Skip to main content
Advertisement
  • Loading metrics

Adaptive β-lactam resistance from an inducible efflux pump that is post-translationally regulated by the DjlA co-chaperone

  • Jordan Costafrolaz,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Project administration, Resources, Writing – original draft, Writing – review & editing

    Affiliation Department of Microbiology and Molecular Medicine, Faculty of Medicine/Centre Médical Universitaire, University of Geneva, Geneva, Switzerland

  • Gaël Panis,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Resources, Writing – review & editing

    Affiliation Department of Microbiology and Molecular Medicine, Faculty of Medicine/Centre Médical Universitaire, University of Geneva, Geneva, Switzerland

  • Bastien Casu,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Writing – review & editing

    Affiliation Department of Biology, Institute of Molecular Biology & Biophysics, Eidgenössische Technische Hochschule Zürich, Zürich, Switzerland

  • Silvia Ardissone,

    Roles Conceptualization, Data curation, Formal analysis, Investigation, Methodology, Writing – review & editing

    Affiliation Department of Microbiology and Molecular Medicine, Faculty of Medicine/Centre Médical Universitaire, University of Geneva, Geneva, Switzerland

  • Laurence Degeorges,

    Roles Data curation, Formal analysis, Investigation, Methodology

    Affiliation Department of Microbiology and Molecular Medicine, Faculty of Medicine/Centre Médical Universitaire, University of Geneva, Geneva, Switzerland

  • Martin Pilhofer,

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Project administration, Supervision, Writing – review & editing

    Affiliation Department of Biology, Institute of Molecular Biology & Biophysics, Eidgenössische Technische Hochschule Zürich, Zürich, Switzerland

  • Patrick H. Viollier

    Roles Conceptualization, Data curation, Formal analysis, Funding acquisition, Investigation, Methodology, Project administration, Supervision, Writing – original draft, Writing – review & editing

    patrick.viollier@unige.ch

    Affiliation Department of Microbiology and Molecular Medicine, Faculty of Medicine/Centre Médical Universitaire, University of Geneva, Geneva, Switzerland

Abstract

The acquisition of multidrug resistance (MDR) determinants jeopardizes treatment of bacterial infections with antibiotics. The tripartite efflux pump AcrAB-NodT confers adaptive MDR in the polarized α-proteobacterium Caulobacter crescentus via transcriptional induction by first-generation quinolone antibiotics. We discovered that overexpression of AcrAB-NodT by mutation or exogenous inducers confers resistance to cephalosporin and penicillin (β-lactam) antibiotics. Combining 2-step mutagenesis-sequencing (Mut-Seq) and cephalosporin-resistant point mutants, we dissected how TipR uses a common operator of the divergent tipR and acrAB-nodT promoter for adaptive and/or potentiated AcrAB-NodT-directed efflux. Chemical screening identified diverse compounds that interfere with DNA binding by TipR or induce its dependent proteolytic turnover. We found that long-term induction of AcrAB-NodT deforms the envelope and that homeostatic control by TipR includes co-induction of the DnaJ-like co-chaperone DjlA, boosting pump assembly and/or capacity in anticipation of envelope stress. Thus, the adaptive MDR regulatory circuitry reconciles drug efflux with co-chaperone function for trans-envelope assemblies and maintenance.

Introduction

Bacterial multidrug resistance (MDR) is jeopardizing treatment of bacterial infections with antibiotics [1,2]. The MDR challenge is particularly grave for diderm (gram-negative) pathogens that already possess an intrinsic MDR determinant: a protective outer membrane (OM) that prevents entry of soluble antibiotics [3]. The OM is typically structured as an asymmetric lipid bilayer, with an inner leaflet of phospholipids and an outer leaflet containing lipopolysaccharide (LPS), a charged glycolipid. By contrast, the inner (or cytoplasmic) membrane (IM) is a phospholipid bilayer, but both IM and OM flank the bacterial cell wall (CW; Fig 1A), an important shape-determining structure that is essential for life under most growth conditions. While the CW is the target of many natural antibiotics, the OM prevents these antibiotics from reaching the CW [4]. Moreover, owing to its unusual lipid composition, the OM is also hydrophilic barrier for hydrophobic antibiotics and as well as large hydrophilic antibiotics that only cross the OM via aqueous pores or nutrient transporters [4,5].

thumbnail
Fig 1. NAL-dependent derepression of AcrAB-NodT and TipR confers adaptive β-lactam resistance.

(A) Schematic of the tipR/acrAB-nodT locus and the repression of the divergent promoter by TipR, which is antagonized by the quinolone antibiotic Nal. The AcrAB-NodT components assemble into a tripartite complex spanning the IM, CW (grey shading), and OM to expel the diverse compounds indicated on the right via efflux. (B) Antibiograms of Caulobacter crescentus strains on PYE. Antibiotic discs, from top left to bottom right: Pivmecillinam 20 μg, Mecillinam 10 μg, Ampicillin 100 μg, Ceftazidime 40 μg, Meropenem 10 μg, Amoxicillin 4 μg, Cephalothin 30 μg, Cefepime 30 μg, Piperacillin 100 μg, Oxacillin 5 μg, Doripenem 10 μg, Cefotaxime 30 μg, Moxalactam 30 μg, Imipenem 10 μg, Aztreonam 30 μg, Cephalexin 40 μg, Cefsulodin 30 μg. Nal induction performed at a concentration of 20 μg/mL. (C) Dot plot representation of the RNA-Seq analysis realised on C. crescentus NA1000 (WT). RNA was extracted from biological replicates of exponential phase cells harvested after induction with and without Nal 20 μg/mL for 30 minutes in PYE. Each dot represents the fold-enrichment of genes expression across the genome (represented as linear starting from the origin of replication) in the Nal-induced condition. The data from the analysis are deposited in S1 Data. (D) Activity of β-galactosidase of the PacrA-lacZ in different mutants of C. crescentus. The « + » indicates induction by Nal 10 μg/mL for 2 hours. All levels are indicated in percentage of expression regarding the basal level of NA1000 (WT) without induction. The data from the analysis are deposited in S2 Data. (E) Immunoblot with anti-AcrA and anti-TipR antibodies performed on NA1000 (WT) and tipR mutant (tipR::Tn) cells with and without induction with of Nal 20 μg/mL for 2 hours. (F) Representation of all TipR binding sites on the chromosome of C. crescentus (represented as linear starting from the origin of replication). The black line represents the RPM obtained in the condition without induction, while the blue line represent the RPM after 30 minutes of treatment with 20 μg/mL of Nal. Name above indicate the locus. The data from the analysis are deposited in S3 Data. CW, cell wall; IM, inner membrane; Nal, nalidixic acid; OM, outer membrane; PYE, peptone-yeast extract; RPM, reads per million; WT, wild-type.

https://doi.org/10.1371/journal.pbio.3002040.g001

Additional protection against antibiotics that manage to traverse the OM comes from inducible and envelope-spanning multidrug efflux pumps, such as members of the tripartite resistance-nodulation-division (RND) transporter family [5,6]. RND pumps can expel a myriad of different noxious molecules (including antibiotics) from the cell using the proton motive force (PMF). Because RND pump expression can be induced by small molecules, they essentially function confer adaptive MDR. In the uninduced state, transcriptional repressors minimize expression of efflux pumps by binding to their promoters to prevent firing. However, small molecule inducers interact with these repressors, dislodging them from the promoter to relieve repression of the efflux pump gene [7,8] (Fig 1A). Thus, these transcriptional repressors act as single-component chemical sensors for the expulsion of toxic molecules that are encountered and produced, for example, by competing microbes. Although induction of AcrAB-NodT pumps rids the cell from toxic molecules, the massive synthesis and assembly of new trans-envelope complexes can also pose (sometimes even lethal) stress to the envelope [9,10]. Thus, protective measures are needed in anticipation of such imbalances, deformations, and/or prolonged stress of the envelope. One possible remedy to avert these problems is to shut off AcrAB-NodT synthesis rapidly after expulsion of the inducer by reinstating promoter repression, but additional mechanisms likely exist.

The best-studied RND-type multidrug efflux pump AcrAB-TolC from the gram-negative γ-proteobacterium Escherichia coli has a wide specificity of efflux substrates, including clinically important antimicrobial compounds, dyes, fatty acids, monovalent and bivalent cationic lipophilic antiseptics and disinfectants, detergents, and solvents [5,8,11]. The tripartite system is encoded in the acrAB-tolC operon that is repressed by the cis-encoded transcriptional repressor AcrR [12]. We recently described the RND-type efflux pump, AcrAB-NodT, which confers adaptive resistance in the oligotrophic α-proteobacterium Caulobacter crescentus, a model system for cell cycle and polarity studies. C. crescentus AcrAB-NodT is expressed from a promoter (PacrA) that is inducible by the first-generation quinolone nalidixic acid (NAL) and related molecules [9,13], and this AcrAB-NodT induction by NAL can be lethal to certain envelope mutants, for example, mutants lacking the TipN polarity factor [9] or the conserved TrcR protein, an RNA polymerase (RNAP)-associated factor [10]. PacrA is repressed by TipR, a TetR-like DNA-binding protein that is divergently encoded and expressed from a shared promoter region that spans 167 nucleotides between the predicted start codons of AcrA and TipR (Fig 1A).

TetR-type transcriptional repressors are characterised by an N-terminal DNA-binding domain with a helix-turn-helix (HTH) motif and a C-terminal regulatory domain interacting with ligands [14]. Binding of an inducer to the C-terminal domain induces a conformational change that ultimately causes its dissociation from the target DNA, liberating the promoter for transcriptional activation by RNAP [15,16]. TetR-like proteins are known to regulate a wide range of cellular functions, ranging from osmotic and general stress homeostasis, envelope maintenance, virulence gene expression, metabolism and synthesis of antimicrobial compounds, MDR, and expression of efflux pumps [14,17,18]. Whether TipR also has such broad sensory and protective properties has not been explored.

Here, we first report that NAL-based induction of AcrAB-NodT in C. crescentus confers cross-protection against penicillins and cephalosporins, β-lactam antibiotics that target CW biosynthesis enzymes. We exploit this relationship in a deep mutational analysis to dissect the molecular determinants and mechanism responsible for TipR-dependent perception of various novel inducers identified in our chemical screens. We also discovered that TipR not only acts in cis to control its own expression and that of AcrAB-NodT but additionally targets a distant promoter to repress expression of DjlA, a DnaJ-like co-chaperone that interacts with the DnaK protein folding chaperone [19,20]. Although C. crescentus DjlA does not encode a trans-membrane anchor as E. coli DjlA, we found that long-term AcrAB-NodT overexpression perturbs envelope integrity in C. crescentus and that DjlA augments AcrAB-NodT function by interacting with AcrA, thus reconciling the genetically wired co-induction of DjlA and AcrAB via TipR. This circuit design ensures a transient burst of efflux pump and co-chaperone co-expression to prevent efflux pump protein aggregation, while also transiently augmenting efflux capacity in intoxicated cells.

Results

NAL-induced adaptive resistance to β-lactams requires AcrAB-NodT

C. crescentus Δbla mutant cells lack the chromosomally encoded metallo-β-lactamase (MBL) and are, thus, unable to grow on plates supplemented with certain β-lactams, including the cephalosporin cephalothin (CEF, 10 μg/mL, CEF10) or the ureidopenicillin piperacillin (PIR, 40 μg/mL, PIR40). To search for determinants conferring β-lactam resistance in cells lacking MBL, we conducted a transposon (Tn) mutagenesis of Δbla mutant cells, delivering the Tn by intergeneric conjugation from an E. coli donor. Initially, we counterselected the donor on plates containing NAL (20 μg/mL, NAL20), but we unexpectedly observed that Δbla cells grew on PIR40 or CEF10 plates supplemented with NAL20. NAL is not known to affect β-lactam resistance directly, as it targets the A subunit of DNA gyrase (GyrA) in susceptible bacteria. However, since C. crescentus is naturally resistant to NAL owing to the polymorphism F96D in GyrA [9], we speculated that NAL activates a cryptic or an alternative β-lactam resistance mechanism in C. crescentus. To confirm that NAL confers β-lactam resistance in Δbla cells, we conducted Kirby–Bauer-based diffusion assays using β-lactam discs placed on plates with or without NAL20. Indeed, NAL induces resistance of Δbla cells to various cephalosporins (Fig 1B: CEF, cefotaxime and cephalexin) and penicillins (PIR, ampicillin, and amoxicillin), yet it provided little protection against carbapenems (meropenem, imipenem, and doripenem).

To dissect the genetic basis for the NAL-induced resistance to these β-lactams, we mutagenized Δbla cells with a himar1 Tn-conferring kanamycin (KAN) resistance, counterselecting the Tn-delivering E. coli donor cells with colistin or aztreonam (instead of NAL) because C. crescentus is also naturally resistant to these antibiotics. We selected resistant mutants on plates containing KAN20 and PIR40 and mapped the Tn insertions of 2 different clones to the tipR gene (Fig 2A and 2C) [9,13]. Next, we backcrossed the tipR::Tn mutation into Δbla cells and compared the adaptive β-lactam resistance conferred by NAL in Δbla;tipR::Tn double-mutant cells to that of Δbla parental cells by Kirby–Bauer antibiotic disc diffusion assays (Fig 1B), as well as by growth on plates containing CEF10 or PIR40 (Fig 2C). These tests revealed an identical β-lactam resistance spectrum between the Δbla strain grown in the presence of NAL and the Δbla;tipR::Tn strain without NAL, indicating that NAL responsiveness is lost without TipR (Figs 1B and 2C), the transcriptional repressor of PacrA.

thumbnail
Fig 2. Deep mutational dissection of the tipR-acrAB-nodT locus.

(A) Positions of mutation in tipR and the acrAB-nodT operon conferring resistance to cephalothin (10 μg/mL) in Δbla cells. The list of mutations is a combination of sequences from the Mut-Seq experiment and single-clone sequencing. The line represents, from the tallest to the smallest, the loss of start codon, new stop codon, missense mutations, frameshift, and a synonymous variant. (B) Intergenic region between acrA and tipR. The inverted repeats region is a hot spot for mutations that confer resistance to cephalothin (10 μg/mL) to Δbla cells. The * represent SNPs; (α, σ, γ) are double mutations per strain and (i) insertion. Position of the putative −10 and −35 sequences was predicted using SoftBerry. (C) Resistance of Δbla cells and various mutant derivatives to cephalothin or piperacillin. Strains were straked plates containing 10 μg/mL or 40 μg/mL of cephalothin (CEF10 or CEF40), or 40 μg/mL of piperacllin (PIR40) with or without NAL (10 μg/mL, NAL10) and grown for 72 hours at 30°C. Note that ΔABT is used as abbreviation for the ΔacrAB-nodT mutation.

https://doi.org/10.1371/journal.pbio.3002040.g002

RNA-Seq analysis of NAL-induced cells revealed a 20-fold induction of acrAB-nodT transcripts compared to the (uninduced) control state (Figs 1CB and S2A and S1 Data) and a 7-fold induction of tipR transcripts, but also a 2-fold induction of transcripts encoding the DnaJ-like co-chaperone DjlA (CCNA_02218; see below). Moreover, several transcripts encoding components of respiratory chain/cytochrome biosynthesis pathway (CCNA_01767, CCNA_2850, CCNA_3065) were up-regulated between 1.9- and 1.6-fold. Thus, the transcriptional response to NAL predominantly affects the acrAB-nodT and tipR locus. No signature of a general transcriptional response to altered DNA supercoiling or DNA damage was apparent, a response that typically occurs in the presence of certain quinolone antibiotics. We hypothesized that the induction of the djlA transcript could reflect a direct interplay with TipR and AcrAB-NodT (see below), whereas the weak up-regulation of transcripts encoding respiratory proteins likely stems from increased respiratory demand owing to PMF consumption by elevated AcrAB-NodT-mediated efflux arising from TipR derepression.

To confirm that NAL interferes with TipR-mediated repression of PacrA, we conducted β-galactosidase (LacZ)-based promoter probe assays (Fig 1D and S2 Data) in cells transformed with a PacrA-lacZ transcriptional promoter probe plasmid (pPacrA-lacZ). In wild-type (WT) cells, the addition of NAL lead to a 4-fold induction of PacrA-lacZ activity (405%) relative to the reference state without NAL (set to 100%). By contrast, PacrA-lacZ activity was 427% in tipR::Tn cells grown without NAL and remained unchanged upon the addition of NAL. Matching results were obtained by monitoring the accumulation of AcrA by immunoblotting using polyclonal antibodies to AcrA (Figs 1E and S3A). AcrA abundance was strongly induced upon the addition of NAL in WT cells. By contrast, AcrA steady-state levels are elevated in tipR::Tn cells and not further augmented by the addition of NAL. We conclude that NAL induction of PacrA is likely governed through inactivation or removal of TipR. Moreover, and consistent with the induction of tipR mRNA by RNA-Seq, immunoblotting using polyclonal antibodies to TipR revealed a congruent accumulation of TipR and AcrA after 20 minutes during a time course analysis with NAL10 (S3B Fig), suggesting that acrA and tipR promoters are simultaneously derepressed upon the addition of NAL.

Next, we conducted ChIP-Seq experiments using polyclonal antibodies to TipR in pursuit of TipRs promoter occupancy in vivo. We discovered that TipR significantly binds 4 chromosomal targets in vivo. Importantly, TipR occupancy of these sites is reduced by ≈500% after exposing cells to NAL (for 30 minutes; Figs 1F and S2A and S3 Data), indicating that NAL antagonizes repression of PacrA and PtipR by TipR in vivo, allowing recruitment of RNAP (and TrcR [10]). The coordinated induction of AcrAB-NodT and TipR by NAL points towards a homeostatic control mechanism in which the adaptive and transient resistance to β-lactam antibiotics is conferred by induction of AcrAB-NodT, followed by TipR synthesis to reestablish the repressed ground state once the inducers have been expelled by AcrAB-NodT.

In support of this notion, we found that NAL no longer confers adaptive β-lactam resistance in Δbla ΔacrAB-nodT cells (Figs 2C and S1). While this result indicates that the adaptive resistance is mediated by NAL-based induction of AcrAB-NodT, the basal level expression of AcrAB-NodT in the uninduced state suffices to confer resistance to the fourth-generation cephalosporin cefepime. Such residual constitutive resistance by AcrAB-NodT is also detectable for low levels of cephalothin (first generation) and cefotaxime (third generation; S1B Fig). In the case of the former, we asked if AcrAB-NodT could even confer resistance to higher levels of cephalothin than the resistance arising from inactivation of tipR (CEF10). To this end, we conducted a second step selection for CEF40 (cephalothin 40 μg/mL) resistance of Δbla;tipR::Tn cells and isolated several such mutants, each harbouring single missense mutations in acrB (Q177P, L241P, Q528H, P565A, E566K, N676Y, I762F; Figs 2C and S2B). We surmise that these mutations likely increase the specificity, capacity, or abundance of AcrAB-NodT to expel cephalothin (Fig 2C) and that they specifically point to AcrB as a key plasticity determinant enabling high-level MDR and the main element underlying NAL-dependent adaptive resistance (to cephalothin) in C. crescentus.

Ectopic long-term AcrAB-NodT induction perturbs envelope structure and function

Having dissected the role of AcrAB-NodT in adaptive β-lactam resistance, we probed for elevated resistance conferred by the tipR::Tn mutation compared to the parent by Kirby–Bauer-based antibiotic disc diffusion assays using a diverse panel of discs with different classes of antibiotics. We observed increased resistance toward the quinolones sparfloxacin and ciprofloxacin as well as to the macrolide antibiotics erythromycin and azithromycin, likely owing to increased efflux conferred by AcrAB-NodT overexpression (S3 Fig). Surprisingly, these experiments also revealed that tipR::Tn cells are more susceptible CW-targeting antibiotics such as vancomycin, teicoplanin, bacitracin, and fosfomycin (S4 Fig). As the first 3 molecules are large soluble antibiotics that are usually poorly active against gram-negative bacteria due to the OM, which prevents their entry into cells [21], we wondered whether envelope integrity might be compromised in tipR::Tn cells, facilitating the entry of these antibiotics. Indeed, phase contrast microscopy shows frequent envelope blebs and other cell deformations on tipR::Tn cells. While such blebs were seen less frequently when AcrAB-NodT induction occurred through NAL than through inactivation of TipR or through simple AcrAB-NodT overexpression form a plasmid (S4A Fig, red arrows), we suspect that NAL-dependent induction could be a transient phenomenon and/or the induction is attenuated because of the expulsion of NAL by AcrAB-NodT. In support of this idea, we found that the envelope deformations do not require the efflux activity of the pump, as cells exposed to the efflux pump inhibitor 1-(1-Naphthylmethyl)-piperazine (NMP) still harbour these deformations (S5A Fig). Thus, a strong induction and subsequent assembly of this trans-envelope structure is sufficient to disfigure cells.

To investigate the nature of those perturbations in detail, we imaged cells by cryo-electron tomography (cryoET). We imaged Δbla cells overexpressing AcrAB-NodT either via tipR::Tn mutation and via a multicopy plasmid carrying acrAB-nodT (pSRK-acrAB-nodT) coupled with native induction of AcrAB-NodT from the chromosomal locus by NAL10 for maximal overexpression. The cryo-electron tomograms of Δbla cells showed a canonical cell envelope architecture, with a continuous IM, OM, and a paracrystalline surface (S-layer) layer (S5B Fig). In contrast, Δbla;tipR::Tn and Δbla;pSRK-acrAB-nodT cells revealed abnormal cell shapes, manifesting in bulging of the entire cell envelope (S6A Fig). Furthermore, these mutants also showed significant perturbations of the structure of the cell envelope, including membrane blebs and diverse types of vesiculations of the membranes (S5B and S6 Figs). While the precise content and origin of those vesicles remains unknown, we suspect that they contain phospholipid- and/or LPS-derived OM material and/or soluble periplasmic content.

In summary, these structural defects are consistent with increased susceptibility toward antibiotics that are typically kept at bay by the OM, and they highlight the necessity for tight and transiently regulated control by TipR for an efflux system that should be induced only under life threatening conditions.

Genetic dissection of TipR and its target

To dissect the molecular mechanism of induction by TipR, we took advantage of the chemical-genetic relationship between AcrAB-NodT induction and CEF10 resistance. With the goal of discovering loss-of-function mutations in tipR, or potentially its (unknown) target sequence in PacrA, we isolated spontaneous CEF10-resistant point mutants by plating Δbla cells on CEF10 plates. After collecting approximately 10,000 CEF10-resistant Δbla colonies, we PCR amplified tipR and acrAB-nodT from the mutant pool and deep sequenced the PCR products. Sequence analysis (Fig 2A and S4 Data) of this Mut-Seq dataset indeed revealed missense, nonsense, and frameshift mutations scattered throughout tipR, with a slight hotspot in the region predicted to encode the N-terminal DNA-binding domain. The fact that nonsense and frameshift mutations are also found in the C-terminal half of the protein indicates that these residues are also important for TipR function. Interestingly, we also found 14 (gain-of-function) mutations within the acrAB-nodT coding sequence that presumably increase the stability, structure, and/or affinity of AcrAB-NodT components towards CEF (matching the CEF40 experiments described above).

Next, we isolated individual CEF10-resistant Δbla mutant clones from this pool and sequenced their genomes. Three TipR missense mutants were chosen for further study, each encoding a different mutation in TipR:S53R (in the DNA-binding domain), as well as E119V and L204Q, respectively in the central region and C-terminal domain (Fig 2A, corresponding to mutants Δbla;tipR-S53R, Δbla;tipR-E119V, and Δbla;tipR-S53R). We confirmed the strong increase in PacrA activity in all these mutants using pPacrA-lacZ-based assays (S7 Fig and S2 Data).

In addition to these loss-of-function mutations in TipR, we also obtained clones with a mutation in the operator site for the promoter of tipR (PtipR, since tipR and acrAB-nodT are transcribed divergently from the same intergenic region; Fig 2B) that would prevent TipR from binding and repressing PacrA. Since the TipR-binding site sequence is shorter than the 615-nucleotide tipR gene, mutations in the former would likely surface less frequently than nonsense mutations in tipR. Nonetheless, genome sequencing of CEF10-resistant mutants revealed 2 strains with mutations in PacrA (strains Δbla PacrA-RIR and Δbla PacrA-IRup), attesting to the strength and specificity of the selection. To favour the enrichment of additional loss-of-function mutations in the TipR-binding site (Fig 2B), we modified our starting genotype for a revised screen. To this end, we conducted the CEF10-resistant selection with Δbla cells carrying a multicopy plasmid expressing TipR (pSRK-Gm-tipR) to eliminate mutants with a dysfunctional tipR gene that would otherwise surface. With this strategy, we uncovered strains with additional PacrA mutations, 80% of which reside in a region of dyad symmetry comprising an 11-base pair inverted repeated (IR, 5′-TGAGAATGAAC-3′; Fig 2B), 2 adjacent mutations are 11 nucleotides upstream of the IR (IRup mutation), and another mutation lies 33 nucleotides downstream of the IR (mutation IRdw).

Lastly, we chose the following IR mutant strains for follow-up confirmation studies: Δbla PacrA-MIR (middle inverted repeat), Δbla PacrA-RIR (right inverted repeat), Δbla PacrA-2IR (2 mutations inverted repeat) and Δbla PacrA-IIR (insertion inverted repeat; Fig 2B), as well as one with a mutation upstream of the IR (Δbla PacrA-IRup) and one downstream of it (Δbla PacrA-IRdw; Fig 2B). Immunoblotting using polyclonal antibodies to AcrA and TipR confirmed that AcrA is overexpressed in all selected mutants compared to the Δbla parent (S8A–S8D Fig), explaining why these mutants grew on CEF10 plates. TipR was also overproduced in all mutants to similar levels, except for strains Δbla PacrA-IRup and Δbla tipR-L204Q (S8A and S8B Fig) in which the TipR levels resemble the basal amount of the Δbla parental strain. The addition of NAL led to the induction of TipR in these 2 mutants, albeit to lower steady-state levels compared to Δbla cells, whereas AcrA levels were not substantially affected. This result suggests that the effect of the mutation differs for PacrA versus PtipR and/or that NAL affects the abundance of TipR through a post-translational mechanism. Below, we provide additional evidence for the latter model.

In summary, the strong selection for CEF10 resistance in Δbla cells unveiled mutational hotspots in the region predicted to be required for DNA binding by TipR, but also in the IR of in the suspected promoter target of TipR.

Defining the TipR target sequence and functional residues in TipR

We used electrophoretic mobility assays (EMSAs) to confirm that TipR binds the IR in vitro. In these experiments, a Cy5-labelled PacrA probe centred on the IR was mixed with recombinant (untagged) TipR purified in 2 steps from an E. coli overexpression strain (see Experimental procedures). We observed a retardation of the labelled probe by TipR in a concentration-dependent fashion, even in the presence of nonspecific calf thymus competitor DNA (Fig 3A). EMSA experiments with E. coli extracts containing WT TipR also revealed retardation of Cy5-labelled PacrA compared to E. coli extracts without TipR. Next, we probed E. coli extracts each containing a different TipR mutant described above (E119V, S53R, or L204Q) for binding of PacrA by EMSA. While the L204Q variant retards PacrA akin to WT TipR, the E119V or S53R mutations abolish PacrA binding (Fig 3A). Immunoblots conducted with C. crescentus cell extracts revealed the E119V and S53R variants to accumulate to the same steady-state level as WT TipR upon NAL induction (S8A Fig), indicating that the E119V and S53R mutations promote DNA binding. By contrast, the abundance of TipR-L204Q is severely reduced, even in the presence of NAL (S8A Fig), suggesting that this mutation interferes with protein stability rather than with DNA-binding activity as shown by antibiotic chase experiments, revealing a reduced half-life of TipR-L204Q (Fig 3B).

thumbnail
Fig 3. Determinants in TipR and PacrA required for repression.

(A) EMSA with 200 ng of a Cy5-labelled acrA promoter (PacrA) probe in the presences of increasing concentration of cell extract from E. coli overexpressing different mutated TipR protein from 0.031 μM to 16 μM. (B) Immunoblot showing the result of a 1-hour chase experiment using chloramphenicol (5 μg/mL) on exponential NA1000 (WT) and Δbla;tipR-L204Q cells in PYE. (C) EMSA with 200 ng of different Cy5-labelled PacrA mutants analysed for retardation by gradual concentration of purified TipR protein from a concentration of 0.0156 μM to 16 μM. (D) EMSA with 100 ng of Cy5-labelled acrA with gradual concentration of heparin-purified TipR protein from a concentration of 0.25 μM to 8 μM. (E) EMSA with 100 ng of Cy5-labelled djlA, ccrM, and acrA promoters (PdjlA, PccrM, and PacrA) with gradual concentration of purified TipR protein at a concentration of 1.125 μM to 18 μM. (F) Activity of β-galactosidase of the PdjlA-lacZ or PccrM-lacZ in WT C. crescentus. The « + » indicates induction by Nal (10 μg/mL) for 2 hours. All levels are indicated in percentage of expression regarding the basal level of NA1000 (WT) without induction. The data from the analysis are deposited in S2 Data. EMSA, electrophoretic mobility assay; PYE, peptone-yeast extract; WT, wild-type.

https://doi.org/10.1371/journal.pbio.3002040.g003

Having identfied determinants in TipR that promote DNA binding, we next determined the effects of the isolated promoter mutations. To this end, PacrA promoter mutants with alterations in the IR were labelled with Cy3 and incubated with purified WT TipR in EMSAs (Fig 3C). While PacrA-RIR is no longer retarded by TipR in EMSAs, we only observed a partial loss of binding by TipR to PacrA-MIR. As expected, TipR still retarded the PacrA-IRup probe that carries a mutation outside the IR. Thus, these EMSAs validate the requirement of the IR as the likely TipR docking site in PacrA. With these results, we investigated further how PacrA-IRup can cause overexpression of the pump without crippling DNA binding of TipR. EMSAs with a new Cy5-PacrA probe combining the PacrA-RIR mutant (that abolish binding of TipR) with the PacrA-IRup mutation, EMSA (Fig 3D) revealed that PacrA-IRup increased the affinity of TipR for the PacrA-RIR probe, suggesting that PacrA-IRup mutation creates a secondary docking site upstream of the IR.

Indeed, the consensus target sequence derived for TipR (S9 Fig) resembles the motif created by the PacrA-IRup mutation. In search for additional binding sites of TipR, we conducted ChIP-Seq experiments (S3 Data) and found that in vivo TipR binds 4 sites on the chromosome: the promoter of acrAB-nodT and tipR, the promoter of the djlA gene predicted to encode a DnaJ-like co-chaperone, as well as a site upstream of the DNA methyltransferase gene ccrM and another site in between the qor and rho genes oriented towards each other. The RNA-Seq experiments described above confirmed the NAL induction of 3 genes neighbouring two of the TipR binding sites (acrAB-nodT, tipR, and djlA), and ChIP-Seq experiments conducted with NAL-treated cells revealed that the binding of TipR to all 4 chromosomal sites is impaired in vivo (Figs 1F and S10). However, the binding of TipR to PacrA was approximately 250% more efficient than to the other sites, suggesting that the IR in PacrA represents the preferred recognition sequence of TipR. Nonetheless, we scanned the other target sites for sequences comparable to the IR and detected one upstream of djlA and of ccrM, delivering the consensus sequence 5′-WTGaGWMtGAWC-3′ by MEME analysis (S9 Fig). Using this consensus, we analysed the intergenic sequence between the qor and rho genes and found a single sequence that could explain the presence of TipR at this location; however, the configuration of those genes facing each other makes a TipR-dependent transcriptional regulation of these 2 genes unlikely.

Next, we tested the binding ability of TipR to djlA and ccrM promoters in vitro by EMSA using purified recombinant TipR. As seen in Fig 3E, TipR was able to retard the djlA promoter, but not the promoter of ccrM. Moreover, LacZ-based promoter probe assays with PdjlA-lacZ and PccrM -lacZ reporters revealed an induction of PdjlA in WT cells in response to NAL, but no significant induction of PccrM-lacZ (Fig 3F and S2 Data). Since TipR binds PccrM weakly in vivo, likely because of the divergence of its IR from the preferred TipR target sequence, the weak interaction might result in a TipR-promoter complex too fragile to be maintained during electrophoresis. Moreover, in the absence of a significant induction of transcription by NAL in our RNA-Seq experiment (Fig 1C), we speculate that TipR is a poor repressor of PccrM in vivo, yet it clearly down-regulates PdjlA.

Screening for chemical inducers of PacrA

Having identified the IR of PacrA as the target of TipR’s DNA-binding activity, we engineered a reporter for rapid screening of chemical libraries for compounds that induce PacrA (independent of requirement of AcrAB-NodT efflux function as in CEF10 resistance screen above). To this end, we fused to the promoterless nptII kanamycin resistance gene to PacrA, creating a convenient readout for molecules like NAL that can activate PacrA in vivo scored as kanamycin resistance. We transformed the resulting PacrA-nptII promoter probe reporter plasmid (pPacrA-nptII) into WT cells, embedded the resulting reporter cells in soft agar containing kanamycin (10 μg/mL, KAN10) overlaid on KAN10 plates, and then spotted 4 μL drops of each compound (at 10 μM) of the Maybridge chemical library for inducers of PacrA on these seeded indicator plates. Small molecule inducers were identified by their ability to promote growth of cells around the disc owing to activation of PacrA-nptII by the compound(s) (S11 Fig). This chemical screen along with tests of candidate compounds reported in the literature for other systems, unearthed 14 different inducers of PacrA-nptII (Figs 4A and S11) that were subsequently grouped into 2 classes based upon pPacrA-lacZ inducer strength as determined by LacZ activity measurements. One group of strong inducers (triggering PacrA-lacZ activity exceeding 230% activity relative to WT) includes the quinolones NAL, flumequine (FLU), and sparfloxacin, quinolone precursors such as chloroquine and dyes such as rhodamine 6G (Rh6) and crystal violet (CV). The weak inducers include the dye malachite green (MG), DNA intercalants such as acridine orange (AO), ethidium bromide (EtBr), and SYBR safe, but also various other compounds such as glafenine and doxylamine. We also noticed that dopamine and aminolevulinic acid fail to induce PacrA-lacZ after 3 hours; however, they allowed growth on kanamycin plates with PacrA-nptII after overnight incubation (Figs 4A and S11 and S2 Data), suggesting that they act slowly, through another pathway, or that they counteract kanamycin in an alternative way.

thumbnail
Fig 4. Different chemicals relieve repression of PacrA by TipR.

(A) Expression of β-galactosidase assayed from PacrA-lacZ in NA1000 (WT) after 3 hours of induction by different compounds. All concentrations are in μg/mL. All levels are indicated as percentage of expression regarding the basal level of the control: NA1000 (WT) without induction. The data from the analysis are deposited in S2 Data. (B) Immunoblot revealing the abundance of TipR during a 2-hour chase experiment with chloramphenicol (5 μg/mL) added to exponential tipR::Tn;pSRK-tipR cells grown in PYE or and without selected inducers: Nal (10 μg/mL), Flu (10 μg/mL), Rh6 (2 μg/mL), CV (2 μg/mL), EtBr (2 μg/mL), and MG (2 μg/mL). (C) EMSA using 200 ng of Cy5-tagged DNA of the promoting region of acrA with 8 μM of heparin-purified TipR, supplemented with an increasing concentration of selected inducers. All quantities are in μg/mL. (D) β-galactosidase measurements in C. crescentus WT and different mutants containing PacrA-lacZ before or after induction with Nal (10 μg/mL), Rh6 (Rho, 2 μg/mL Rho), CV (2 μg/mL), and MG (2 μg/mL) for 2 hours. All measurements are expressed as percentage of expression relative to basal activity of the control: WT or ΔsocB cells before induction. The data from the analysis are deposited in S2 Data. (E) Immunoblot analysis of cell extracts from different protease mutants performed with polyclonal antibodies to TipR and to AcrA. All inductions (+) were performed after 2 hours of induction with Nal (10 μg/mL). Shown as loading control are blots with polyclonal antibodies to CCNA_00163. (F) Immunoblot following a 2-hour chase experiment with chloramphenicol (5 μg/mL) on exponential cells of WT with a clpP* inserted at the Pxyl locus. Cells were grown in PYE, followed by the addition of glucose or xylose, and, subsequently, the inducer was added: Nal (10 μg/mL), and, finally, protein synthesis was blocked with chloramphenicol. CV, crystal violet; EMSA, electrophoretic mobility assay; EtBr, ethidium bromide; Flu, flumequine; MG, malachite green; Nal, nalidixic acid; PYE, peptone-yeast extract; Rh6, rhodamine 6G; WT, wild-type.

https://doi.org/10.1371/journal.pbio.3002040.g004

Some of the identified molecules are known substrates of the enterobacterial AcrAB-TolC RND pump and would thus likely be exported to reduce the active concentration in cells. Hence, we wondered if induction of PacrA would be accentuated in cells lacking AcrAB-NodT (S12A Fig and S2 Data) that can no longer expel these molecules, resulting in their accumulation to higher intracellular levels. While the level of PacrA induction with compounds like NAL or EtBr was similar in WT versus ΔacrAB-nodT cells, AO clearly showed a 4-fold elevated level of induction compared to the 2-fold induction in WT cells. Thus, the intracellular concentration modulates the level of acrAB-nodT expression. To eliminate the possibility that the compounds cause a nonspecific global increase of gene expression or LacZ enzymatic activity, we tested Pbla-lacZ reporter expressing LacZ expressed from the constitutive promoter of the C. crescentus MBL gene CCNA_02223 and found that these compounds did not alter LacZ activity from Pbla-lacZ (S12B Fig and S2 Data).

Different inducers control TipR at the post-translational level

To determine if the new inducers act in the same way on TipR as NAL, we first explored if they affect TipR stability in vivo. To this end, we uncoupled TipR synthesis from its inducible transcription by expressing it from the IPTG-controllable Plac promoter (of E. coli) on plasmid pSRK-tipR in tipR::Tn cells. AcrAB-NodT is still inducible from PacrA in these cells, but TipR synthesis is independent from the tester compounds such as NAL, Rh6, and MG (S13A and S13B Fig). We then compared the half-life of TipR in these cells after exposing them to the chemical inducers. To stop translation of TipR, we treated cells with high levels of the protein synthesis inhibitor chloramphenicol, a phenicol antibiotic that is (likely) not a substrate of the AcrAB-NodT efflux pump. In the presence of the quinolones NAL or FLU, TipR is rapidly turned over (Fig 4B), reminiscent of the instability observed for TipR-L204Q. Such an induced instability was not observed with the other PacrA-activating compounds.

Next, we examined the effect of these compounds on the DNA binding ability of purified TipR by EMSA (Fig 4C). While at physiologically relevant concentrations, NAL or FLU did not perturb retardation of the Cy-labelled PacrA, other inducers such as EtBr or MG interfered with binding of TipR to PacrA. Rh6 and CV affect TipR binding in the same manner, but less efficiently in vitro compared to EtBr or MG. To assess the specificity of these compounds on TipR, rather than interfering with protein–DNA interactions in a general fashion, we conducted control EMSA using the histone-like IHF protein and its target DNA sequence attR (S14 Fig) as probe. Increasing concentration of CV did not prevent IHF binding even at a concentration sufficient to change the charge of the probe resulting in its upward migration in the gel during electrophoresis. Therefore, CV, Rh6, EtBr, and MG apparently interfere with TipR’s ability to bind the IR in PacrA, explaining the derepression of PacrA by these compounds in vivo.

To rule out the possibility that induction of AcrAB-NodT by NAL or FLU acts through the SOS (DNA damage) response by corrupting DNA gyrase (encoded by gyrAB) as in E. coli, we confirmed that inhibition of GyrA by the fluoroquinolone ciprofloxacin (CIP) or of GyrB by the aminocoumarin novobiocin (NOV) alone did not induce of PacrA (S15 Fig and S2 Data). Moreover, NAL and CIP did not synergise to enhance PacrA induction, yet NAL and NOV together resulted in elevated PacrA-lacZ induction. Consistent with these findings, we next assayed PacrA-lacZ activity in WT cells expressing an additional GyrA variant: GyrA from Brucella melitensis or a mutant variant of C. crescentus GyrA, GyrA*(F96D) that is NAL sensitive and causes DNA damage and the SOS response in the presence of NAL. C. crescentus WT cells expressing either of these GyrA variants from pMT335 still induce PacrA upon the addition of NAL (S15 Fig and S2 Data), even more efficiently than WT cells. In fact, the level of NAL induction in cells expressing the NAL-sensitive forms of GyrA attains the level of induction when NAL and NOV are added jointly to WT cells, indicating that Gyrase inhibition can enhance AcrAB-NodT induction, but only after repression of PacrA by TipR has been relieved by the addition of NAL.

Prompted by the observation that, at physiological concentrations, NAL and FLU reduce the half-life of TipR without interfering with TipR binding to PacrA in vitro in EMSAs (Fig 4B), we speculated that a transient destabilization by a dedicated protease could account for the induction of AcrAB-NodT. Using a candidate approach, we then tested for protease(s) conferring instability of TipR by immunoblotting using polyclonal antibodies to TipR and AcrA (S16 Fig). While no significant changes in AcrA or TipR steady-state levels was observed in ftsH and lon mutant cells compared to WT cells, loss of ClpP or ClpX led to an accumulation of AcrA and a reduction in TipR levels in the absence of NAL. Indeed, previous work showed that AcrA and AcrB are substrates of the ClpP degradation machinery [22]. Yet, AcrA and TipR are still inducible upon exposure of clpP and clpX mutant cells to NAL (Figs 4E and S13), but promoter probe assays using the PacrA-lacZ reporter revealed a strong reduction of LacZ induction when clpP and clpX mutant are exposed to NAL, Rh6, CV, or MG compared to WT cells (Fig 4D and S2 Data). Taken together, our findings suggest that the ClpXP protease promotes high level of induction of acrAB-nodT and tipR, whereas inactivation of the clpA chaperone gene in WT cells did not cause a major AcrA or TipR abundance before after induction with NAL (Fig 4E).

To test if ClpP controls the half-life of AcrA and TipR, we conducted antibiotic chase experiments in cells expressing a dominant negative version of ClpP (ClpP*) in which the catalytic serine is mutated to alanine (Figs 4F and S11A). This ClpP* variant is expressed from the xylose-inducible Pxyl promoter at the xylX locus (xylX::Pxyl-clpP*). Immunoblotting using antibodies to TipR and AcrA revealed that TipR is turned over normally upon the addition NAL in the presence of glucose (to repress expression of ClpP* from Pxyl). However, when cells are grown in PYE containing xylose to induce of the dominant negative ClpP* variant), TipR stability is increased. Lastly, to demonstrate the interaction of TipR and AcrA with ClpXP biochemically, we used a GFP Trap matrix to pull down a ClpX-YFP fusion protein from lysates prepared from WT xylX:.Pxyl-ClpX-YFP cells before or after induction with NAL. Immunoblotting revealed the presence of TipR and AcrA in the pulled down material (S17 Fig) regardless of the presence of NAL.

In summary, while some compounds dislodge TipR from the IR sequence, others like first-generation quinolones act to destabilize TipR. The ClpXP protease confers partial instability of TipR in the presence of NAL, and it also reduces TipR and AcrA steady-state levels in the absence of NAL. Lastly, we discovered that ClpXP is required for efficient transcriptional induction of acrAB-nodT.

Co-induction and interaction of DjlA and AcrAB-NodT

Since DjlA, a putative DnaJ-like co-chaperone, is co-induced with AcrAB-NodT and TipR, we investigated whether DjlA interacts with TipR or AcrAB-NodT or affects their function. We first examined whether DjlA overexpression alters induction of PacrA or TipR and AcrA steady-state levels upon addition of NAL. Neither PacrA-lacZ-activity nor AcrA or TipR abundance was affected as a function of DjlA, regardless of whether NAL was present or not (Figs 5A and S18 and S2 Data). However, when we probed for AcrAB-NodT-mediated efflux activity, we found that ectopic expression of DjlA enhances resistance to several β-lactam antibiotics belonging to the cephalosporin and penicillin substrates of AcrAB-NodT (S19 Fig). To assess if DjlA also promotes AcrAB-NodT-dependent efflux of other substrates, we conducted survival (efficiency of plating (EOP)) assays on plates containing EtBr6 (EtBr, 6 μg/mL) and found that cells ectopically expressing DjlA (from pMT335-djlA) exhibited an elevated EOP compared to WT cells with the empty vector (Fig 5B). The positive effect of DjlA was also observed by a reduction of ΔdjlA cells, survival on EtBr4 (EtBr, 4 μg/mL) plates. This effect was restored in the presence of a complementing plasmid (pMT335-djlA) but completely lost when AcrAB-NodT was inactivated (i.e., in ΔacrAB-nodT cells; Fig 5B and 5C). Together, these results support the view that the activity or assembly of the pump, and thus the ability to expel EtBr, is enhanced by DjlA. Interestingly, forcing overexpression of AcrAB-NodT in ΔdjlA cells by introducing the tipR::Tn mutation or by the addition of NAL can mitigate the reduced EOP on EtBr4 plates, likely because AcrAB-NodT levels are no longer limiting for expulsion of EtBr to enable growth.

thumbnail
Fig 5. Role of DjlA co-chaperone in the TipR-dependent control loop.

(A) Measurement of β-galactosidase expressed from PacrA-lacZ in different C. crescentus ΔdjlA mutants before or after a 2-hour induction with Nal [10 μg/mL (+) and/or vanillate 100 μM (V)]. All measurements are indicated as percentage relative to the basal activity of the WT control before induction. The data from the analysis are deposited in S2 Data. (B) EOP assay determined by 10-fold serial dilutions of ΔdjlA::pNPTS and ΔacrAB-nodT mutants on plates containing EtBr (6 μg/mL) to probe for efflux pump activity. All strains contain pMT335 plasmid or a derivative (pMT335-djlA) expressing DjlA, grown on Van at 100 μM and Gent 1 μg/mL. (C) EOP assay by 10-fold serial dilutions of ΔacrAB-nodT and ΔdjlA::pNPTS mutants on plates with EtBr (4 μg/mL) to probe for pump efficiency. Nal induction was performed at 10 μg/mL, Van at 100 μM, and Gent in μg/mL. (D) ChIP-Seq analysis of DjlA-HA expressed from the xylX locus in WT cells before and after NAL induction. DjlA-HA was pulled down using an anti-HA affinity matrix. The acrA-tipR promoter region was the only region showing a >2.9-fold increase in abundance compared to the control ChIP-seq performed with cells that were treated with NAL for 30 minutes or not treated. The data from the analysis are deposited in S3 Data. (E) Immunoblot analysis of aggregated cellular proteins in various strains: A = ΔdjlA::pNPTS, B = WT, and C = WT pMTdjlA. Inductions (+) were performed for 2 hours with Nal (10 μg/mL). The total fraction contains the cell lysate. The insoluble fraction corresponds to the material that remained in the pellet after several extractions with 1% Triton X-100. Strain C was grown in PYE with Gent (1 μg/mL) and Van (100 μM). (F) AcrAB-NodT enables growth of C. crescentus in the vicinity of an unknown fungus. C. crescentus strains grown in contact with an unknown fungus on a PYE plate. The fungus was pregrown on the plate for 2 weeks, after which the C. crescentus strains were streaked from the edge of the plate towards the centre until in contact with the fungus. The plate was then incubated at 30°C for another 72 hours. The fungus had been isolated in the laboratory as a contaminant on a PYE plate. AcrAB-NodT encoded in the C. crescentus chromosome protects against (a) compound(s) produced by the fungus to enable growth of C. crescentus in the vicinity of the fungus. (G) Proposed model of the acrAB-nodT operon and djlA gene regulation, induction and function. EOP, efficiency of plating; EtBr, ethidium bromide; Gent, gentamicin; Nal, nalidixic acid; PYE, peptone-yeast extract; Van, vanillate; WT, wild-type.

https://doi.org/10.1371/journal.pbio.3002040.g005

Since co-chaperones like DjlA typically assist in protein folding and/or disaggregation using the conserved DnaK chaperone, we constructed 2 DjlA variants with mutations in the HPD domain (H187A and H187Q) that is necessary for proper DnaK ATPase activity. Those 2 constructs were put under the control of the inducible Pvan promoter on the pMT335 plasmid and transformed in strains lacking the endogenous djlA gene. The complementation assay performed on EtBr6 agar plate demonstrates that the unmodified allele can complement the absence of the chromosomally encoded DjlA, while neither DjlA H187A or H187Q restore the phenotype (S20 Fig). Hence, the DnaK proper function is fully required for DjlA co-chaperone activity.

We reasoned that co-induction of DjlA and AcrAB-NodT by NAL may serve to manage a stress condition (for the envelope) imposed by the massive synthesis and assembly of the AcrAB-NodT structure, perhaps facilitating disaggregation of AcrA before secretion from the cytoplasm where DjlA is located. In support of this idea, pull-down experiments (S21A Fig) using a DjlA variant with a C-terminal HA-tag (DjlA-HA) revealed that DjlA directly or indirectly associates with AcrA and TipR. We conducted a similar experiment using the DjlA-H187A-HA variant in which a key residue affect the DnaK chaperone cycle is inactivated and found that the interaction of this DjlA-HA variant with AcrA and TipR still occurs, suggesting that the interaction of DjlA with AcrA and TipR does not depend on DnaK function (S21B Fig).

The interaction between DjlA-HA and TipR was further confirmed in ChIP-Seq (chromatin immunoprecipitation) experiments (Figs 5D and S22) revealing DjlA-HA can associate with TipR at the PacrA chromosomal site in vivo following exposure to NAL. In these experiments, crosslinked chromatin was prepared from DjlA-HA expressing cells after NAL treatment, followed by immunoprecipitation of DjlA-HA with bound chromatin using an anti-HA affinity matrix. Deep sequence analysis of the precipitated chromatin revealed a near 3-fold increase in PacrA abundance when DjlA-HA was immunoprecipitated from NAL-treated cells versus untreated cells. No association with PacrA was observed in control ChIP-Seq experiments with other HA-tagged DnaJ-like proteins. Moreover, the specific occupancy of DjlA-HA at PacrA required the presence of TipR (S22 Fig), as revealed by ChIP-seq experiments conducted in which tipR::Tn cells expressing DjlA-HA from the xylX locus had been treated with NAL. This result indicates that DjlA-HA specifically occupies PacrA, that this association requires TipR, and that it is enhanced by NAL.

Finally, knowing that DjlA can interact with TipR and AcrA, we investigated whether DjlA also affects aggregation of TipR or AcrA. To test this, we first induced WT cells, ΔdjlA cells, and DjlA-overexpressing cells (WT+ pMT335-djlA) with NAL and then separated extracts into soluble fractions and those that remain insoluble in Triton X-100 by centrifugation (Figs 5E and S23). Immunoblotting revealed approximately 40% less aggregated AcrA in the insoluble fraction from WT cells overexpressing DjlA after NAL10 induction, while conversely, there was a 40% increase of AcrA aggregates induced by NAL10 in ΔdjlA cells. By contrast, no TipR was detectable in the insoluble fraction. To determine if DjlA also acts on other (envelope) proteins, we conducted LC-MS/MS (liquid chromatography followed by tandem mass spectrometry) analyses of the DjlA-HA pull-downs described above (S24 Fig). Pull-downs of DjlA-HA expressing cells grown in the absence of NAL revealed 8 of the 14 most expressed TonB-dependent receptors (TBDRs) and porins destined for the OM as clients of DjlA. In the presence of NAL10, there was a drastic reduction in the number of OM proteins (OMPs) with a concomitant increase in abundance of AcrA, AcrB, and NodT, suggesting that DjlA has higher affinity for efflux pump components than for OMPs, which is consistent with its role in enhancing efflux activity by preventing its aggregation.

Discussion

Here, we elucidated an integrated tripartite homeostatic loop (Fig 5G) controlling the transient induction of the AcrAB-NodT efflux pump in C. crescentus, the TipR repressor, and the DjlA co-chaperone from promoters that are TipR targets. We show that AcrAB-NodT induction confers adaptive resistance to β-lactam antibiotics from its TipR-repressed promoter that can be triggered by different classes of chemicals: Some are antibiotics, while others are general noxious molecules. Our work reveals 2 mechanisms controlling TipR derepression, with one class of compounds interfering with DNA binding of TipR, while others act on TipR stability. Importantly, we uncovered corresponding mutations in TipR by forward genetics that recapitulate these 2 divergent induction mechanisms. Our deep mutational scanning using forward selection for β-lactam resistance reinforced these findings and clearly unveiled 3 major functional domains in the TipR primary structure (Fig 2A): the N-terminal DNA binding, the central domain, and the extreme C-terminus for (de)stabilization. Mutations in each of these domains can lead to derepression of PacrA with overexpression of AcrAB-NodT and cephalosporin resistance. Lastly, our 2-step antibiotic selection regime and Mut-Seq showed that a second (elevated) level of resistance can still be acquired in tipR mutant cells that overexpress AcrAB-NodT by missense mutations in AcrB, augmenting cephalothin resistance to 3-orders of magnitude compared to regular AcrAB-NodT overexpression strains.

Although inducible overexpression of AcrAB-NodT is critical for survival against certain drugs or toxic molecules (also against natural molecules; Fig 5F), it also comes at a cost for envelope integrity and (cell shape) homeostasis, as well as OM barrier function. We discovered that long-term overexpression of the AcrAB-NodT envelope-spanning system results in strong OM perturbations that can sensitize cells towards antibiotics such as the CW-targeting antibiotic vancomycin that normally cannot traverse the OM. This barrier function is apparently compromised by excessive and/or long-term efflux pump (activity), perhaps owing to the OM detachment from other envelope layers membrane blebs and periplasmic vesicles that occur as revealed by cryoET. Such transient alterations from AcrAB overexpression could cause a discontinuous OM or openings through which drugs like vancomycin can subsequently enter. Thus, to minimize envelope integrity defects caused by long-term expression of AcrAB-NodT, cells must ensure that its expression is transient and that the repressed ground state can reestablish as quickly as possible after induction. Our findings in C. crescentus imply that clinical isolates whose MDR phenotype arises from massive overexpression of efflux pumps may also experience OM integrity defects, at least in the presence of antibiotics that induce the efflux pumps. In this case, such MDR isolates will be sensitized towards antibiotics like vancomycin when administered with inducers of efflux pumps.

The genetic linkage of the cytoplasmic co-chaperone DjlA with AcrAB-NodT and TipR is further reinforced by the biochemical interactions. Yet, we also found that DjlA also binds several TBDRs and other OMPs (see S24 Fig). Yet, upon AcrAB-NodT induction by NAL, DjlA is drawn to the preferred AcrAB-NodT substrate, in favour over other OMPs. Since DnaJ-type co-chaperones can sometimes (partially) substitute for one another [23,24], it is possible that AcrAB-NodT secretion or assembly depends on other co-chaperones. It is conceivable that such possible dependence of AcrAB-NodT on such co-chaperones is magnified in the absence of DjlA. Interestingly, DjlA has been implicated in promoting trans-envelope assembly and the envelope stress response in Legionella pneumophila and in E. coli, respectively [25,26]. Both these DjlA orthologs feature an N-terminal transmembrane domain, whereas no such segment is recognizable in C. crescentus DjlA [27]. Nonetheless, C. crescentus DjlA is shorter than typical γ-proteobacterial DjlA, and it is possible that it relies on another membrane association mechanism to capture and deliver its preferred substrates. The substrate preference switch of C. crescentus DjlA towards AcrAB-NodT after NAL induction may accentuate the OM problems arising from extended efflux pump overexpression and assembly, perhaps leading to a reduction of insertion of OMPs that tether the OM to the CW and resulting in membrane blebs. In principle, we cannot rule out that blebbing arises from excessive translocation of phospholipids from the IM to the OM upon AcrAB-NodT overproduction, creating an overload that could lead to the OM invaginations or export of another substrate from the IM that affects OM tethering.

Our finding that the TipR is also induced by NAL, along with AcrAB-NodT and DjlA, is consistent with the view that an underlying tripartite homeostatic control loop sets the stage for reestablishing repression, for example, as soon as the inducing compound is expelled from the cytoplasm (Fig 5G). While repression by TipR will then terminate transcription of acrAB-nodT, the cell must also reduce the levels of the induced AcrAB-NodT protein by proteolysis rather than by dilution due to cellular division, which is substantially slower than proteolysis. Indeed, AcrA is substrate of the ClpXP protease [22], and our findings provide further evidence for interactions between TipR, AcrA, and the degradasome. The ATP-dependent protease ClpP restrict the passage to its proteolytic chamber only small, unfolded peptides, through narrow pores. To degrade bigger molecules, the proteasome needs ClpX that recognize specific substrates, directly or using adaptors to unfold the targeted protein using ATP. This then allows the denatured polypeptide chain to be translocated to the ClpP inner chamber where it gets degraded. One explanation could be that quinolones recruit an adaptor to TipR, or the antibiotic molecules block proper folding of the repressor, rendering them recognisable by the ClpXP machinery.

In the absence of a known natural inducer, we sought and discovered 14 mostly synthetic molecules that trigger derepression of TipR. Among the inducers described in this study, 2 molecules possess a curious behaviour: dopamine and aminolevulinic acid that do not induce AcrAB-NodT expression in the short term. While aminolevulinic acid can be used as a natural heme precursor, dopamine is a catecholamine characterised by an iron-binding domain. Both molecules are highly unstable and change conformation after time, in particular aminolevulinic acid that spontaneously dimerises into porphobilinogen, pseudoporphobilinogen or 2,5-dicarboxyethylpyrazine forming non-heme precursors [28]. As those molecules share structural similarities with efflux pump substrates, we speculate that aminolevulinic acid is not an inducer in this form, but its degradation or dimer products might trigger expression of the efflux pump, perhaps through an iron stress response. In support of this idea, our RNA-Seq analysis of NAL-treated cells shows an increase of expression in the transcript encoding the 5-aminolevulinic acid synthase, hinting at a purpose of acrAB-nodT expression control and inducers, with aminolevulinic acid derivatives activating the AcrAB-NodT expression to export an unusable product from the cell.

AcrAB-NodT likely exports natural products and metabolites upon induction with AcrB acting as key linchpin [29], a finding also implied from or CEF40-resistant acrB alleles. Chemical warfare between microbes typically involves a myriad of toxic molecules, sometimes targeting different compartments, and, thus, protection against many molecules at once can be signalled through a single inducer entering the cytoplasm. Indeed, an unidentified fungus isolated in the lab prevents growth of C. crescentus cells lacking AcrAB-NodT in proximity, but not cells expressing it (Fig 5F), thereby illustrating the importance of AcrAB-NodT in protecting bacteria towards natural toxic metabolites that likely also induce the system. However, we also noted that the contribution of AcrAB-NodT-mediated efflux is also detectable without NAL induction in assays of WT and for ΔacrAB-nodT cells using discs containing the cephalosporins cefepime or cefotaxime (S1B Fig). AcrAB-NodT also confers resistance to other antibiotics including erythromycin or other noxious compounds such as EtBr in the uninduced state, indicating that AcrAB-NodT also expels compounds other than β-lactams. β-lactams do not induce AcrAB-NodT expression, a circumstance that is likely explained by the fact that these drugs are not known to enter the cytoplasm, unlike smaller molecules like EtBr or NAL. While this property explains why resistance to β-lactams in C. crescentus is an adaptive trait, it may also confer multiresistance to a wide range of toxic molecules with a minimal range of inducers of the efflux pump.

Collectively, our findings explain why AcrAB-NodT control on at least 2 levels (TipR and DjlA) is critical for transient efflux activation that protects bacterial cells in the wild from noxious molecules, but also from antibiotic treatment in the clinical setting.

Materials and methods

Growth conditions

All C. crescentus strains were cultivated in peptone-yeast extract (PYE) and incubated at 30°C. ФCr30-mediated generalized transductions were done as described [30]. E. coli strains were grown in lysogeny broth at 37°C. All media were supplemented with the appropriate antibiotics or indicated inducer. The strain used for cloning is E. coli EC100D, while the strain used for protein expression is E. coli BL21(DE3). As both WT (NA1000) and Δbla C. crescentus cells are naturally resistant to colistin (COL, 4 μg/mL) or aztreonam (AZT, 3 μg/mL), COL or AZT was used to counterselect E. coli Tn delivery strains for Tn (himar1) mutagenesis encoded on plasmid pHPV414 [31]. The himar1 encodes resistance to kanamycin, and, therefore, the PYE plates were supplemented with kanamycin (20 μg/mL, KAN20) and CEF10 or PIR40. Strains, plasmids, primers, and synthetic genes are detailed in S6 Data.

β-galactosidase assays

All β-galactosidase assays were done using freshly electroporated strains harboring the pLac290-derived promoter probe plasmids to quantify the activity of the transcriptional fusions between the studied promoters and lacZ. The assays were done at room temperature. Between 50 and 100 μL of cells (OD600nm of 0.3 to 0.7) were lysed in 30 μL of chloroform and vigorously mixed with Z buffer (60 mM Na2HPO4; 40 mM NaH2PO4; 10 mM KCl and 1 mM MgSO4 (pH 7)) to obtain a final volume of 800 μL. Followed by the addition of 200 μl of ONPG (Ortho-nitrophenyl-β-D-galactopyranoside, at 4 mg/mL in 0.1 M potassium phosphate (pH 7)) to begin the reaction. Assays were stopped using 500 μL of 1 M Na2CO3 when the solution turned light yellow. The OD420nm of the supernatant was collected and used to calculate the Miller units as follows: U = (OD420 * 1,000) / (OD600 * t(min) * v(ml)). Error was calculated as standard deviation from at least 3 biological replicates (S2 Data).

Kirby–Bauer disk diffusion susceptibility test

In a 14-cm diameter Petri dish, 50 mL of 1.5% agar media supplemented with indicated antibiotic and/or inducer were cast. After the polymerization, 12 mL of 0.375% agar media maintained at 50°C were mixed with 400 μL of overnight bacterial culture prior to be spread evenly on top of the solid media. Antibiotics discs were from Bio-Rad or made in the lab using sterile disks. The pictures were taken after the plates were incubated overnight at appropriate temperate. Images were captured on a Bio-Rad illuminator and with the Image Lab 4.1 software.

Efficiency of plating (EOP) assays

According to the OD600, the same quantity cells from an overnight culture were placed in the first line of a 96-well plates. All the wells were filled with 180 μL of appropriate media except the first line that was adjusted to 200 μL. The dilutions were done with 20 μL of the first line of wells transferred to the following line using a multipipette; this step was repeated until the 8 lanes were done. Finally, 5 μL of each dilution was dropped on a Petri dish with the solid media supplemented with the appropriate chemical(s).

Microscopy

Around 2 μL of an exponential phase culture of C. crescentus grown in PYE supplemented with indicated chemicals were spotted on a thin layer of 1% agarose pad for immobilization. Contrast microscopy pictures were taken using a phase 100× objective with an oil interface (Zeiss, alpha plan achromatic 100×/1.46 oil phase 3) using an Axio Imager M2 microscope (Zeiss), with appropriate filter (Visitron Sys-tems GmbH) and a cooled CCD camera (Photometrics, CoolSNAP HQ2) controlled via the Metamorph software (Molecular Devices).

Electrophoretic mobility shift assay (EMSA)

Cy5-labeled probes were generated by PCR with primers chemically modified with the fluorescent dye. The probes were purified by agarose gel electrophoresis. The reaction took place in a buffer containing: 40 mM Tris-HCl (pH 7.6), 60 mM KCl, and 0.1% glycerol, 240 μg of calf DNA, 800 μg of BSA, and 200 ng of labeled probe. After the addition of the indicated quantity of purified protein, the samples were placed in the dark for 30 minutes at room temperature. After a prerun of 30 minutes, the samples were loaded and migrated by electrophoresis in a nondenaturing 4% Tris-Bore-EDTA acrylamide (19:1) gel. After migration, the samples were observed in a Bio-Rad illuminator (Chemidoc MP) using the preinstalled parameters for Cy5 imaging; pictures were collected through Image Lab 4.1 software (Bio-Rad).

Chemical library screening

The Maybridge Chemical Library was delivered aliquoted into 96-well plates and dissolved in 50% DMSO. Petri dish of 22 cm containing PYE kanamycin 10 μg/mL were prepared using the Kirby–Bauer technique described above, with a NA1000 strain carrying the pLac290 plasmid harboring a fusion between the PacrA and the nptII kanamycin resistance gene. Potential inducers were detected when a growth area occur at the position of a drop after 48 hours of incubation at 30°C.

Immunoblot analysis/chase

The appropriate strain was grown for 2 to 4 hours at 30°C under constant agitation up to OD600nm of 0.4 to 0.6, then the inducer was added, and the cells were grown for 2 additional hours. For chase experiment, the strain was grown similarly up to OD600nm of 0.3 to 0.5, then chloramphenicol 5 μg/mL was supplemented with vigorous mixing, immediately followed by the addition of mentioned inducer. Protein samples from exponentially growing cells were separated on a SDS–polyacrylamide (37.5:1) gel electrophoresis and blotted on 0.45 μm pore PolyVinyliDenFluoride (PVDF) membranes (Immobilon-P from Sigma Aldrich). Membranes were blocked for 2 hours with 1× Tris-buffered saline (TBS) (50 mM Tris-HCl, 150 mM NaCl [pH 8]) that contain 0.1% Tween-20% and 8% powdered milk, followed by overnight incubation with the primary antibodies diluted in the same milk solution. The polyclonal antisera to AcrA (1:15,000 dilution), TipR (1:5,000 dilution), and CCNA_00164 (1:20,000 dilution) were used. The detection of primary antibodies was done using HRP-conjugated donkey anti-rabbit antibody (Jackson ImmunoResearch) with Western Blotting Detection System (Immobilon from Milipore), and an imaging was performed in Bio-Rad illuminator (Chemidoc MP, Bio-Rad).

Protein purification

TipR protein was expressed from pET21 in E. coli BL21(DE3)/pLysS. Cells were cultivated in LB at 37°C and induced by 1 mM of IPTG for 4 hours, up to an OD600nm of 0.5. The bacteria were harvested at 8,000 RPM for 30 minutes at 4°C. The pellet was resuspended in 25 mL of buffer (Tris-HCL 40 mM (pH 7.6), 50 mM KCl) before to be sonicated in a water–ice bath (15 cycles of 30 seconds ON, 30 seconds OFF). After centrifugation at 5,000g for 20 minutes at 4°C and filtration through 0.22 μm filters, the proteins contained in the supernatant were purified under native conditions successively using successively Qsepharose and Heparin columns (HiTrap systems from Cytia) according to manufacturer manual. Protein concentration was determined using the Bradford quantification method. Proteins were stored at −80°C in 40 mM Tris-HCl (pH 7.6), 50 mM KCl, and 10% glycerol.

Two-step resistance selection and Mut-seq

To isolate the Δbla;tipR::Tn derivatives that grow on CEF40 (Δbla;tipR::Tn acrAB-nodT*), Δbla;tipR::Tn cells were plated on CEF40, resistant mutant clones were pooled, and ФCr30-lysates were prepared from this pool. The tipR::Tn allele was then retransduced into Δbla cells, and transductants were selected on plates with KAN20 and then tested for growth on plates with CEF40. The acrAB-nodT locus was sequenced in several such CEF40 -resistant clones.

To isolate CEF10-resistant clones for Mut-Seq, the Δbla strain was spread on PYE cephalothin 10 μg/mL and incubated at 30°C for 48 hours. Around 10,000 colonies were harvested and pooled prior to DNA extraction using Ready-Lyse Lysozyme (Epicentre Lucigen) and DNAzol (Thermo Fisher). PCR were performed using the Q5 High-Fidelity DNA Polymerase according to manufacturer instructions with a maximum number of 15 cycles. The amplicons were purified with GeneJET Gel Extraction Kit (Thermo Fisher) and sent to sequencing at the iGE3 genomic platform in the CMU of the University of Geneva. The mixed amplicons were sent to the Genomic platform iGE3 at the University of Geneva. Library preparation and sequencing were done using a HiSeq 2500 with 50-bp paired-end reads. Data analysis was done using Burrows–Wheeler Alignment Tool version 0.7.5a and Samtools version 1.2 prior to be blast against the C. crescentus NA1000 reference genome (NC_011916.1). SNPs with an abundance lower than 5% were considered background (PCR amplification errors) and excluded from the analysis. The analysis is presented in S4 Data.

RNA extraction, deep- sequencing, and bioinformatics analysis

Overnight cultures in PYE of C. crescentus NA1000 (WT) were freshly restarted in 10 mL of PYE (starting OD660nm approximately 0.05) and incubated at 30°C under agitation to reach an OD600nm of 0.5. NAL-treated cultures were supplemented with 20 μg/mL of antibiotic 30 minutes before being harvested. First, to provide immediate stabilization of RNA, each cell cultures (4 ml) are treated with 2 volumes of RNA Protect Bacteria reagent (Qiagen, Switzerland) according to the manufacturer’s protocol. Then, cells were lysed in a Ready-Lyse lysozyme solution (Epicentre Technologies) according to manufacturer’s instructions, and lysates were homogenized through QiaShredder columns (Qiagen, Switzerland). The RNeasy Mini Kit (Qiagen, Switzerland) was used for total RNA extraction according to the manufacturer’s protocol, which included a first DNase treatment using the On-column DNase I digestion kit (Qiagen, Switzerland). Extracted total RNA was subjected to a second DNase treatment using Promega RQ1 at 1 unit/μg of RNA. Another total RNA cleanup was performed after the DNase treatment. The RNA concentration was measured using a Nanodrop 1000 spectrophotometer (Thermo Scientific, USA), and RNA quality was assessed using a 2100 Bioanalyzer Instrument (Agilent Technologies, USA). Two independent biological replicates were analyzed per condition.

RNA-Seq library preparation and sequencing were performed at Fasteris SA (Geneva, Switzerland). Bacterial rRNAs were removed from each total RNA samples, and RNA-Seq libraries were prepared using the Ovation Complete Prokaryotic RNA-Seq library system kit according to the manufacturer’s instructions. Single-end runs were performed on an Illumina NextSeq 500 instrument (50 cycles), yielding several million reads (stored as fastq files). Using the web-based analysis platform Galaxy (https://usegalaxy.org), the single-end sequenced reads quality was checked (FastQC, Galaxy Version 0.72), and reads were mapped (Bowtie2, Galaxy Version 2.3.4.3) to the C. crescentus NA1000 genome (NC_011916.1). The counts with the number of reads mapping to each gene feature were prepared using the htseq-count software, Galaxy Version 0.9.1. For each experimental series, the counts normalization and the statistical differential expression analysis was performed using the DESeq2 software, Galaxy Version 2.11.40.6. Sequence data (S1, S3, and S4 Data) have been deposited to the Gene Expression Omnibus (GEO) database (GSE225489 accession).

Chromatin immunoprecipitation coupled to deep sequencing (ChIP-Seq) and data analysis

Overnight cultures in PYE of C. crescentus NA1000(WT) or Δbla; PxylX::PxyldjlA-HA were freshly restarted in 80 mL of PYE (starting OD600nm of 0.05) and incubated at 30°C under agitation with 0.3% xylose when necessary. Cultures were treated with 20 μg/mL of NAL 30 minutes before fixation. Cultures of exponentially growing cells (OD600nm of 0.5) were supplemented with 10 μM sodium phosphate buffer (pH 7.6) and then treated with formaldehyde (1% final concentration) at room temperature for 10 minutes to achieve crosslinking. Subsequently, the cultures were incubated for an additional 30 minutes on ice and washed 3 times in phosphate-buffered saline (PBS, pH 7.4). The resulting cell pellets were stored at −80°C. After resuspension of the cells in TES buffer (10 mM Tris-HCl (pH 7.5), 1 mM EDTA, 100 mM NaCl) containing 10 mM of DTT, the cell resuspensions were incubated in the presence of Ready-Lyse lysozyme solution (Epicentre, Madison, WI) for 10 minutes at 37°C, according to the manufacturer’s instructions. Lysates were sonicated (Bioruptor Pico) at 4°C using 15 bursts of 30 seconds to shear DNA fragments to an average length of 0.2 to 0.5 kbp and cleared by centrifugation at 14,000 rpm for 2 minutes at 4°C. The volume of the lysates was then adjusted (relative to the protein concentration) to 1 mL using ChIP buffer (0.01% SDS, 1.1% Triton X-84 100, 1.2 mM EDTA, 16.7 mM Tris-HCl [pH 8.1], 167 mM NaCl) containing protease inhibitors (Roche) and precleared with 80 μl of Protein-A agarose (Roche, www.roche.com) and 100 μg BSA. Approximately 5% of each precleared lysate were reserved as total input samples (negative control samples). The precleared lysates were then incubated overnight at 4°C with purified polyclonal rabbit anti-TipR antibodies (1:400 dilution) [9] or monoclonal rabbit anti-HA antibodies (1:250 dilution) (Clone 114-2C-7, Merck Millipore). The immunocomplexes were captured after incubation with Protein-A agarose beads (presaturated with BSA) during a 4-hour incubation at 4°C and then washed subsequently with low salt washing buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-HCl (pH 8.1), 150 mM NaCl), with high salt washing buffer (0.1% SDS, 1% Triton X-100, 2 mM EDTA, 20 mM Tris-HCl (pH 8.1), 500 mM NaCl), with LiCl washing buffer (0.25 M LiCl, 1% NP-40, 1% deoxycholate, 1 mM EDTA, 10 mM Tris-HCl (pH 8.1)) and, finally, twice with TE buffer (10 mM Tris-HCl (pH 8.1), 1 mM EDTA). The immunocomplexes were eluted from the Protein-A agarose beads with 2 times 250 μL elution buffer (SDS 1%, 0.1 M NaHCO3, freshly prepared) and then, just like total input samples, incubated overnight with 300 mM NaCl at 65°C to reverse the crosslinks. The samples were then treated with 2 μg of Proteinase K for 2 hours at 45°C in 40 mM EDTA and 40 mM Tris-HCl (pH 6.5). DNA was extracted using phenol:chloroform:isoamyl alcohol (25:24:1), ethanol-precipitated using 20 μg of glycogen as a carrier and resuspended in 50 μL of DNAse/RNAse free water.

Immunoprecipitated chromatins were used to prepare sample libraries used for deep sequencing at Fasteris SA (Geneva, Switzerland). ChIP-Seq libraries were prepared using the DNA Sample Prep Kit (Illumina) following manufacturer instructions. Single-end run was performed on an Illumina Next-Generation DNA sequencing instruments (NextSeq High); 50 cycles were performed and yielded several million reads per sequenced samples. The single-end sequence reads stored in FastQ files were mapped against the genome of C. crescentus NA1000 (NC_011916.1) using Bowtie2 version 2.4.2.+galaxy0 available on the web-based analysis platform Galaxy (https://usegalaxy.org) to generate the standard genomic position format files (BAM). ChIP-Seq reads sequencing and alignment statistics are detailed in S3 Data. Then, BAM files were imported into SeqMonk version 1.47.2 (http://www.bioinformatics.babraham.ac.uk/projects/seqmonk/) to build ChIP-Seq-normalized sequence read profiles. Briefly, the genome was subdivided into 50 bp, and for every probe, we calculated the number of reads per probe as a function of the total number of reads (per million, using the Read Count Quantitation option). Using the web-based analysis platform Galaxy (https://usegalaxy.org), TipR ChIP-Seq peaks were called using MACS2 Version 2.1.1.20160309.6 (no broad regions option) relative to the total input DNA samples. The q-value (false discovery rate (FDR)) cutoff for called peaks was 0.05. Peaks were rank ordered according to their fold-enrichment values (S3 Data). Peaks with a fold-enrichment values >4 for TipR were retained for further analysis. Consensus sequences common to the 4 enriched TipR-associated loci were identified by scanning peak sequences (+ or − 75 bp relative to the coordinates of the peak summit) for conserved motifs using MEME (http://meme-suite.org/) [32]. Sequence data (S3 Data) have been deposited to the Gene Expression Omnibus (GEO) database (GSE225489 series).

Purification of anti-TipR antibodies

We eluted high-affinity polyclonal antibodies to TipR from an immunoblot. To this end, 200 μg of purified TipR protein was blotted onto a PVDF membrane, and the blot was incubated overnight in 5 ml of anti-TipR serum at 4°C. The membrane was then washed 4 times with 5% fat-free milk and once with PBS. A band corresponding to the size of TipR was then cut out from the membrane and placed in a 15-ml falcon tube. The band was then incubated with 1 mL of buffer 1 [50 mM Glycine (adjusted to pH 2.8 with HCl), 500 mM NaCl] for 3 minutes prior to neutralization with 8 μL of Tris-HCl (pH 8.2). The eluate was harvested, and the membrane was incubated again in 1 mL of buffer 2 [50 mM Glycine (adjusted to pH 2.2 with HCl), 500 mM NaCl] for 3 minutes prior to neutralization with 16 μL of Tris-HCl (pH 8.2). The second eluate was combined with the first and used as purified antibodies at a 1:500 dilution for immunoblotting and 1:400 dilution for ChIP-Seq experiments.

Plunge freezing of Caulobacter crescentus

C. crescentus cells were mixed with 10 nm Protein A conjugated colloidal gold particles (1:10 v/v, Cytodiagnostics), and 4 μl of the mixture was applied to a glow-discharged copper EM grid (R2/1 or R2/2, Quantifoil). The grid was automatically backside blotted for 4 to 6 seconds in a Mark IV Vitrobot (Thermo Fisher Scientific) by using a Teflon sheet on the front pad and plunge frozen in a liquid ethane–propane mixture (37%/63%) cooled by a liquid nitrogen bath. Frozen grids were stored in liquid nitrogen until loaded into the microscope.

Cryo-electron tomography

C. crescentus cells were imaged by cryoET [33]. Images were recorded on Titan Krios 300 kV microscopes (Thermo Fisher Scientific) equipped with a Quantum LS imaging filter operated at a 20-eV slit width and K2 or K3 Summit direct electron detectors (Gatan). Tilt series were collected using a bidirectional tilt-scheme from −60 to +60° in 2° increments. Total dose was 130 to 150 e/Å, and defocus was kept at −8 μm. Tilt series were acquired using SerialEM [34], drift corrected using alignframes, reconstructed, and segmented manually using the IMOD program suite [35]. To enhance contrast, tomograms were deconvolved with a Wiener-like filter [36].

In vivo protein aggregation assay

Exactly, not around 40 mL of exponentially growing cells (OD600 between 0.4 and 0.6) were rapidly cooled in ice bath and pelleted by centrifugation (6,000g, 10 minutes). All steps were performed at 4°C, and buffers were all supplemented with protease inhibitor cocktail from Roche (cOmplete tablets, EASYpack). The pellets were washed once in buffer A (50 mM Tris/HCl (pH 8.0), 150 mM NaCl) and resuspended in 300 μL of buffer A supplemented with 100 U/mL Ready-Lyse Lysozyme (Epicentre Lucigen), 10 μg/mL of DNAse I (Roche), and 10 μg/mL of RNAse A (Invitrogen). Cells were lysed in a Bioruptor (Diagenode) (set to high, 15 cycles for 30 seconds at 4°C). Lysates were centrifuged (5,000g, 10 minutes) 2 times to remove nonlysed cells. The protein concentration of the lysate was quantified by Bradford assay, and an aliquot was harvested as the Total fraction. The remaining samples were centrifuged (14,000g, 30 minutes) to pellet the insoluble protein fraction. The supernatant was kept as the Soluble fraction. The pellets were then resuspended in 300 μL buffer A supplemented with 1% (v/v) Triton X-100 with and incubated for 1 hour on ice with regular vortexing prior to the sonication in a Bioruptor (set to high, 1 cycle for 30 seconds) followed by a centrifugation (20,000g, 20 minutes) for washing. This procedure was repeated 3 times. The protein pellet was resuspended in 100 μL 1xSDS loading buffer prior to heating to 95°C for 10 minutes (S5 Data).

Immunoprecipitation

A volume of 50 mL of exponentially growing cells were harvested by centrifugation (20 minutes, 4,500 rpm at 4°C). The pellets were washed in 50 mL 1xPBS and centrifuged for 15 minutes with 6,000 rpm at 4°C and then washed once in 1 mL 1xPBS before centrifugation (for 5 minutes, 14,000 rpm, at 4°C). The pellets were resuspended in 1 mL TES (10 mM Tris-HCl (pH 7.5); 1 mM EDTA; 100 mM NaCl) containing protease inhibitor (cOmplete tablets, EASYpack from Roche) at 2 tabs per 50 mL and 100 U of Ready-lyse/ml of culture (Ready-Lyse Lysozyme from Epicentre Lucigen) prior to 10 minutes of incubation at room temperature. Each sample were supplemented with 50 μL NP-40 10% (AppliChem); 1 μL EDTA 0.5 M (pH 8.0) (Sigma); 10 μL MgCl2 1 M; 10 μL DNAse I (Roche) 10 mg/mL; 10 μL RNAse A (Invitrogen) 10 mg/mL and incubated 20 minutes at room temperature with constant agitation. The nonlysed cells were removed by centrifugation (10 minutes, 8,000 rpm at room temperature). The supernatant was harvested and centrifuged for 20 minutes with 14,000 rpm at 4°C. The remaining supernatant was kept as IP input.

For each sample, 25 μL of GFP-trap (GFP-Trap_A from Chromotek) or anti-HA affinity matrix beads (Roche) was rinsed 4 times with 1 mL TES buffer and centrifugation 2 minutes at 3,000 rpm. Then, 25 μL of beads were added to each IP input and left overnight at 4°C with agitation. The beads were harvested by centrifugation (2 minutes at 3,000 rpm) and washed 4 times with wash buffer (10 mM Tris-HCl 7.5; 150 mM NaCl; 0.5 mM EDTA; 0.5% n-dodecyl-β-D-maltoside (DDM)). The beads were then resuspended in 40 μL of 2× SDS sample buffer [125 mM Tris-HCl (pH 6.8); 4% SDS; 20% glycerol; 10% β-mercaptoethanol; 0.004% bromophenol blue] and incubated 10 minutes at 95°C. Each sample was centrifuged for 2 minutes at 3,000 rpm, and the supernatant was collected (without pipetting the beads). Samples of 5 or 10 μL of each collected fraction was analyzed by SDS-PAGE and Coomassie or immunoblotting.

Supporting information

S1 Fig. Resistance of tipR mutants to β-lactams.

(A) Growth of C. crescentus mutant strains on plates containing NAL10 (nalidixic acid, 10 μg/mL), CEF10 (cephalothin, 10 μg/mL), and/or PIR40 (piperacillin, 40 μg/mL) for 3 days. (B) Antibiograms of C. crescentus strains on PYE. Antibiotic discs, from top left to bottom right: Pivmecillinam 20 μg, Mecillinam 10 μg, Ampicillin 100 μg, Ceftazidime 40 μg, Meropenem 10 μg, Amoxicillin 4 μg, Cephalothin 30 μg, Cefepime 30 μg, Piperacillin 100 μg, Oxacillin 5 μg, Doripenem 10 μg, Cefotaxime 30 μg, Moxalactam 30 μg, Imipenem 10 μg, Aztreonam 30 μg, Cephalexin 40 μg, Cefsulodin 30 μg.

https://doi.org/10.1371/journal.pbio.3002040.s001

(PDF)

S2 Fig. RNA-Seq analysis of the tipR-acrAB-nodT locus.

(A) Representation of the reads [represented as reads per million (RPM)] obtained from the RNA-Seq experiment covering the tipR and acrAB-nodT region in the NA1000 strain (WT). Induction was performed on exponentially grown cells in PYE for 30 minutes with nalidixic acid (Nal, 20μg/mL). Red curves represent the reads in forward orientation (FW); blue curves represent the reads in reverse orientation (RV). The data from the analysis are deposited in S1 Data. (B) Scheme showing the location of the AcrB mutations conferring high level of cephalothin resistance (CEF40) to Δbla;tipR::Tn cells.

https://doi.org/10.1371/journal.pbio.3002040.s002

(PDF)

S3 Fig. Induction of AcrA and TipR by NAL.

(A) Immunoblot with polyclonal antibodies to AcrA and to TipR to probe extracts from NA1000 (WT) and tipR mutant (tipR::Tn) cells exponentially grown in PYE before and after 2 hours of induction of Nal (20 μg/mL). Loading control represents CCNA_00163 revealed with antibodies to CCNA_00163. (B) Immunoblot polyclonal antibodies to AcrA and to TipR to probe extracts of cells harvested during a time course experiment on NA1000 (WT) for 2 hours in the presence and absence of nalidixic acid (Nal, 10 μg/mL) in PYE.

https://doi.org/10.1371/journal.pbio.3002040.s003

(PDF)

S4 Fig. Antibiotic sensitivity of cells overexpressing AcrAB-NodT.

Kirby–Bauer-based disc diffusion assays to establish antibiograms of C. crescentus strains on PYE. Antibiotic discs, from top left to bottom right: Piramycin 100 μg, Vancomycin 30 μg, Flumequine 30 μg, Sparfloxacin 5 μg, Norfloxacin 5 μg, Kanamycin 20 μg, Neomycin 30 μg, Colistin 50 μg, Ciprofloxacin 5 μg, Teicoplanin 30 μg, Bacitracin 130 μg, Trimethoprim 5 μg, Azithromycin 15 μg, Rifampicin 30 μg, Clindamycin 2 μg, Erythromycin 15 μg, Fosfomycin 50 μg. Note: Kanamycin and neomycin resistance is conferred by the nptII gene located the transposon inserted in the tipR gene. Red circles demarcate the growth boundary for the reference strain.

https://doi.org/10.1371/journal.pbio.3002040.s004

(PDF)

S5 Fig. Envelope defects caused by long-term overexpression of AcrAB-NodT.

(A) Light microscopy (phase contrast) images of NA1000 (WT) and tipR::Tn cells during exponential growth in PYE in the presence or absence of Nalidixic acid (Nal, 10 μg/mL), 1-(1-Naphtylmethyl)-piperazine (NMP quantities are indicated. (B) Cryo-ET images C. crescentus Δbla cells (left), Δbla; tipR::Tn (center), and Δbla; pSRK-acrAB-nodT (right). Scale bar, 250 nm. CP, cytoplasm; IM, inner membrane; OM, outer membrane.

https://doi.org/10.1371/journal.pbio.3002040.s005

(PDF)

S6 Fig. Membrane deformation induced by AcrAB-NodT overexpression.

Representative cryo-tomographic slices (left) and 3D renderings (right) of C. crescentus Δbla;tipR::Tn (A) and Δbla;pSRK-acrAB-nodT (B) cells. Scale bar, 250 nm. CP, cytoplasm; IM (dark blue), inner membrane; OM (blue), outer membrane; S-Layer (cyan); vesicles (purple).

https://doi.org/10.1371/journal.pbio.3002040.s006

(PDF)

S7 Fig. PacrA activity in various tipR mutants.

β-galactosidase activity expressed from PacrA-lacZ in various mutants. Induction (+) was for 2 hours with nalidixic acid (Nal, 10 μg/mL). All levels are indicated as percentage of expression regarding the basal level of the WT (NA1000) without induction. The data from the analysis are deposited in S2 Data. All strains carry additionally the pPacrA-lacZ promoter probe plasmid.

https://doi.org/10.1371/journal.pbio.3002040.s007

(PDF)

S8 Fig. Abundance of TipR and AcrA in various tipR and PacrA point mutants.

Immunoblots probed with polyclonal antibodies to AcrA and to TipR in extracts of different PacrA (B, C, and D) or tipR (A) mutants. All inductions (+) were performed after 2 hours of treatment with 10 μg/mL of Nal on exponentially grown cells in PYE. Blots were also probed with antibodies to CCNA_00163 as loading control.

https://doi.org/10.1371/journal.pbio.3002040.s008

(PDF)

S9 Fig. Weblogo analysis of the TipR target DNA sequence.

An inverted repeat (IR) sequences was detected at each of the 3 TipR-binding positions on the chromosome, plus a half-site at the fourth target site. Consensus sequence of the IRs based on the sequences detected at position 399827, 927547, and 2365267 on the chromosome. This consensus has been identified using MEME software (mean P value: 4.48 × 10−6) and drawn by WebLogo (crooks).

https://doi.org/10.1371/journal.pbio.3002040.s009

(PDF)

S10 Fig. Occupancy of TipR on its chromosomal targets as determined by ChIP-Seq.

Representation of the reads [in reads per million (RPM)] obtained from the ChIP-Seq analyses covering the TipR binding regions. Positions are indicated under the graphic. Induction was performed for 30 minutes with Nal (20 μg/mL) in PYE (blue line) compared with the not induced condition (black line). The yellow boxes indicate the positions of the putative TipR binding site. The data from the analysis are deposited in S3 Data.

https://doi.org/10.1371/journal.pbio.3002040.s010

(PDF)

S11 Fig. Screen for chemical inducers of PacrA using the PacrA-nptII reporter.

Reporter assay of PacrA activity fused with the nptII gene (PacrA-nptII conferring kanamycin resistance) cloned on plasmid plac290 (pPacrA-nptII). To identify inducers, chemicals were spotted on WT cells carrying the pPacrA-nptII reporter plasmid embedded on soft agar on PYE plates both containing with kanamycin (10 μg/mL). Plates were incubated for 2 days at 30°C.

https://doi.org/10.1371/journal.pbio.3002040.s011

(PDF)

S12 Fig. Quantification of PacrA-lacZ activity by different chemical inducers.

(A) β-galactosidase activity using the PacrA-lacZ in the ΔacrAB-nodT cells (A) or Pbla-lacZ in WT cells. (B) Inductions were performed for 3 hours. All levels are indicated in percentage of expression regarding the basal level of the uninduced state. The data from the analysis are deposited in S2 Data.

https://doi.org/10.1371/journal.pbio.3002040.s012

(PDF)

S13 Fig. AcrA abundance upon the addition of chemical inducers of PacrA.

(A) Immunoblots probed with polyclonal antibodies to AcrA and to TipR to detect AcrA and TipR in tipR::Tn cells complemented with the pSRK-tipR plasmid, induced with the indicated amount of IPTG (in mM) and Nalidixic acid (Nal, 10 μg/mL) for 2 hours. (B) Immunoblot using polyclonal antibodies to AcrA and to TipR to probe extracts of multiple mutant strains grown in PYE with IPTG 0.5 mM, with and without 2 hours induction of Nal (10 μg/mL), rhodamine 6G (Rh6, 2 μg/mL), and malachite green (MG, 2 μg/mL).

https://doi.org/10.1371/journal.pbio.3002040.s013

(PDF)

S14 Fig. Crystal violet does not affect binding of IHF to its target.

EMSA with 4 μM of IHF protein and 200 ng of Cy5-labelled attR DNA as probe. All quantities of crystal violet (CV) indicated are in μg/mL.

https://doi.org/10.1371/journal.pbio.3002040.s014

(PDF)

S15 Fig. Effect of chemical inducers and gyrA variants on PacrA-lacZ.

β-galactosidase activity measurements from PacrA-lacZ in NA1000 (WT) carrying the pMT335 or a derivative to express a gyrAF96N (GyrA*) or gyrA from Brucella melitensis (Bm) grown in PYE Van 50 μM. Antibiotics were used at the following concentrations: nalidixic acid (Nal, 10 μg/mL), ciprofloxacin (Cip, 2 μg/mL), rhodamine 6G (Rho, 2 μg/mL), novobiocin (Novo, 10 μg/mL). All levels are indicated in percentage of expression regarding the basal level of WT before induction. The data from the analysis can be found in S2 Data.

https://doi.org/10.1371/journal.pbio.3002040.s015

(PDF)

S16 Fig. Abundance of AcrA and TipR in protease mutants.

Immunoblot analysis using polyclonal antibodies to AcrA and to TipR to determine the steady-state levels of AcrA and TipR in various protease mutants of C. crescentus, before and after induction with nalidixic acid (Nal, +, 10 μg/mL) for 2 hours in PYE. Loading control is anti-CCNA_00164. All samples were loaded on the same immunoblot.

https://doi.org/10.1371/journal.pbio.3002040.s016

(PDF)

S17 Fig. Pull-down of ClpX-YFP.

Coomassie Blue–stained PAGE (12% gel) (left) and immunoblot (right) of a ClpX-YFP co-immunoprecipitation (GFP Trap Matrix) with polyclonal antibodies to TipR and to AcrA. Induction with (+) or without (−) Nal (10 μg/mL).

https://doi.org/10.1371/journal.pbio.3002040.s017

(PDF)

S18 Fig. Induction of AcrA and TipR in cells lacking DjlA.

Immunoblot analysis of extracts from WT and co-chaperone mutants using polyclonal antibodies to TipR and to AcrA. All inductions (+) were performed for 2 hours with 10 μg/mL of nalidixic acid (Nal). Strains carrying the pMT335 are induced with vanillate 100 μM (Van). Immunoblots performed with antibodies to CCNA_00163 serve as loading controls.

https://doi.org/10.1371/journal.pbio.3002040.s018

(PDF)

S19 Fig. Antibiogram of cells lacking or overexpressing DjlA.

Antibiograms of C. crescentus strains using antibiotic discs, from top left to bottom right, Mecillinam 10 μg, Ceftazidime 40 μg, Meropenem 10 μg, Amoxicillin 20 μg, Cephalothin 30 μg, Cefepime 30 μg, Piperacillin 100 μg, Oxacillin 5 μg, Doripenem 10 μg, Cefotaxime 30 μg, Moxalactam 30 μg, Imipenem 10 μg, Aztreonam 30 μg, Cephalexin 40 μg, Cefsulodin 30 μg. Nal induction performed at 10 μg/mL. All plates with strains carrying a pMT335 or pMT335-djlA are supplemented with vanillate 100 μM.

https://doi.org/10.1371/journal.pbio.3002040.s019

(PDF)

S20 Fig. EOP assays with DjlA expressed from plasmids in WT and mutant cells.

Efficiency of plating (EOP) assay determined by 10-fold serial dilutions of ΔdjlA mutant on plates containing ethidium bromide (EtBr, 6 μg/mL) to probe for efflux pump activity. All strains contain pMT335 plasmid or a derivative (pMT335-djlA, pMT335-djlAH187A, and pMT335-djlAH187Q) expressing DjlA, grown on vanillate (Van) at 100 μM and gentamicin (Gent) at 1 μg/mL.

https://doi.org/10.1371/journal.pbio.3002040.s020

(PDF)

S21 Fig. Pull-down of DjlA-HA.

(A) Coomassie staining after 12% PAGE (left) and immunoblotting (right) of a DjlA-HA co-immunoprecipitation eluate (from anti HA affinity matrix) probed with polyclonal antibodies to AcrA and to TipR. Induction was with Nal (+) at 10 μg/mL for 2 hours. (B) Immunoblots probed with polyclonal antibodies to TipR and to AcrA after pull-down of HA-tagged protein from extracts of cells expressing DjlA-HA, DjlAH187A-HA, DjlAH187Q-HA, DnaJ1-HA, and DnaJ2-HA using the anti HA affinity matrix. Induction was with Nal (+) at 10 μg/mL for 2 hours.

https://doi.org/10.1371/journal.pbio.3002040.s021

(PDF)

S22 Fig. ChIP-Seq analysis of the divergent promoter for acrAB-nodT and tipR.

(A) Representation of the reads (in reads per million (RPM)) obtained from the ChIP-Seq analyses covering the TipR binding regions. Positions are indicated under the graphic. Induction was performed for 30 minutes with Nalidixic acid (Nal) 20 μg/mL in PYE (blue line) compared with the not induced condition (black line). The data from the analysis are deposited in S23 Fig. (B) Immunoblotting of extracts from cells expressing DjlA-HA, DnaJ1 to 5-HA and ClpX-HA used in the ChIP-Seq experiment. The blots were probed with polyclonal antibodies to TipR and monoclonal antibodies to the HA tag.

https://doi.org/10.1371/journal.pbio.3002040.s022

(PDF)

S23 Fig. Quantification of AcrA aggregation in various mutants.

Graphical representation of 4 replicates of aggregation of assay quantified with ImageJ from immunoblotting of AcrA in soluble versus insoluble cell lysates as determined by immunoblotting using polyclonal antibodies to AcrA. All inductions (+) were performed for 2 hours with 10 μg/mL of Nal. Strains carrying the pMT335 are induced with vanillate (Van, 100 μM). All samples are normalized regarding the quantification of uninduced WT set at 100%. The data from the analysis are deposited in S5 Data.

https://doi.org/10.1371/journal.pbio.3002040.s023

(PDF)

S24 Fig. LC-MS/MS analysis of DjlA-HA pull-downs.

Graphical representation of the total spectrum count of various proteins detected by LC-MS/MS upon immunoprecipitation of DjlA-HA from soluble cell lysates using anti-HA affinity matrix. The lysates used were from WT expressing DjlA-HA plus (blue) and minus (grey) NAL, and a WT control without DjlA-HA (red). All the TonB-dependant receptors (TBDRs) are annotated with their respective CCNA_#. The data from the LC-MS/MS analysis are deposited in S5 Data.

https://doi.org/10.1371/journal.pbio.3002040.s024

(PDF)

S1 Data. RNA-Seq analysis after induction with NAL.

https://doi.org/10.1371/journal.pbio.3002040.s025

(XLSX)

S2 Data. Quantification of promoter probe assays.

https://doi.org/10.1371/journal.pbio.3002040.s026

(XLSX)

S3 Data. ChIP-Seq analysis after induction with NAL.

https://doi.org/10.1371/journal.pbio.3002040.s027

(XLSX)

S4 Data. Mut-Seq analysis of acrAB-nodT selected by cephalothin resistance of Δbla cells.

https://doi.org/10.1371/journal.pbio.3002040.s028

(XLSX)

S5 Data. Quantification of AcrA aggregation assay.

https://doi.org/10.1371/journal.pbio.3002040.s029

(XLSX)

S6 Data. List of strains, plasmids, primers, and synthetic genes.

https://doi.org/10.1371/journal.pbio.3002040.s030

(DOCX)

Acknowledgments

We acknowledge instrument access at the imaging platform ScopeM at ETH Zürich. JC wishes to acknowledge Sabine Quindou, Jamy Gourmaud, and Frédéric Courant for nurturing an entire generation’s curiosity.

References

  1. 1. Karam G, Chastre J, Wilcox MH, Vincent JL. Antibiotic strategies in the era of multidrug resistance. Crit Care. 2016;20(1):136. Epub 2016/06/23. pmid:27329228; PubMed Central PMCID: PMC4916531.
  2. 2. Collaborators GBDAR. Global mortality associated with 33 bacterial pathogens in 2019: a systematic analysis for the Global Burden of Disease Study 2019. Lancet. 2022;400(10369):2221–2248. Epub 2022/11/25. pmid:36423648.
  3. 3. Darby EM, Trampari E, Siasat P, Gaya MS, Alav I, Webber MA, et al. Molecular mechanisms of antibiotic resistance revisited. Nat Rev Microbiol. 2023;21(5):280–295. Epub 2022/11/22. pmid:36411397.
  4. 4. Ruiz N, Kahne D, Silhavy TJ. Advances in understanding bacterial outer-membrane biogenesis. Nat Rev Microbiol 2006;4 (1):57–66. PubMed pmid:16357861.
  5. 5. Nikaido H, Pages JM. Broad-specificity efflux pumps and their role in multidrug resistance of Gram-negative bacteria. FEMS Microbiol Rev. 2012;36(2):340–363. Epub 2011/06/29. pmid:21707670; PubMed Central PMCID: PMC3546547.
  6. 6. Lamut A, Peterlin Mašič L, Kikelj D, Tomašič T. Efflux pump inhibitors of clinically relevant multidrug resistant bacteria. Med Res Rev. 2019;39(6):2460–2504.
  7. 7. Ahmed M, Borsch CM, Taylor SS, Vázquez-Laslop N, Neyfakh AA. A protein that activates expression of a multidrug efflux transporter upon binding the transporter substrates. J Biol Chem. 1994;269(45):28506–28513. Epub 1994/11/11. pmid:7961792.
  8. 8. Li XZ, Plesiat P, Nikaido H. The challenge of efflux-mediated antibiotic resistance in Gram-negative bacteria. Clin Microbiol Rev. 2015;28(2):337–418. Epub 2015/03/20. pmid:25788514; PubMed Central PMCID: PMC4402952.
  9. 9. Kirkpatrick CL, Viollier PH. Synthetic interaction between the TipN polarity factor and an AcrAB-family efflux pump implicates cell polarity in bacterial drug resistance. Chem Biol. 2014;21(5):657–665. Epub 2014/04/15. pmid:24726830.
  10. 10. Delaby M, Varesio LM, Degeorges L, Crosson S, Viollier PH. The DUF1013 protein TrcR tracks with RNA polymerase to control the bacterial cell cycle and protect against antibiotics. Proc Natl Acad Sci U S A. 2021;118(8). Epub 2021/02/20. pmid:33602809.
  11. 11. Blair JM, Piddock LJ. Structure, function and inhibition of RND efflux pumps in Gram-negative bacteria: an update. Curr Opin Microbiol. 2009;12(5):512–519. Epub 2009/08/12. pmid:19664953.
  12. 12. Deng W, Li C, Xie J. The underling mechanism of bacterial TetR/AcrR family transcriptional repressors. Cell Signal. 2013;25(7):1608–1613. Epub 2013/04/23. pmid:23602932.
  13. 13. Kirkpatrick CL, Martins D, Redder P, Frandi A, Mignolet J, Chapalay JB, et al. Growth control switch by a DNA-damage-inducible toxin-antitoxin system in Caulobacter crescentus. Nat Microbiol. 2016;1:16008. Epub 2016/08/31. pmid:27572440.
  14. 14. Ramos JL, Martinez-Bueno M, Molina-Henares AJ, Teran W, Watanabe K, Zhang X, et al. The TetR family of transcriptional repressors. Microbiol Mol Biol Rev. 2005;69(2):326–356. pmid:15944459.
  15. 15. Orth P, Cordes F, Schnappinger D, Hillen W, Saenger W, Hinrichs W. Conformational changes of the Tet repressor induced by tetracycline trapping. J Mol Biol. 1998;279(2):439–447. Epub 1998/06/27. pmid:9642048.
  16. 16. Tiebel B, Garke K, Hillen W. Observing conformational and activity changes of Tet repressor in vivo. Nat Struct Biol. 2000;7(6):479–481.
  17. 17. Cuthbertson L, Nodwell JR. The TetR family of regulators. Microbiol Mol Biol Rev. 2013;77(3):440–475. pmid:24006471.
  18. 18. Colclough AL, Scadden J, Blair JMA. TetR-family transcription factors in Gram-negative bacteria: conservation, variation and implications for efflux-mediated antimicrobial resistance. BMC Genomics. 2019;20(1):731. pmid:31606035
  19. 19. Castanie-Cornet MP, Bruel N, Genevaux P. Chaperone networking facilitates protein targeting to the bacterial cytoplasmic membrane. Biochim Biophys Acta. 2014;1843(8):1442–1456. Epub 2013/11/26. pmid:24269840.
  20. 20. Mayer MP, Kityk R. Insights into the molecular mechanism of allostery in Hsp70s. Front Mol Biosci. 2015;2:58. pmid:26539440.
  21. 21. Muheim C, Götzke H, Eriksson AU, Lindberg S, Lauritsen I, Nørholm MHH, et al. Increasing the permeability of Escherichia coli using MAC13243. Sci Rep. 2017;7(1):17629. pmid:29247166
  22. 22. Bhat NH, Vass RH, Stoddard PR, Shin DK, Chien P. Identification of ClpP substrates in Caulobacter crescentus reveals a role for regulated proteolysis in bacterial development. Mol Microbiol. 2013;88(6):1083–1092. pmid:23647068; PubMed Central PMCID: PMC3681837
  23. 23. Genevaux P, Schwager F, Georgopoulos C, Kelley WL. The djlA gene acts synergistically with dnaJ in promoting Escherichia coli growth. J Bacteriol. 2001;183(19):5747–5750. Epub 2001/09/07. pmid:11544239; PubMed Central PMCID: PMC95468.
  24. 24. Genevaux P, Wawrzynow A, Zylicz M, Georgopoulos C, Kelley WL. DjlA is a third DnaK co-chaperone of Escherichia coli, and DjlA-mediated induction of colanic acid capsule requires DjlA-DnaK interaction. J Biol Chem. 2001;276(11):7906–7912. Epub 2000/12/07. pmid:11106641.
  25. 25. Castanie-Cornet MP, Cam K, Jacq A. RcsF is an outer membrane lipoprotein involved in the RcsCDB phosphorelay signaling pathway in Escherichia coli. J Bacteriol. 2006;188(12):4264–4270. Epub 2006/06/03. pmid:16740933; PubMed Central PMCID: PMC1482940.
  26. 26. Vincent CD, Buscher BA, Friedman JR, Williams LA, Bardill P, Vogel JP. Identification of non-dot/icm suppressors of the Legionella pneumophila DeltadotL lethality phenotype. J Bacteriol. 2006;188(23):8231–8243. Epub 2006/09/26. pmid:16997951; PubMed Central PMCID: PMC1698199.
  27. 27. Nierman WC, Feldblyum TV, Laub MT, Paulsen IT, Nelson KE, Eisen J, et al. Complete genome sequence of Caulobacter crescentus. Proc Natl Acad Sci U S A. 2001;98(7):4136–4141.
  28. 28. Tewari KM, Eggleston IM. Chemical approaches for the enhancement of 5-aminolevulinic acid-based photodynamic therapy and photodiagnosis. Photochem Photobiol Sci. 2018;17(11):1553–1572. Epub 2018/10/18. pmid:30328459.
  29. 29. Alav I, Bavro VN, Blair JMA. A role for the periplasmic adaptor protein AcrA in vetting substrate access to the RND efflux transporter AcrB. Sci Rep. 2022;12(1):4752. Epub 2022/03/21. pmid:35306531; PubMed Central PMCID: PMC8934357.
  30. 30. Ely B. Genetics of Caulobacter crescentus. Methods Enzymol. 1991;204:372–384.
  31. 31. Viollier PH, Thanbichler M, McGrath PT, West L, Meewan M, McAdams HH, et al. Rapid and sequential movement of individual chromosomal loci to specific subcellular locations during bacterial DNA replication. Proc Natl Acad Sci U S A. 2004;101(25):9257–9262. Epub 2004/06/05. [pii]. pmid:15178755; PubMed Central PMCID: PMC438963.
  32. 32. Bailey TL, Boden M, Buske FA, Frith M, Grant CE, Clementi L, et al. MEME SUITE: tools for motif discovery and searching. Nucleic Acids Res. 2009;37(Web Server issue):W202-W208. Epub 2009/05/22. pmid:19458158; PubMed Central PMCID: PMC2703892.
  33. 33. Weiss GL, Medeiros JM, Pilhofer M. In Situ Imaging of Bacterial Secretion Systems by Electron Cryotomography. Methods Mol Biol. 2017;1615:353–375. Epub 2017/07/02. pmid:28667625.
  34. 34. Mastronarde DN. Automated electron microscope tomography using robust prediction of specimen movements. J Struct Biol. 2005;152(1):36–51. Epub 2005/09/27. [pii] pmid:16182563.
  35. 35. Kremer JR, Mastronarde DN, McIntosh JR. Computer visualization of three-dimensional image data using IMOD. J Struct Biol. 1996;116 (1):71–76. pmid:8742726.
  36. 36. Tegunov D, Cramer P. Real-time cryo-electron microscopy data preprocessing with Warp. Nat Methods. 2019;16(11):1146–1152. pmid:31591575