regulation of IncX plasmid conjugation.
Gene duplication has long been recognized as a major driver for genome evolution (102, 103). Duplicated genes, known as paralogs, provide functional redundancy that temporarily relaxes purifying selection, allowing for sequence divergence that leads to gene pseudogenization (becoming non-functional), conservation (maintaining the same function), subfunctionalization (performing a subset of roles of the ancestral gene), or neofunctionalization (performing a new function) (103). Our results suggest that this duplication-divergence framework may explain the evolution of Sfx and the unique regulatory system of some IncX plasmids.
First, our phylogenetic analysis indicates that Sfx homologs form the sister clade of the H-NS paralog StpA (Fig. 1A), consistent with prior work (35). This relationship suggests that ancestral Sfx may also be a H-NS paralog and likely experienced reduced purifying selection following the duplication. Indeed, our molecular evolution analysis suggests that the extant StpA clade experiences more relaxed selective constraints relative to other H-NS homologs (Fig. S2), presumably due to the lack of a pronounced fitness cost associated with stpA null mutations (77, 104). We propose that this recessive phenotype of StpA, and likely other H-NS paralogs, potentiates the evolution of functionally distinct H-NS proteins.
Both neutral evolution (105) and positive selection (106–108) are known to drive sequence divergence following gene duplication. What drove the divergence of ancestral Sfx following its duplication? Our evolutionary analyses suggest that positive selection contributed to the initial burst of evolution experienced by the ancestral Sfx lineage (Fig. 2). However, neutral evolution dominated in the clade’s later evolution (Table S3), implying the loss of functional redundancy following lineage divergence. We propose that this later neutral evolution of the Sfx clade may be facilitated by the proliferation of homologs across IncX plasmids, particularly those with the MPFT-type T4SS, with plasmid incompatibility curbing gene recombination and a higher gene copy number potentially boosting mutation rates (109, 110).
Having established the mechanism of Sfx clade evolution, we next explored the functional consequences of these mutations. Indeed, many of the positively selected sites we document align with H-NS regions important for dimerization, oligomerization, interaction with Hha, DNA binding, and environmental sensing (Table 1). Some of these variations underpin the functional divergence between Sfx and chromosomal H-NS. Notably, Sfx, but not H-NS nor StpA, can repress conjugation of R6K (Fig. 3D), a feature that likely favored the migration/persistence of ancestral Sfx onto/on IncX plasmids. The C-terminal DNA-binding domain of Sfx is necessary for conjugation repression. Additionally, Sfx-mediated repression is enhanced by its interaction with chromosomal Hha and requires Sfx’s N-terminal dimerization domain (Fig. 6A). This co-repression may involve Hha’s ability to promote DNA bridging and generate topological stress (78, 111), inhibiting conjugation gene transcription and/or prohibiting the initial unwinding of the donor strand needed for conjugative transfer. Conversely, plasmid-encoded topoisomerase III (topB), which frequently colocalizes with sfx (36, 37), could relieve this topological stress and promote conjugation. Given that similar Hha-H-NS co-repression is observed in the IncHI plasmid R27 (73) and that Hha homologs are often co-localized with H-NS homologs in plasmids (89), conjugation regulation by H-NS/Hha pairs may be a common mechanism.
After demonstrating the functional importance of Sfx’s C-terminal DNA-binding domain, we further explored the biochemical basis of its significance. Our gain-of-function experiments indicate that the linker region is important for Sfx’s conjugation repression activity (Fig. 5C). Positive selection within the linker region notably redistributed its positive charge (Fig. 2B). Although the functional impact of this charge variation is unclear—since similar changes in chromosomal H-NS did not significantly impact its regulatory activity (64)—it is plausible that the distinct charge distribution, longer length, and increased proline-induced rigidity (41) of the Sfx linker could collectively contribute to its unique repression activity.
The RGR AT-hook motif is another distinct feature of the Sfx clade (Fig. 1A) and contributes to the regulatory activity of Sfx’s C-terminal DNA binding domain (Fig. 5C). The first position of this motif displays signatures of positive selection (Fig. 2A). Remarkably, altering just one amino acid from QGR to RGR in chromosomal H-NS enables it to partially repress R6K conjugation (Fig. 5C), demonstrating that small sequence changes can confer large functional shifts. Interestingly, the RGR AT-hook motif is typically found in distantly related H-NS-like proteins encoded by bacteria with GC-rich genomes. These include Lsr2 carried by Mycobacteria (62-70% GC) (112) and Bv3F carried by Burkholderia (40) (59.2%–68.9% GC; from NCBI genome). These H-NS-like proteins show a heightened affinity for higher GC% sequences than those with the QGR motif (40, 66), enabling them to better differentiate between core and foreign genes in high GC% genomic contexts (66, 113).
Curiously, Sfx homologs are not found in high GC% genomic contexts (Fig. S5B). For instance, R6K has a GC% of 45.3%, slightly lower than E. coli BW25113 (50.8%; NCBI accession: CP009273.1). Furthermore, IncX plasmids carrying Sfx homologs display an unusual base composition: their insertion elements (foreign cargo genes) are higher in GC% than their core genes (median GC% difference of 7.4%; Fig. S5A and B), unlike previously characterized plasmids (114, 115). We hypothesize that Sfx’s RGR AT-hook motif and the plasmid’s atypical GC composition may facilitate better regulation of plasmid conjugation. First, the transfer operon of R6K is less AT-rich than typical H-NS binding sites (56.6% AT from taxA to tiv11 vs. 61.4% AT for average H-NS-bound site [116]). Therefore, the RGR motif may enable Sfx to bind and repress transcription more effectively in these regions, where the higher GC content prohibits efficient regulation by chromosomal H-NS. This bias toward Sfx-mediated regulation, rather than H-NS-mediated, could confer several selective advantages, including optimized timing of conjugation independent of H-NS activity and adaptability to hosts without H-NS homologs. The elevated GC% of insertion elements also suggests potential decoupling of insertion element regulation from conjugation, enabling appropriate cargo gene expression that enhances plasmid carrier fitness. Overall, Sfx represents a unique instance of an RGR-bearing H-NS homolog found in a lower GC% genomic context. Future experiments examining the exact mechanism of how the RGR AT-hook motif facilitates conjugation regulation could yield invaluable insights.
Conjugation repression is common among characterized plasmids (7), and the fitness cost of constitutive conjugation is well-documented in numerous plasmid systems (9, 10, 96–98, 117). However, our results did not reveal any apparent fitness penalty associated with Sfx loss. E. coli carrying R6K or R6K∆sfx display similar growth dynamics in nutrient-rich (LB) and -poor (M9) environments (Fig. 4B through E; Fig. S4), and R6K∆sfx carriers exhibit growth defects only under highly artificial conditions in which chromosomal H-NS is also lost (Fig. 4D). This lack of fitness cost is also found on a larger time scale, wherein R6K∆sfx is stably maintained in the absence of antibiotic selection, consistent with previous studies on IncX3 plasmids (34), without selecting for mutations affecting conjugation efficiency (Fig. 7). Our findings suggest that conjugatively derepressed plasmids may have a fitness advantage under certain conditions. Indeed, plasmids harboring “superspreader” mutations that increase conjugation efficiency have been isolated from clinical and environmental sources (118, 119). Our findings also imply that other selective forces, not measured in laboratory conditions, may impose a selective advantage to keeping plasmid conjugation rates low, such as susceptibility to bacteriophages that attach to conjugal pili (12–15).
In summary, our study provides insights into the molecular evolution and functionality of plasmid-encoded H-NS homologs. We demonstrate the evolutionary and functional divergence of Sfx from chromosomal H-NS and its unique role in plasmid conjugation repression. The altered protein-protein and protein-DNA interaction interfaces of Sfx, driven partially by positive selection, likely contribute to its distinct functionality. Despite the conservation of Sfx among IncX plasmids, its loss does not impose an obvious fitness cost, reflecting the complexity of selection forces shaping plasmid evolution. Our results, therefore, highlight the sequence and functional diversity within the H-NS family and underscore the pivotal role that protein evolution plays in plasmid biology.
MATERIALS AND METHODS
Strains and plasmids
The strains, plasmids, and oligonucleotides used in this study are listed in Table S4 through S6, respectively. E. coli DH5α (NEB) was used for cloning and plasmid propagation. Except for the bacterial two-hybrid assays, all experiments were performed using the E. coli BW25113 or strains from the Keio collection when specified (120). For all conjugation assays, the recipient strain was a spontaneous NaN3r EcoR25 strain derived in-house. E. coli BTH101 (F−, cya-99, araD139, galE15, galK16, rpsL1 [Strr], hsdR2, mcrA1, mcrB1) and T18/T25-fragment containing plasmid backbones (pKT25, pKNT25, pUT18, pUT18C) are a kind gift from Véronique Taylor (University of Toronto) and were used for the bacterial two-hybrid assays. The domain-swapping, complementation, and bacterial two-hybrid plasmids were constructed via Gibson assembly using PCR-amplified vector backbone and fragments from R6K or E. coli BW25113 genomic DNA following the manufacturer’s instructions. Mutations and small fragment substitutions in plasmids were introduced via site-directed mutagenesis using a homebrew KLD enzyme mix. Plasmids were transformed into E. coli using electroporation or heat shock. R6K∆sfx is a kind gift from Irina Artsimovitch (Ohio State University). All plasmid constructs were verified via Sanger or Nanopore sequencing (Plasmidsaurus).
Media and culture conditions
E. coli strains were routinely grown in LB liquid media (BD Difco LB Broth Miller) or LB agar (BD Difco LB Agar Miller) at 37˚C. Liquid cultures were grown with shaking (200 rpm). When required for plasmid maintenance, the growth medium was supplemented with 100 µg/mL streptomycin, 50 µg/mL kanamycin, 100 µg/mL ampicillin, 20 µg/mL chloramphenicol, and/or 0.01% NaN3. For growth curve, biofilm, and serial passaging, strains were grown in M9 (BD Difco Bacto M9 Minimal Salts) + 0.2% glucose when specified.
Constructing phylogeny of Sfx, H-NS, and StpA
A Sfx, H-NS, and StpA phylogeny was inferred via the maximum-likelihood method using IQ-TREE2 (121). The best codon substitution model was chosen using ModelFinder (122), and branch confidence was assessed via 5,000 rounds of Ultrafast Bootstrap Approximation (123) and 1,000 bootstrap replicates of SH-aLRT (124). For visualization purposes, the tree was rooted at the Ler protein, but all molecular evolution analyses were performed on an unrooted phylogeny that does not contain the Ler protein sequence. Sequence taxonomy was derived based on the subfamily classification proposed by Alnajar and Gupta (125), whereas other sequence characteristics (genetic context, AT-hook motif) were retrieved from NCBI or derived from the protein alignment and mapped onto the phylogeny using the ggtree package (126).
Conservation and alignment of chromosomal and plasmid-borne H-NS homologs
We aligned the amino acid sequences of Sfx (NCBI accession ID: WP_001282381.1), Acr2 (NCBI accession ID: WP_000651490.1), StpA (Uniprot accession: P0ACG1), Sfh (Uniprot accession: Q8GKU0), and H-NS (Uniprot accession: P0ACF8) using MAFFT G-INS-i (127). The conservation within Sfx homologs was derived using the ConSurf webserver (128) with default parameters.
Mapping Sfx orthologs across IncX plasmids
We retrieved 975 IncX plasmids, as classified by PlasmidFinder (129), from PLSDB (130) in March of 2023. One plasmid was arbitrarily selected from each unique PlasmidFinder lineage to produce 55 representative sequences. For each lineage representative, we reannotated the sequences using PGAP (131) and derived plasmid attributes (replicon, relaxase, MPF, OriT, predicted mobility, predicted host range) using MOB-suite (132). To identify H-NS orthologs and co-occurring proteins, we performed orthologous clustering of all 55 lineages using OrthoFinder with default parameters (133). The AT-hook motif of the H-NS homologs was derived by performing multiple sequence alignment using the DECIPHER package (134) and further validated by visual inspection. The attributes of the representative plasmids were clustered via hierarchical clustering using Gower distance and visualized using the R package ggtree (126).
Generating Sfx, H-NS, and StpA sequence alignment for molecular evolution analysis
The NCBI nr/nt database (135) is queried using tblastn in October 2022 with default search parameters to retrieve nucleotide sequence homologous to Sfx from the R6K plasmid (NCBI accession ID: WP_001282381.1), H-NS from E. coli K12 (Uniprot ID: P0ACF8), and StpA from E. coli K12 (Uniprot ID: P0ACG1). We removed all hits that are <50% similar to the seed template sequences and retrieved the coding sequences corresponding to the remaining hits using an in-house script. The resulting sequences were filtered to eliminate short (<88 amino acids for Sfx homologs, <100 amino acids for H-NS and StpA homologs) and overly long peptides (>200 amino acids for all sequences) that are indicative of pseudogenes and/or misannotation. The nucleotide sequences were clustered using MMseqs2 (136) to identify sequence-level representatives (95% identity, 80% coverage, coverage mode = 1). The sequence representatives were aligned using the Guidance2 web server (137) (codon model, MAFFT with maximum 1,000 cycles of iterations) and trimmed to eliminate any sequences with <0.6 confidence score and columns with <0.87 confidence score. To aid with phylogeny visualization, the Ler protein from E. coli O157:H7 str. Sakai (GenBank accession: BAB38011.2), a H-NS-like protein that is distantly related to Sfx, H-NS, and StpA (138) (pairwise sequence identity ~20%–25%) was chosen as the outgroup and aligned with all sequences.
Molecular evolution analysis using PAML
We used PAML 4.10.6 (139) to explore the variation in selective constraints across Sfx, H-NS, and StpA homologs. Branch-sites (140) and clade models (141) were fitted using the codeml program, and likelihood ratio tests were used to explore whether using more complex models leads to a statistically significant improvement in model fit (142). We used the Akaike information criterion (AIC) as an evaluation metric for comparison across non-nested models, as suggested by Weadick and Chang (142). Residues under positive selection were detected using NEB and BEB analyses (46).
AlphaFold structural prediction of Sfx, H-NS, and StpA
Tetramers and dimers of Sfx, H-NS, and StpA were predicted using ColabFold (Alphafold2-multimer) (143). Template information of each sequence set was fetched from the PDB70 database, and the predicted structures were relaxed using the amber force fields. The secondary structure was derived using POLYVIEW-2D (144) and visualized using the R package gggenes (145), whereas the 3D structure was modeled in PyMOL (146).
Constructing WebLogo of Sfx, H-NS, and StpA homologs
Sequences within the Sfx, H-NS, and StpA clades (defined based on phylogeny) were aligned (within-clade) using the MAFFT E-INS-i algorithm (127). The sequence alignments were manually trimmed using Pfaat to remove gappy regions (147), and the amino acid frequency at each position was visualized using the WebLogo webserver (148).
Filter paper conjugation assay
For all experiments, one biological replicate of the recipient strain and two biological replicates of the donor strains were grown overnight in LB (recipient: supplemented with 0.01% NaN3; donor: supplemented with 100 µg/mL of streptomycin, and 20 µg/mL chloramphenicol if harboring pHSG576 constructs) at 37˚C with shaking (200 rpm). The conjugation efficiency on filter paper was assessed as follows: 2 OD600 U of donor and recipient (spontaneous NaN3r EcoR25) overnight cultures were pelleted (4,500 × g for 5 min at room temperature), washed with sterile phosphate-buffered saline (PBS), and combined at equal volume. In addition, 20 µL of the culture mixture was spotted on MF-Millipore Membrane Filter (0.22 µm pore size) overlaid on LB agar and incubated at 30°C for 3 h. The bacterial spots were washed off by vortexing the filter paper discs in 100 µL of sterile PBS for 20 s. The donor and transconjugant populations were determined by spotting serially diluted bacterial mixtures on LB agar supplemented with 100 µg/mL ampicillin and LB agar supplemented with 100 µg/mL ampicillin and 0.01% NaN3, respectively. Each conjugation experiment was repeated at least twice.
Biofilm conjugation assay
Donor and recipient strains were grown and processed via the same method outlined in “Filter paper conjugation assay.” For each strain, washed overnights of three biological replicates were diluted to a final OD600 of 0.03 in 150 µL of M9 + 0.2% glucose media in a 96-well polystyrene plate (Sarstedt). The similarity in donor cell input (for conjugation assays) or R6K/R6K∆sfx carrier (for biofilm competition assay) was validated by plating serially diluted inoculum on LB agar supplemented with 100 µg/mL streptomycin (which selects for both R6K and R6K∆sfx carriers) and/or 50 µg/mL kanamycin (which selects for only R6K∆sfx carriers). Biofilm formation was allowed to occur statically at 30˚C for 24 h. After 24 h, the planktonic cultures were aspirated and quantified via selective plating (see below for conditions). The biofilm was gently washed once with 150 µL of sterile PBS to remove non-adherent cells and resuspended in 40 µL of sterile PBS. To assess donor, recipient, and transconjugant population, the planktonic cultures/resuspended biofilms were serial diluted and spotted on LB agar supplemented with 100 µg/mL ampicillin, or 0.01% NaN3, or 100 µg/mL ampicillin and 0.01% NaN3, respectively. Each biofilm conjugation experiment was repeated at least twice.
Motility assay
Overnight E. coli cultures grown at 37˚C with shaking were subcultured 1:100× in 5 mL of LB with no antibiotics to an OD600 of ~0.6. Six microliters of mid-log culture was aspirated into the middle of 0.3% LB agar, dried at room temperature for 1 h, and incubated at 30˚C for 24 h. The motility assay was repeated twice.
Growth curve analysis
One OD600 unit of overnight E. coli cultures were pelleted (4,500 × g for 5 min at room temperature), washed with sterile PBS, and diluted to a final OD600 of 0.02 in 200 µL of growth media (LB or M9 + 0.2% glucose supplemented with 20 µg/ml chloramphenicol if strains carry pHSG576 constructs). All cells were seeded in 96-well polystyrene plates (Sarstedt), sealed with Breathe-Easy sealing membrane (Sigma Aldrich), and grown at 37˚C for 24 h with shaking every 15 min. OD620 was monitored every 15 min by an S&P growth robot. All growth curve assays were repeated at least three times.
Bacterial two-hybrid assay
The coding sequences of hns, stpA, hha, and cnu were amplified via PCR from E. coli BW25113 genomic DNA, whereas the coding sequences of sfx was amplified from R6K. The PCR products were cloned into bacterial two-hybrid vectors (pUT18, pUT18C, pKT25, pKNT25) using Gibson assembly following the manufacturer’s instructions and validated by Sanger sequencing. All H-NS homolog constructs (Sfx, H-NS, and StpA) were fused to the adenylate cyclase fragment at its C-terminus to preserve the N-terminal dimerization domain. High-copy vectors (pUT18C) encoding for Hha and Cnu are constructed with only the N-terminal T18 fusion (given that C-terminal fusion disrupts H-NS-Hha interaction [94]), whereas both N-terminal and C-terminal chimeric proteins were constructed for low-copy vectors. To detect protein-protein interactions, we heat-shocked the bacterial two-hybrid vectors into E. coli BTH101 and plated them on LB supplemented with 100 µg/mL ampicillin and 50 µg/mL kanamycin to select for transformants. Three biological replicates of each strain were grown overnight in LB supplemented with 100 µg/mL ampicillin and 50 µg/mL kanamycin at 37°C with shaking (200 rpm). Expression of the chimeric proteins was then induced by spotting 5 µL of overnight cultures on LB agar supplemented with 100 µg/mL ampicillin, 50 µg/mL kanamycin, 40 µg/mL X-gal, and 50 mM isopropyl β-D-1-thiogalactopyranoside (IPTG). The plates are incubated at 30°C for 24 h (for strains carrying pUT18::hns (M1-G85), pUT18::stpA, pUT18::sfx, pUT18::sfx-N-truncated) or 48 h (for strains carrying pUT18C::hha and pUT18C::cnu). Each bacterial two-hybrid screen was repeated twice.
Serial passaging and assessing passaged strain conjugation efficiency
Six independent lineages of E. coli carrying R6K or R6K∆sfx were serially passaged in LB or M9 + 0.2% glucose for 20 days (~200 generations). All strains were grown in 5 mL of media in 20 mm glass tubes at 37˚C with shaking (200 rpm). Every 24 h, the cultures were diluted 1:1,000× in fresh media. The proportion of plasmid-retaining cells was assessed by plating serially diluted cultures on LB agar (quantify total bacterial population) and LB agar supplemented with 100 µg/mL streptomycin (which selects for only plasmid carriers). To assess whether mutations accumulated on the R6K/R6K∆sfx plasmid during passaging, single colonies of passaged strain (from frozen DMSO stocks) were grown overnight in LB supplemented with 100 µg/mL streptomycin. Plasmids were harvested using the GeneJET Plasmid Miniprep Kit following the manufacturer’s protocol and sequenced using Nanopore sequencing (Plasmidsaurus). The effect of serial passaging on plasmid conjugation was assessed by taking single colonies from day 0 (ancestral strain) and day 20 (passaged strain) DMSO stocks preserved at −80˚C and evaluating the filter paper conjugation efficiency of two biological replicates, each with two technical replicates (filter paper conjugation experiment was performed as specified above).
GC content calculation of IS and non-IS regions in IncX plasmids
We used the ISfinder algorithm (blastn, default parameters) (149) to search for putative IS regions across 22 IncX plasmid representatives with H-NS homologs. The resulting hits were filtered to retain only higher confidence (e-value ≤ 0.1) and longer sequences (>100 bp). An in-house script was then used to calculate the GC% of all putative IS and non-IS regions.
ACKNOWLEDGMENTS
We thank members of the Navarre Lab, Jordan Lin (University of Toronto), Dr. Irina Artsimovitch (Ohio State University), Dr. Barbara Funnell (University of Toronto), and Dr. Kamna Singh for their feedback on the manuscript and technical support. We would also like to thank Dr. Véronique Taylor (University of Toronto) for her help in setting up the bacterial two-hybrid assay and sharing the strains and plasmids. Finally, we thank Dr. Belinda Chang (University of Toronto) for her guidance and critical feedback on the molecular evolution analysis.
This work was funded by grants from the Natural Sciences and Engineering Research Council of Canada (NSERC: RGPIN-2020-06015 and USRA 24/25).
Contributor Information
William Wiley Navarre, Email: william.navarre@utoronto.ca.
Kelly T. Hughes, University of Utah, Salt Lake City, Utah, USA
SUPPLEMENTAL MATERIAL
The following material is available online at https://doi.org/10.1128/mbio.02089-24.
Tables S1 to S6.
Figures S1 to S5.
ASM does not own the copyrights to Supplemental Material that may be linked to, or accessed through, an article. The authors have granted ASM a non-exclusive, world-wide license to publish the Supplemental Material files. Please contact the corresponding author directly for reuse.
REFERENCES
- 1. Rodríguez-Beltrán J, DelaFuente J, León-Sampedro R, MacLean RC, San Millán Á. 2021. Beyond horizontal gene transfer: the role of plasmids in bacterial evolution. Nat Rev Microbiol 19:347–359. doi: 10.1038/s41579-020-00497-1 [DOI] [PubMed] [Google Scholar]
- 2. Wang H, Peng N, Shah SA, Huang L, She Q. 2015. Archaeal extrachromosomal genetic elements. Microbiol Mol Biol Rev 79:117–152. doi: 10.1128/MMBR.00042-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Extrachromosomal elements in lower eukaryotes. 1986. Springer US, Boston, MA. Available from: http://link.springer.com/10.1007/978-1-4684-5251-8 [Google Scholar]
- 4. MacLean RC, San Millan A. 2015. Microbial evolution: towards resolving the plasmid paradox. Curr Biol 25:R764–R767. doi: 10.1016/j.cub.2015.07.006 [DOI] [PubMed] [Google Scholar]
- 5. Coluzzi C, Garcillán-Barcia MP, de la Cruz F, Rocha EPC. 2022. Evolution of plasmid mobility: origin and fate of conjugative and nonconjugative plasmids. Mol Biol Evol 39:msac115. doi: 10.1093/molbev/msac115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6. Ares-Arroyo M, Nucci A, Rocha EPC. 2024. Identification of novel origins of transfer across bacterial plasmids. bioRxiv. doi: 10.1101/2024.01.30.577996 [DOI] [Google Scholar]
- 7. Ghigo JM. 2001. Natural conjugative plasmids induce bacterial biofilm development.. Nature New Biol 412:442–445. doi: 10.1038/35086581 [DOI] [PubMed] [Google Scholar]
- 8. Singh PK, Meijer WJJ. 2014. Diverse regulatory circuits for transfer of conjugative elements. FEMS Microbiol Lett 358:119–128. doi: 10.1111/1574-6968.12526 [DOI] [PubMed] [Google Scholar]
- 9. San Millan A, MacLean RC. 2017. Fitness costs of plasmids: a limit to plasmid transmission. Microbiol Spectr 5. doi: 10.1128/microbiolspec.MTBP-0016-2017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10. Zahrl D, Wagner M, Bischof K, Koraimann G. 2006. Expression and assembly of a functional type IV secretion system elicit extracytoplasmic and cytoplasmic stress responses in Escherichia coli. J Bacteriol 188:6611–6621. doi: 10.1128/JB.00632-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11. Yang X, Ma Q, Wood TK. 2008. The R1 conjugative plasmid increases Escherichia coli biofilm formation through an envelope stress response. Appl Environ Microbiol 74:2690–2699. doi: 10.1128/AEM.02809-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Bradley DE, Coetzee JN, Bothma T, Hedges RW. 1981. Phage X: A plasmid-dependent, broad host range, filamentous bacterial virus. J Gen Microbiol 126:389–396. doi: 10.1099/00221287-126-2-389 [DOI] [PubMed] [Google Scholar]
- 13. Jalasvuori M, Friman V-P, Nieminen A, Bamford JKH, Buckling A. 2011. Bacteriophage selection against a plasmid-encoded sex apparatus leads to the loss of antibiotic-resistance plasmids. Biol Lett 7:902–905. doi: 10.1098/rsbl.2011.0384 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Ojala V, Laitalainen J, Jalasvuori M. 2013. Fight evolution with evolution: plasmid-dependent phages with a wide host range prevent the spread of antibiotic resistance. Evol Appl 6:925–932. doi: 10.1111/eva.12076 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Quinones-Olvera N, Owen SV, McCully LM, Marin MG, Rand EA, Fan AC, Martins Dosumu OJ, Paul K, Sanchez Castaño CE, Petherbridge R, Paull JS, Baym M. 2023. Diverse and abundant phages exploit conjugative plasmids. bioRxiv:2023.03.19.532758. doi: 10.1101/2023.03.19.532758 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. de la Cruz F, Frost LS, Meyer RJ, Zechner EL. 2010. Conjugative DNA metabolism in Gram-negative bacteria. FEMS Microbiol Rev 34:18–40. doi: 10.1111/j.1574-6976.2009.00195.x [DOI] [PubMed] [Google Scholar]
- 17. Fuqua WC, Winans SC. 1994. A LuxR-LuxI type regulatory system activates Agrobacterium Ti plasmid conjugal transfer in the presence of a plant tumor metabolite. J Bacteriol 176:2796–2806. doi: 10.1128/jb.176.10.2796-2806.1994 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Moncalián G, Grandoso G, Llosa M, de la Cruz F. 1997. OriT-processing and regulatory roles of TrwA protein in plasmid R388 conjugation. J Mol Biol 270:188–200. doi: 10.1006/jmbi.1997.1082 [DOI] [PubMed] [Google Scholar]
- 19. Strohmaier H, Noiges R, Kotschan S, Sawers G, Högenauer G, Zechner EL, Koraimann G. 1998. Signal transduction and bacterial conjugation: characterization of the role of ArcA in regulating conjugative transfer of the resistance plasmid R1. J Mol Biol 277:309–316. doi: 10.1006/jmbi.1997.1598 [DOI] [PubMed] [Google Scholar]
- 20. Ding H, Yip CB, Hynes MF. 2013. Genetic characterization of a novel rhizobial plasmid conjugation system in Rhizobium leguminosarum bv. viciae strain VF39SM. J Bacteriol 195:328–339. doi: 10.1128/JB.01234-12 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. Carraro N, Matteau D, Luo P, Rodrigue S, Burrus V. 2014. The master activator of IncA/C conjugative plasmids stimulates genomic islands and multidrug resistance dissemination. PLOS Genet 10:e1004714. doi: 10.1371/journal.pgen.1004714 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Bañuelos-Vazquez LA, Torres Tejerizo G, Brom S. 2017. Regulation of conjugative transfer of plasmids and integrative conjugative elements. Plasmid 91:82–89. doi: 10.1016/j.plasmid.2017.04.002 [DOI] [PubMed] [Google Scholar]
- 23. Meijer WJJ, Boer DR, Ares S, Alfonso C, Rojo F, Luque-Ortega JR, Wu LJ. 2021. Multiple layered control of the conjugation process of the Bacillus subtilis plasmid pLS20. Front Mol Biosci 8:648468. doi: 10.3389/fmolb.2021.648468 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Will WR, Frost LS. 2006. Characterization of the opposing roles of H-NS and TraJ in transcriptional regulation of the F-plasmid tra operon. J Bacteriol 188:507–514. doi: 10.1128/JB.188.2.507-514.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Koraimann G, Teferle K, Markolin G, Woger W, Högenauer G. 1996. The FinOP repressor system of plasmid R1: analysis of the antisense RNA control of traJ expression and conjugative DNA transfer. Mol Microbiol 21:811–821. doi: 10.1046/j.1365-2958.1996.361401.x [DOI] [PubMed] [Google Scholar]
- 26. Zahrl D, Wagner A, Tscherner M, Koraimann G. 2007. GroEL plays a central role in stress-induced negative regulation of bacterial conjugation by promoting proteolytic degradation of the activator protein TraJ. J Bacteriol 189:5885–5894. doi: 10.1128/JB.00005-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Arold ST, Leonard PG, Parkinson GN, Ladbury JE. 2010. H-NS forms a superhelical protein scaffold for DNA condensation. Proc Natl Acad Sci U S A 107:15728–15732. doi: 10.1073/pnas.1006966107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Navarre WW, McClelland M, Libby SJ, Fang FC. 2007. Silencing of xenogeneic DNA by H-NS-facilitation of lateral gene transfer in bacteria by a defense system that recognizes foreign DNA. Genes Dev 21:1456–1471. doi: 10.1101/gad.1543107 [DOI] [PubMed] [Google Scholar]
- 29. Will WR, Lu J, Frost LS. 2004. The role of H-NS in silencing F transfer gene expression during entry into stationary phase. Mol Microbiol 54:769–782. doi: 10.1111/j.1365-2958.2004.04303.x [DOI] [PubMed] [Google Scholar]
- 30. Lang KS, Johnson TJ. 2016. Characterization of Acr2, an H-NS-like protein encoded on A/C2-type plasmids. Plasmid 87–88:17–27. doi: 10.1016/j.plasmid.2016.07.004 [DOI] [PubMed] [Google Scholar]
- 31. Dillon SC, Cameron ADS, Hokamp K, Lucchini S, Hinton JCD, Dorman CJ. 2010. Genome-wide analysis of the H-NS and Sfh regulatory networks in Salmonella Typhimurium identifies a plasmid-encoded transcription silencing mechanism. Mol Microbiol 76:1250–1265. doi: 10.1111/j.1365-2958.2010.07173.x [DOI] [PubMed] [Google Scholar]
- 32. Baños RC, Aznar S, Madrid C, Juárez A. 2011. Differential functional properties of chromosomal- and plasmid-encoded H-NS proteins. Res Microbiol 162:382–385. doi: 10.1016/j.resmic.2011.02.003 [DOI] [PubMed] [Google Scholar]
- 33. Baomo L, Lili S, Moran RA, van Schaik W, Chao Z. 2021. Temperature-regulated IncX3 plasmid characteristics and the role of plasmid-encoded H-NS in thermoregulation. Front Microbiol 12:765492. doi: 10.3389/fmicb.2021.765492 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Liu B, Shui L, Zhou K, Jiang Y, Li X, Guan J, Li Q, Zhuo C. 2020. Impact of plasmid-encoded H-NS-like protein on blaNDM-1-bearing IncX3 plasmid in Escherichia coli. J Infect Dis 221:S229–S236. doi: 10.1093/infdis/jiz567 [DOI] [PubMed] [Google Scholar]
- 35. Fitzgerald S, Kary SC, Alshabib EY, MacKenzie KD, Stoebel DM, Chao TC, Cameron ADS. 2020. Redefining the H-NS protein family: a diversity of specialized core and accessory forms exhibit hierarchical transcriptional network integration. Nucleic Acids Res 48:10184–10198. doi: 10.1093/nar/gkaa709 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Johnson TJ, Bielak EM, Fortini D, Hansen LH, Hasman H, Debroy C, Nolan LK, Carattoli A. 2012. Expansion of the IncX plasmid family for improved identification and typing of novel plasmids in drug-resistant Enterobacteriaceae. Plasmid 68:43–50. doi: 10.1016/j.plasmid.2012.03.001 [DOI] [PubMed] [Google Scholar]
- 37. Guo X, Chen R, Wang Q, Li C, Ge H, Qiao J, Li Y. 2022. Global prevalence, characteristics, and future prospects of IncX3 plasmids: A review. Front Microbiol 13:979558. doi: 10.3389/fmicb.2022.979558 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Ali SS, Xia B, Liu J, Navarre WW. 2012. Silencing of foreign DNA in bacteria. Curr Opin Microbiol 15:175–181. doi: 10.1016/j.mib.2011.12.014 [DOI] [PubMed] [Google Scholar]
- 39. Kołodziej M, Trojanowski D, Bury K, Hołówka J, Matysik W, Kąkolewska H, Feddersen H, Giacomelli G, Konieczny I, Bramkamp M, Zakrzewska-Czerwińska J. 2021. Lsr2, a nucleoid-associated protein influencing mycobacterial cell cycle. Sci Rep 11:2910. doi: 10.1038/s41598-021-82295-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Duan B, Ding P, Navarre WW, Liu J, Xia B. 2021. Xenogeneic silencing and bacterial genome evolution: mechanisms for DNA recognition imply multifaceted roles of Xenogeneic silencers. Mol Biol Evol 38:4135–4148. doi: 10.1093/molbev/msab136 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Fernández-de-Alba C, Berrow NS, Garcia-Castellanos R, García J, Pons M. 2013. On the origin of the selectivity of plasmidic H-NS towards horizontally acquired DNA: linking H-NS oligomerization and cooperative DNA binding. J Mol Biol 425:2347–2358. doi: 10.1016/j.jmb.2013.03.006 [DOI] [PubMed] [Google Scholar]
- 42. Page AJ, Cummins CA, Hunt M, Wong VK, Reuter S, Holden MTG, Fookes M, Falush D, Keane JA, Parkhill J. 2015. Roary: rapid large-scale prokaryote pan genome analysis. Bioinformatics 31:3691–3693. doi: 10.1093/bioinformatics/btv421 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Smillie C, Garcillán-Barcia MP, Francia MV, Rocha EPC, de la Cruz F. 2010. Mobility of plasmids. Microbiol Mol Biol Rev 74:434–452. doi: 10.1128/MMBR.00020-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Douarre P-E, Mallet L, Radomski N, Felten A, Mistou M-Y. 2020. Analysis of COMPASS, a new comprehensive plasmid database revealed prevalence of multireplicon and extensive diversity of incf plasmids. Front Microbiol 11:483. doi: 10.3389/fmicb.2020.00483 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45. Zhang J, Nielsen R, Yang Z. 2005. Evaluation of an improved branch-site likelihood method for detecting positive selection at the molecular level. Mol Biol Evol 22:2472–2479. doi: 10.1093/molbev/msi237 [DOI] [PubMed] [Google Scholar]
- 46. Yang Z, Wong WSW, Nielsen R. 2005. Bayes empirical Bayes inference of amino acid sites under positive selection. Mol Biol Evol 22:1107–1118. doi: 10.1093/molbev/msi097 [DOI] [PubMed] [Google Scholar]
- 47. García J, Madrid C, Cendra M, Juárez A, Pons M. 2009. N9L and L9N mutations toggle Hha binding and hemolysin regulation by Escherichia coli and Vibrio cholerae H-NS. FEBS Lett 583:2911–2916. doi: 10.1016/j.febslet.2009.07.054 [DOI] [PubMed] [Google Scholar]
- 48. Yamanaka Y, Winardhi RS, Yamauchi E, Nishiyama S-I, Sowa Y, Yan J, Kawagishi I, Ishihama A, Yamamoto K. 2018. Dimerization site 2 of the bacterial DNA-binding protein H-NS is required for gene silencing and stiffened nucleoprotein filament formation. J Biol Chem 293:9496–9505. doi: 10.1074/jbc.RA117.001425 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Shahul Hameed UF, Liao C, Radhakrishnan AK, Huser F, Aljedani SS, Zhao X, Momin AA, Melo FA, Guo X, Brooks C, Li Y, Cui X, Gao X, Ladbury JE, Jaremko Ł, Jaremko M, Li J, Arold ST. 2019. H-NS uses an autoinhibitory conformational switch for environment-controlled gene silencing. Nucleic Acids Res 47:2666–2680. doi: 10.1093/nar/gky1299 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50. Zhao X, Shahul Hameed UF, Kharchenko V, Liao C, Huser F, Remington JM, Radhakrishnan AK, Jaremko M, Jaremko Ł, Arold ST, Li J. 2021. Molecular basis for the adaptive evolution of environment-sensing by H-NS proteins. Elife 10:e57467. doi: 10.7554/eLife.57467 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51. Madrid C, Nieto JM, Paytubi S, Falconi M, Gualerzi CO, Juárez A. 2002. Temperature- and H-NS-dependent regulation of a plasmid-encoded virulence operon expressing Escherichia coli hemolysin. J Bacteriol 184:5058–5066. doi: 10.1128/JB.184.18.5058-5066.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Ono S, Goldberg MD, Olsson T, Esposito D, Hinton JCD, Ladbury JE. 2005. H-NS is a part of a thermally controlled mechanism for bacterial gene regulation. Biochem J 391:203–213. doi: 10.1042/BJ20050453 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53. White-Ziegler CA, Davis TR. 2009. Genome-wide identification of H-NS-controlled, temperature-regulated genes in Escherichia coli K-12. J Bacteriol 191:1106–1110. doi: 10.1128/JB.00599-08 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Cordeiro TN, García J, Bernadó P, Millet O, Pons M. 2015. A three-protein charge zipper stabilizes a complex modulating bacterial gene silencing. J Biol Chem 290:21200–21212. doi: 10.1074/jbc.M114.630400 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Hu L, Kong W, Yang D, Han Q, Guo L, Shi Y. 2019. Threonine phosphorylation fine-tunes the regulatory activity of histone-like nucleoid structuring protein in Salmonella transcription. Front Microbiol 10. doi: 10.3389/fmicb.2019.01515 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Esposito D, Petrovic A, Harris R, Ono S, Eccleston JF, Mbabaali A, Haq I, Higgins CF, Hinton JCD, Driscoll PC, Ladbury JE. 2002. H-NS oligomerization domain structure reveals the mechanism for high order self-association of the intact protein. J Mol Biol 324:841–850. doi: 10.1016/s0022-2836(02)01141-5 [DOI] [PubMed] [Google Scholar]
- 57. Vreede J, Dame RTh. 2012. Predicting the effect of Ions on the conformation of the H-NS dimerization domain. Biophys J 103:89–98. doi: 10.1016/j.bpj.2012.05.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Johansson J, Eriksson S, Sondén B, Wai SN, Uhlin BE. 2001. Heteromeric interactions among nucleoid-associated bacterial proteins: localization of StpA-stabilizing regions in H-NS of Escherichia coli. J Bacteriol 183:2343–2347. doi: 10.1128/JB.183.7.2343-2347.2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Ueguchi C, Seto C, Suzuki T, Mizuno T. 1997. Clarification of the dimerization domain and its functional significance for the *Escherichia coli* nucleoid protein H-NS. J Mol Biol 274:145–151. doi: 10.1006/jmbi.1997.1381 [DOI] [PubMed] [Google Scholar]
- 60. Wang W, Li G-W, Chen C, Xie XS, Zhuang X. 2011. Chromosome organization by a nucleoid-associated protein in live bacteria. Science 333:1445–1449. doi: 10.1126/science.1204697 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Nye MB, Taylor RK. 2003. Vibrio cholerae H-NS domain structure and function with respect to transcriptional repression of ToxR regulon genes reveals differences among H-NS family members. Mol Microbiol 50:427–444. doi: 10.1046/j.1365-2958.2003.03701.x [DOI] [PubMed] [Google Scholar]
- 62. Giangrossi M, Wintraecken K, Spurio R, de Vries R. 2014. Probing the relation between protein–protein interactions and DNA binding for a linker mutant of the bacterial nucleoid protein H-NS. Biochimica et Biophysica Acta (BBA) - Prot Proteom 1844:339–345. doi: 10.1016/j.bbapap.2013.11.010 [DOI] [PubMed] [Google Scholar]
- 63. Shindo H, Ohnuki A, Ginba H, Katoh E, Ueguchi C, Mizuno T, Yamazaki T. 1999. Identification of the DNA binding surface of H-NS protein from Escherichia coli by heteronuclear NMR spectroscopy. FEBS Lett 455:63–69. doi: 10.1016/s0014-5793(99)00862-5 [DOI] [PubMed] [Google Scholar]
- 64. Gao Y, Foo YH, Winardhi RS, Tang Q, Yan J, Kenney LJ. 2017. Charged residues in the *H-NS* linker drive DNA binding and gene silencing in single cells. Proc Natl Acad Sci U S A 114:12560–12565. doi: 10.1073/pnas.1716721114 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Ueguchi C, Suzuki T, Yoshida T, Tanaka K, Mizuno T. 1996. Systematic mutational analysis revealing the functional domain organization of Escherichia coli nucleoid protein H-NS. J Mol Biol 263:149–162. doi: 10.1006/jmbi.1996.0566 [DOI] [PubMed] [Google Scholar]
- 66. Gordon BRG, Li Y, Cote A, Weirauch MT, Ding P, Hughes TR, Navarre WW, Xia B, Liu J. 2011. Structural basis for recognition of AT-rich DNA by unrelated xenogeneic silencing proteins. Proc Natl Acad Sci U S A 108:10690–10695. doi: 10.1073/pnas.1102544108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Donato GM, Kawula TH. 1999. Phenotypic analysis of random hns mutations differentiate DNA-binding activity from properties of fima promoter inversion modulation and bacterial motility. J Bacteriol 181:941–948. doi: 10.1128/JB.181.3.941-948.1999 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68. Zhang L, Li S, Liu Q, Duan C, Li J, Sun H, Xu L, Yang Q, Wang Y, Shen X. 2023. Cyclic di-GMP rescues H-NS–mediated silencing of bacterial type VI secretion systems. Res Squ. doi: 10.21203/rs.3.rs-3177235/v1 [DOI] [Google Scholar]
- 69. Leonard PG. 2008. Structural and biophysical characterisation of the histone-like nucleoid structuring protein from enteric bacteria. University College London. [Google Scholar]
- 70. Dilweg IW, Dame RT. 2018. Post-translational modification of nucleoid-associated proteins: an extra layer of functional modulation in bacteria? Biochem Soc Trans 46:1381–1392. doi: 10.1042/BST20180488 [DOI] [PubMed] [Google Scholar]
- 71. Dong H, Zhao Y, Bi C, Han Y, Zhang J, Bai X, Zhai G, Zhang H, Tian S, Hu D, Xu L, Zhang K. 2022. TmcA functions as a lysine 2-hydroxyisobutyryltransferase to regulate transcription. Nat Chem Biol 18:142–151. doi: 10.1038/s41589-021-00906-3 [DOI] [PubMed] [Google Scholar]
- 72. Christensen BB, Sternberg C, Andersen JB, Eberl L, Møller S, Givskov M, Molin S. 1998. Establishment of new genetic traits in a microbial biofilm community. Appl Environ Microbiol 64:2247–2255. doi: 10.1128/AEM.64.6.2247-2255.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Forns N, Baños RC, Balsalobre C, Juárez A, Madrid C. 2005. Temperature-dependent conjugative transfer of R27: role of chromosome- and plasmid-encoded Hha and H-NS proteins. J Bacteriol 187:3950–3959. doi: 10.1128/JB.187.12.3950-3959.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74. Beloin C, Deighan P, Doyle M, Dorman CJ. 2003. Shigella flexneri 2a strain 2457T expresses three members of the H-NS-like protein family: characterization of the Sfh protein. Mol Genet Genomics 270:66–77. doi: 10.1007/s00438-003-0897-0 [DOI] [PubMed] [Google Scholar]
- 75. Paul K, Carlquist WC, Blair DF. 2011. Adjusting the spokes of the flagellar motor with the DNA-binding protein H-NS. J Bacteriol 193:5914–5922. doi: 10.1128/JB.05458-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Free A, Williams RM, Dorman CJ. 1998. The StpA protein functions as a molecular adapter to mediate repression of the bgl operon by truncated H-NS in Escherichia coli. J Bacteriol 180:994–997. doi: 10.1128/JB.180.4.994-997.1998 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77. Srinivasan R, Chandraprakash D, Krishnamurthi R, Singh P, Scolari VF, Krishna S, Seshasayee ASN. 2013. Genomic analysis reveals epistatic silencing of “expensive” genes in Escherichia coli K-12. Mol Biosyst 9:2021–2033. doi: 10.1039/c3mb70035f [DOI] [PubMed] [Google Scholar]
- 78. Boudreau BA, Hron DR, Qin L, van der Valk RA, Kotlajich MV, Dame RT, Landick R. 2018. StpA and Hha stimulate pausing by RNA polymerase by promoting DNA-DNA bridging of H-NS filaments. Nucleic Acids Res 46:5525–5546. doi: 10.1093/nar/gky265 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Nieto JM, Madrid C, Prenafeta A, Miquelay E, Balsalobre C, Carrascal M, Juárez A. 2000. Expression of the hemolysin operon in Escherichia coli is modulated by a nucleoid-protein complex that includes the proteins Hha and H-NS. Mol Gen Genet 263:349–358. doi: 10.1007/s004380051178 [DOI] [PubMed] [Google Scholar]
- 80. Kim MS, Bae SH, Yun SH, Lee HJ, Ji SC, Lee JH, Srivastava P, Lee SH, Chae H, Lee Y, Choi BS, Chattoraj DK, Lim HM. 2005. Cnu, a novel oriC-binding protein of Escherichia coli. J Bacteriol 187:6998–7008. doi: 10.1128/JB.187.20.6998-7008.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Ueda T, Takahashi H, Uyar E, Ishikawa S, Ogasawara N, Oshima T. 2013. Functions of the Hha and YdgT proteins in transcriptional silencing by the nucleoid proteins, H-NS and StpA, in Escherichia coli. DNA Res 20:263–271. doi: 10.1093/dnares/dst008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82. Leonard PG, Ono S, Gor J, Perkins SJ, Ladbury JE. 2009. Investigation of the self-association and hetero-association interactions of H-NS and StpA from Enterobacteria. Mol Microbiol 73:165–179. doi: 10.1111/j.1365-2958.2009.06754.x [DOI] [PubMed] [Google Scholar]
- 83. Lim CJ, Whang YR, Kenney LJ, Yan J. 2012. Gene silencing H-NS paralogue StpA forms a rigid protein filament along DNA that blocks DNA accessibility. Nucleic Acids Res 40:3316–3328. doi: 10.1093/nar/gkr1247 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Ali SS, Whitney JC, Stevenson J, Robinson H, Howell PL, Navarre WW. 2013. Structural insights into the regulation of foreign genes in Salmonella by the Hha/H-NS complex. J Biol Chem 288:13356–13369. doi: 10.1074/jbc.M113.455378 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Bae S-H, Liu D, Lim HM, Lee Y, Choi B-S. 2008. Structure of the nucleoid-associated protein Cnu reveals common binding sites for H-NS in Cnu and Hha. Biochemistry 47:1993–2001. doi: 10.1021/bi701914t [DOI] [PubMed] [Google Scholar]
- 86. Sharma VK, Zuerner RL. 2004. Role of hha and ler in transcriptional regulation of the esp operon of enterohemorrhagic Escherichia coli O157:H7. J Bacteriol 186:7290–7301. doi: 10.1128/JB.186.21.7290-7301.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Fukui N, Oshima T, Ueda T, Ogasawara N, Tobe T. 2016. Gene activation through the modulation of nucleoid structures by a horizontally transferred regulator, Pch, in Enterohemorrhagic Escherichia coli. PLOS ONE 11:e0149718. doi: 10.1371/journal.pone.0149718 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Silphaduang U, Mascarenhas M, Karmali M, Coombes BK. 2007. Repression of intracellular virulence factors in Salmonella by the Hha and YdgT nucleoid-associated proteins. J Bacteriol 189:3669–3673. doi: 10.1128/JB.00002-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Takeda T, Yun C-S, Shintani M, Yamane H, Nojiri H. 2011. Distribution of genes encoding nucleoid-associated protein homologs in plasmids. Int J Evol Biol 2011:685015. doi: 10.4061/2011/685015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Narayan A, Gopi S, Fushman D, Naganathan AN. 2019. A binding cooperativity switch driven by synergistic structural swelling of an osmo-regulatory protein pair. Nat Commun 10:1995. doi: 10.1038/s41467-019-10002-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Deighan P, Beloin C, Dorman CJ. 2003. Three-way interactions among the Sfh, StpA and H-NS nucleoid-structuring proteins of Shigella flexneri 2a strain 2457T. Mol Microbiol 48:1401–1416. doi: 10.1046/j.1365-2958.2003.03515.x [DOI] [PubMed] [Google Scholar]
- 92. McGovern V, Higgins NP, Chiz RS, Jaworski A. 1994. H-NS over-expression induces an artificial stationary phase by silencing global transcription. Biochimie 76:1019–1029. doi: 10.1016/0300-9084(94)90026-4 [DOI] [PubMed] [Google Scholar]
- 93. Williams RM, Rimsky S, Buc H. 1996. Probing the structure, function, and interactions of the Escherichia coli H-NS and StpA proteins by using dominant negative derivatives. J Bacteriol 178:4335–4343. doi: 10.1128/jb.178.15.4335-4343.1996 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94. Nieto JM, Madrid C, Miquelay E, Parra JL, Rodríguez S, Juárez A. 2002. Evidence for direct protein-protein interaction between members of the enterobacterial Hha/YmoA and H-NS families of proteins. J Bacteriol 184:629–635. doi: 10.1128/JB.184.3.629-635.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95. Paytubi S, Madrid C, Forns N, Nieto JM, Balsalobre C, Uhlin BE, Juárez A. 2004. YdgT, the Hha paralogue in Escherichia coli, forms heteromeric complexes with H-NS and StpA. Mol Microbiol 54:251–263. doi: 10.1111/j.1365-2958.2004.04268.x [DOI] [PubMed] [Google Scholar]
- 96. Dahlberg C, Chao L. 2003. Amelioration of the cost of conjugative plasmid carriage in Eschericha coli K12. Genetics 165:1641–1649. doi: 10.1093/genetics/165.4.1641 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97. Porse A, Schønning K, Munck C, Sommer MOA. 2016. Survival and evolution of a large multidrug resistance plasmid in new clinical bacterial hosts. Mol Biol Evol 33:2860–2873. doi: 10.1093/molbev/msw163 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98. Turner PE, Cooper VS, Lenski RE. 1998. Tradeoff between horizontal and vertical modes of transmission in bacterial plasmids. Evolution (N Y) 52:315–329. doi: 10.1111/j.1558-5646.1998.tb01634.x [DOI] [PubMed] [Google Scholar]
- 99. Iwasaki W, Takagi T. 2009. Rapid pathway evolution facilitated by horizontal gene transfers across prokaryotic lineages. PLoS Genet 5:e1000402. doi: 10.1371/journal.pgen.1000402 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100. Jackson RW, Vinatzer B, Arnold DL, Dorus S, Murillo J. 2011. The influence of the accessory genome on bacterial pathogen evolution. Mob Genet Elements 1:55–65. doi: 10.4161/mge.1.1.16432 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101. Botelho J, Schulenburg H. 2021. The role of integrative and conjugative elements in antibiotic resistance evolution. Trends Microbiol 29:8–18. doi: 10.1016/j.tim.2020.05.011 [DOI] [PubMed] [Google Scholar]
- 102. Ohno S. 1970. Evolution by gene duplication. Springer, Berlin, Heidelberg. [Google Scholar]
- 103. Zhang J. 2003. Evolution by gene duplication: an update. Trend Ecol Evol 18:292–298. doi: 10.1016/S0169-5347(03)00033-8 [DOI] [Google Scholar]
- 104. Sonden B, Uhlin BE. 1996. Coordinated and differential expression of histone-like proteins in Escherichia coli: regulation and function of the H-NS analog StpA. EMBO J 15:4970–4980. [PMC free article] [PubMed] [Google Scholar]
- 105. Kimura M. 1983. The neutral theory of molecular evolution. Cambridge University Press, Cambridge. [Google Scholar]
- 106. Zhang J, Rosenberg HF, Nei M. 1998. Positive darwinian selection after gene duplication in primate ribonuclease genes. Proc Natl Acad Sci USA 95:3708–3713. doi: 10.1073/pnas.95.7.3708 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107. Hughes AL, Green JA, Garbayo JM, Roberts RM. 2000. Adaptive diversification within a large family of recently duplicated, placentally expressed genes. Proc Natl Acad Sci U S A 97:3319–3323. doi: 10.1073/pnas.97.7.3319 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108. Zhang J, Zhang Y, Rosenberg HF. 2002. Adaptive evolution of a duplicated pancreatic ribonuclease gene in a leaf-eating monkey. Nat Genet 30:411–415. doi: 10.1038/ng852 [DOI] [PubMed] [Google Scholar]
- 109. Santer M, Uecker H. 2020. Evolutionary rescue and drug resistance on multicopy plasmids. Genetics 215:847–868. doi: 10.1534/genetics.119.303012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110. Ilhan J, Kupczok A, Woehle C, Wein T, Hülter NF, Rosenstiel P, Landan G, Mizrahi I, Dagan T. 2019. Segregational drift and the interplay between plasmid copy number and evolvability. Mol Biol Evol 36:472–486. doi: 10.1093/molbev/msy225 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111. Carmona M, Balsalobre C, Muñoa F, Mouriño M, Jubete Y, De la Cruz F, Juárez A. 1993. Escherichia coli hha mutants, DNA supercoiling and expression of the haemolysin genes from the recombinant plasmid pANN202-312. Mol Microbiol 9:1011–1018. doi: 10.1111/j.1365-2958.1993.tb01230.x [DOI] [PubMed] [Google Scholar]
- 112. Cook GM, Berney M, Gebhard S, Heinemann M, Cox RA, Danilchanka O, Niederweis M. 2009. Physiology of Mycobacteria. Adv Microb Physiol 55:81–182, doi: 10.1016/S0065-2911(09)05502-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113. Gordon BRG, Li Y, Wang L, Sintsova A, van Bakel H, Tian S, Navarre WW, Xia B, Liu J. 2010. Lsr2 is a nucleoid-associated protein that targets AT-rich sequences and virulence genes in Mycobacterium tuberculosis. Proc Natl Acad Sci U S A 107:5154–5159. doi: 10.1073/pnas.0913551107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114. Rawlings DE, Tietze E. 2001. Comparative biology of IncQ and IncQ-like plasmids. Microbiol Mol Biol Rev 65:481–496. doi: 10.1128/MMBR.65.4.481-496.2001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115. Ares-Arroyo M, Bernabe-Balas C, Santos-Lopez A, Baquero MR, Prasad KN, Cid D, Martin-Espada C, San Millan A, Gonzalez-Zorn B. 2018. PCR-based analysis of ColE1 plasmids in clinical isolates and metagenomic samples reveals their importance as gene capture platforms. Front Microbiol 9:469. doi: 10.3389/fmicb.2018.00469 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116. Lang B, Blot N, Bouffartigues E, Buckle M, Geertz M, Gualerzi CO, Mavathur R, Muskhelishvili G, Pon CL, Rimsky S, Stella S, Babu MM, Travers A. 2007. High-affinity DNA binding sites for H-NS provide a molecular basis for selective silencing within proteobacterial genomes. Nucleic Acids Res 35:6330–6337. doi: 10.1093/nar/gkm712 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117. Haft RJF, Mittler JE, Traxler B. 2009. Competition favours reduced cost of plasmids to host bacteria. ISME J 3:761–769. doi: 10.1038/ismej.2009.22 [DOI] [PubMed] [Google Scholar]
- 118. Poidevin M, Sato M, Altinoglu I, Delaplace M, Sato C, Yamaichi Y. 2018. Mutation in ESBL plasmid from Escherichia coli O104:H4 leads autoagglutination and enhanced plasmid dissemination. Front Microbiol 9:130. doi: 10.3389/fmicb.2018.00130 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119. Potron A, Poirel L, Nordmann P. 2014. Derepressed transfer properties leading to the efficient spread of the plasmid encoding carbapenemase OXA-48. Antimicrob Agents Chemother 58:467–471. doi: 10.1128/AAC.01344-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120. Baba T, Ara T, Hasegawa M, Takai Y, Okumura Y, Baba M, Datsenko KA, Tomita M, Wanner BL, Mori H. 2006. Construction of Escherichia coli K‐12 in‐frame, single‐gene knockout mutants: the Keio collection. Mol Syst Biol 2:2006. doi: 10.1038/msb4100050 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121. Nguyen L-T, Schmidt HA, von Haeseler A, Minh BQ. 2015. IQ-TREE: a fast and effective stochastic algorithm for estimating maximum-likelihood phylogenies. Mol Biol Evol 32:268–274. doi: 10.1093/molbev/msu300 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122. Kalyaanamoorthy S, Minh BQ, Wong TKF, von Haeseler A, Jermiin LS. 2017. ModelFinder: fast model selection for accurate phylogenetic estimates. Nat Methods 14:587–589. doi: 10.1038/nmeth.4285 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123. Minh BQ, Nguyen MAT, von Haeseler A. 2013. Ultrafast approximation for phylogenetic bootstrap. Mol Biol Evol 30:1188–1195. doi: 10.1093/molbev/mst024 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124. Guindon S, Dufayard JF, Lefort V, Anisimova M, Hordijk W, Gascuel O. 2010. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol 59:307–321. doi: 10.1093/sysbio/syq010 [DOI] [PubMed] [Google Scholar]
- 125. Alnajar S, Gupta RS. 2017. Phylogenomics and comparative genomic studies delineate six main clades within the family Enterobacteriaceae and support the reclassification of several polyphyletic members of the family. Infect Genet Evol 54:108–127. doi: 10.1016/j.meegid.2017.06.024 [DOI] [PubMed] [Google Scholar]
- 126. Yu G, Smith DK, Zhu H, Guan Y, Lam TTY. 2017. Ggtree: an r package for visualization and annotation of phylogenetic trees with their covariates and other associated data. Methods Ecol Evol 8. doi: 10.1111/2041-210X.12628 [DOI] [Google Scholar]
- 127. Katoh K, Misawa K, Kuma K, Miyata T. 2002. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res 30:3059–3066. doi: 10.1093/nar/gkf436 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128. Ashkenazy H, Abadi S, Martz E, Chay O, Mayrose I, Pupko T, Ben-Tal N. 2016. ConSurf 2016: an improved methodology to estimate and visualize evolutionary conservation in macromolecules. Nucleic Acids Res 44:W344–50. doi: 10.1093/nar/gkw408 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129. Carattoli A, Hasman H. 2020. PlasmidFinder and In Silico pMLST: identification and typing of plasmid replicons in whole-genome sequencing (WGS).p 285–294. In Cruz F (ed), Horizontal gene transfer: methods and protocols. Springer US, New York. [DOI] [PubMed] [Google Scholar]
- 130. Schmartz GP, Hartung A, Hirsch P, Kern F, Fehlmann T, Müller R, Keller A. 2022. PLSDB: advancing a comprehensive database of bacterial plasmids. Nucleic Acids Res 50:D273–D278. doi: 10.1093/nar/gkab1111 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131. Tatusova T, DiCuccio M, Badretdin A, Chetvernin V, Nawrocki EP, Zaslavsky L, Lomsadze A, Pruitt KD, Borodovsky M, Ostell J. 2016. NCBI prokaryotic genome annotation pipeline. Nucleic Acids Res 44:6614–6624. doi: 10.1093/nar/gkw569 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132. Robertson J, Nash JHE. 2018. MOB-suite: software tools for clustering, reconstruction and typing of plasmids from draft assemblies. Microb Genom 4:e000206. doi: 10.1099/mgen.0.000206 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133. Emms DM, Kelly S. 2019. OrthoFinder: phylogenetic orthology inference for comparative genomics. Genome Biol 20:238. doi: 10.1186/s13059-019-1832-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134. Wright ES. 2015. DECIPHER: harnessing local sequence context to improve protein multiple sequence alignment. BMC Bioinformatics 16:322. doi: 10.1186/s12859-015-0749-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135. Sayers EW, Bolton EE, Brister JR, Canese K, Chan J, Comeau DC, Connor R, Funk K, Kelly C, Kim S, Madej T, Marchler-Bauer A, Lanczycki C, Lathrop S, Lu Z, Thibaud-Nissen F, Murphy T, Phan L, Skripchenko Y, Tse T, Wang J, Williams R, Trawick BW, Pruitt KD, Sherry ST. 2022. Database resources of the national center for biotechnology information. Nucleic Acids Res 50:D20–D26. doi: 10.1093/nar/gkab1112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136. Steinegger M, Söding J. 2017. MMseqs2 enables sensitive protein sequence searching for the analysis of massive data sets. Nat Biotechnol 35:1026–1028. doi: 10.1038/nbt.3988 [DOI] [PubMed] [Google Scholar]
- 137. Sela I, Ashkenazy H, Katoh K, Pupko T. 2015. GUIDANCE2: accurate detection of unreliable alignment regions accounting for the uncertainty of multiple parameters. Nucleic Acids Res 43:W7–14. doi: 10.1093/nar/gkv318 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 138. Mellies JL, Benison G, McNitt W, Mavor D, Boniface C, Larabee FJ. 2011. Ler of pathogenic Escherichia coli forms toroidal protein-DNA complexes. Microbiology (Reading) 157:1123–1133. doi: 10.1099/mic.0.046094-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 139. Yang Z. 2007. PAML 4: phylogenetic analysis by maximum likelihood. Mol Biol Evol 24:586–591. doi: 10.1093/molbev/msm088 [DOI] [PubMed] [Google Scholar]
- 140. Yang Z, Nielsent R. 2002. Codon-substitution models for detecting molecular adaptation at individual sites along specific lineages. Mol Biol Evol 19:908–917. doi: 10.1093/oxfordjournals.molbev.a004148 [DOI] [PubMed] [Google Scholar]
- 141. Bielawski JP, Yang Z. 2004. A maximum likelihood method for detecting functional divergence at individual codon sites, with application to gene family evolution. J Mol Evol 59:121–132. doi: 10.1007/s00239-004-2597-8 [DOI] [PubMed] [Google Scholar]
- 142. Weadick CJ, Chang BSW. 2012. Complex patterns of divergence among green-sensitive (RH2a) African cichlid opsins revealed by Clade model analyses. BMC Evol Biol 12:206. doi: 10.1186/1471-2148-12-206 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 143. Mirdita M, Schütze K, Moriwaki Y, Heo L, Ovchinnikov S, Steinegger M. 2022. ColabFold: making protein folding accessible to all. Nat Methods 19:679–682. doi: 10.1038/s41592-022-01488-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144. Porollo AA, Adamczak R, Meller J. 2004. POLYVIEW: a flexible visualization tool for structural and functional annotations of proteins. Bioinformatics 20:2460–2462. doi: 10.1093/bioinformatics/bth248 [DOI] [PubMed] [Google Scholar]
- 145. Wilkins D. 2020. gggenes: draw gene arrow maps in “ggplot2". Comprehensive R Archive Network (CRAN). [Google Scholar]
- 146. Delano WL. 2002. The PyMOL molecular graphics system. CCP4 newsl protein crystallogr 40
- 147. Caffrey DR, Dana PH, Mathur V, Ocano M, Hong E-J, Wang YE, Somaroo S, Caffrey BE, Potluri S, Huang ES. 2007. PFAAT version 2.0: a tool for editing, annotating, and analyzing multiple sequence alignments. BMC Bioinformatics 8:381. doi: 10.1186/1471-2105-8-381 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148. Crooks GE, Hon G, Chandonia J-M, Brenner SE. 2004. WebLogo: a sequence logo generator. Genome Res 14:1188–1190. doi: 10.1101/gr.849004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149. Siguier P, Perochon J, Lestrade L, Mahillon J, Chandler M. 2006. ISfinder: the reference centre for bacterial insertion sequences. Nucleic Acids Res 34:D32–6. doi: 10.1093/nar/gkj014 [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Tables S1 to S6.
Figures S1 to S5.