crossmark

Proper Control of Caulobacter crescentus Cell Surface Adhesion Requires the General Protein Chaperone DnaK Daniel S. Eaton,a Sean Crosson,a,b Aretha Fiebiga Department of Biochemistry and Molecular Biology, University of Chicago, Chicago, Illinois, USAa; Department of Microbiology, University of Chicago, Chicago, Illinois, USAb

ABSTRACT

Growth in a surface-attached bacterial community, or biofilm, confers a number of advantages. However, as a biofilm matures, high-density growth imposes stresses on individual cells, and it can become less advantageous for progeny to remain in the community. Thus, bacteria employ a variety of mechanisms to control attachment to and dispersal from surfaces in response to the state of the environment. The freshwater oligotroph Caulobacter crescentus can elaborate a polysaccharide-rich polar organelle, known as the holdfast, which enables permanent surface attachment. Holdfast development is strongly inhibited by the small protein HfiA; mechanisms that control HfiA levels in the cell are not well understood. We have discovered a connection between the essential general protein chaperone, DnaK, and control of C. crescentus holdfast development. C. crescentus mutants partially or completely lacking the C-terminal substrate binding “lid” domain of DnaK exhibit enhanced bulk surface attachment. Partial or complete truncation of the DnaK lid domain increases the probability that any single cell will develop a holdfast by 3to 10-fold. These results are consistent with the observation that steady-state levels of an HfiA fusion protein are significantly diminished in strains that lack the entire lid domain of DnaK. While dispensable for growth, the lid domain of C. crescentus DnaK is required for proper chaperone function, as evidenced by observed dysregulation of HfiA and holdfast development in strains expressing lidless DnaK mutants. We conclude that DnaK is an important molecular determinant of HfiA stability and surface adhesion control. IMPORTANCE

Regulatory control of cell adhesion ensures that bacterial cells can transition between free-living and surface-attached states. We define a role for the essential protein chaperone, DnaK, in the control of Caulobacter crescentus cell adhesion. C. crescentus surface adhesion is mediated by an envelope-attached organelle known as the holdfast. Holdfast development is tightly controlled by HfiA, a small protein inhibitor that directly interacts with a WecG/TagA-family glycosyltransferase required for holdfast biosynthesis. We demonstrate that the C-terminal lid domain of DnaK is not essential for growth but is necessary for proper control of HfiA levels in the cell and for control of holdfast adhesin development.

C

ommunities of surface-attached bacteria, called biofilms, account for the majority of cells on the surface of our planet. Biofilms allow bacteria to engage in metabolic cooperation, share genetic information, and afford protection from chemical and physical stresses in the environment. Moreover, organic biopolymers and ions accumulate at surfaces through a process known as conditioning (1–4); thus, the ability of a bacterial cell to adhere can confer a nutritional advantage, particularly in oligotrophic environments where nutrients are extremely scarce (5). Clearly, the capacity to transition from a free-living, unicellular lifestyle into a surface-attached multicellular community is an important adaptive feature of bacterial physiology that increases the breadth of niches a species can inhabit. The Gram-negative bacterium Caulobacter crescentus thrives in dilute, freshwater ecosystems and can permanently adhere to a number of chemically diverse surfaces via a polar adhesin called the holdfast (6–10). The holdfast organelle is a polysacchariderich matrix which is synthesized via a Wzx/Wzy-dependent secretion system and attached to the cell via localized outer membrane proteins (11–16). Given that holdfast-mediated surface attachment is permanent, it is not surprising that the decision to elaborate a holdfast is a highly regulated process that is controlled by cell cycle, environmental, and physical signals. Newborn swarmer cells are motile and lack holdfasts, consistent with a role in disper-

October 2016 Volume 198 Number 19

sal. During development, or when a swarmer cell encounters a physical surface, holdfasts are elaborated as C. crescentus trades motility for replication and differentiates into a stalked cell (17). A key regulator of holdfast development in Caulobacter is a small, hydrophobic protein, HfiA, which responds to both cell cycle and environmental cues and functions as a potent inhibitor of holdfast synthesis (18). To understand how HfiA functions to inhibit holdfast development, we previously performed a genetic selection for mutants that were insensitive to HfiA, i.e., mutants in which surface adhesion was unaffected by hfiA overexpression (Fig. 1B). This strategy led to the identification of HfsJ, a putative WecG/TagA family glyco-

Received 10 January 2016 Accepted 30 March 2016 Accepted manuscript posted online 4 April 2016 Citation Eaton DS, Crosson S, Fiebig A. 2016. Proper control of Caulobacter crescentus cell surface adhesion requires the general protein chaperone DnaK. J Bacteriol 198:2631–2642. doi:10.1128/JB.00027-16. Editor: G. A. O’Toole, Geisel School of Medicine at Dartmouth Address correspondence to Aretha Fiebig, [email protected]. Supplemental material for this article may be found at http://dx.doi.org/10.1128 /JB.00027-16. Copyright © 2016, American Society for Microbiology. All Rights Reserved.

Journal of Bacteriology

jb.asm.org

2631

Eaton et al.

A

B

HfiA++ (no adhesion)

HfiA++ (mutants adhere)

DnaK HfiA

HfsJ

Holdfast development

wash

adherent mutants repopulate

C

N-

ATPase

Substrate binding

-C

Lid

Q497* K568* I578::Tn5

D Substrate binding pocket

C

N

568 497 587

ATPase domain

Substrate binding domain

FIG 1 Role for DnaK in the control of C. crescentus cell surface adhesion and holdfast development. (A) Model for regulation of holdfast development by DnaK. HfsJ, a putative glycosyltransferase required for synthesis of the adhesive holdfast, functions downstream of HfiA, a potent inhibitor of holdfast development. Data presented in this study support a model in which the general protein chaperone, DnaK, stabilizes HfiA and thereby functions to modulate cell adhesion. (B) Selection strategy to identify adherent mutants that develop holdfast when hfiA is overexpressed (see the text for details). Red dots indicate cells that have elaborated an adhesive holdfast. (C and D) One-dimensional model of DnaK (C) and structural model of DnaK in the ADP-bound, lid-closed conformation (PDB entry 2KH0) (D). The N-terminal ATPase domain is shown in gray. The C-terminal substrate binding domain is rainbow colored from N (blue)- to C (red)-terminal residues. The sites of the DnaK mutations presented in this study are marked with black arrowheads on both models. An asterisk indicates a nonsense mutation.

syltransferase that is necessary for holdfast synthesis. Genetic and biochemical data support a model in which HfsJ is directly inhibited by HfiA (18) (Fig. 1A). From this same genetic selection we also mapped two independent mutant alleles of dnaK, a highly conserved gene encoding an ATP-dependent general protein chaperone that is essential for C. crescentus viability (19, 20). To date, these adhesive dnaK mutant strains have remained uncharacterized. DnaK interacts with a large number of protein substrates (21) to facilitate de novo protein folding, membrane protein targeting, and refolding of denatured proteins (22–25). Substrate recognition is largely based on physiochemical properties. DnaK recognizes short exposed hydrophobic regions, consistent with a role in refolding denatured proteins (26–28). DnaK has an N-terminal ATPase domain followed by a C-terminal substrate-binding domain (Fig. 1C and D). The substrate-binding domain is further divided into a ␤-sheet domain, which forms the substrate binding pocket, and an alpha-helical lid domain at the extreme C terminus (27). The conformation of the substrate binding domain and its affinity for substrate depend on the state of the ATPase domain. In the ATP-bound state, the lid is open and affinity for substrate is low. Upon ATP hydrolysis to ADP the lid closes on the substratebinding domain as shown in Fig. 1D, and the affinity for substrate is dramatically enhanced (29–31). ATP-dependent regulation of substrate affinity is independent of the C-terminal helical lid, although this domain enhances the residence time of substrates on DnaK, which reduces protein aggregation and facilitates substrate folding (32, 33).

2632

jb.asm.org

Both dnaK mutations identified in our genetic selection resulted in truncation of the alpha-helical lid at the extreme C terminus. Given that HfiA is a largely hydrophobic protein, this result suggested that DnaK facilitates stabilization of HfiA and, thus, its ability to function as an inhibitor of holdfast synthesis. Indeed, we demonstrate that truncation of the DnaK lid domain results in increased surface attachment and holdfast development in both hfiA overexpression and wild-type backgrounds. Further, we show that these enhanced adhesion phenotypes are accentuated by deletion of the entire helical lid domain. Consistent with the elevated adhesion phenotypes, the steady-state level of an HfiA fusion protein is reduced in these dnaK mutant strains. Our results support a model in which DnaK, as a general molecular chaperone, plays an important role in stabilizing HfiA in the cell and thereby affects the capacity of C. crescentus to adhere to surfaces in a regulated manner. Our results provide evidence that the nonessential lid domain of DnaK plays a crucial role in ensuring HfiA function as an environmental and cell cycle-controlled regulator of holdfast development and surface attachment. MATERIALS AND METHODS Growth conditions. C. crescentus cells were cultured in either PYE (0.2% peptone, 0.1% yeast extract, 1 mM MgSO4, 0.5 mM CaCl2) or M2 salts (49), supplemented with 0.15% xylose as the carbon source, at 30°C. To induce expression from the Pxyl promoter in cells grown in PYE, the culture was supplemented with 0.1% xylose. Antibiotics were used at the following concentrations in liquid and solid medium, respectively: kana-

Journal of Bacteriology

October 2016 Volume 198 Number 19

DnaK Affects Caulobacter Surface Attachment

TABLE 1 Plasmids used in this study Name

Description

Comments and/or primers used to amplify inserted sequencea

Reference or source

pMT805

pBXMCS-6

51

pMT685

pXCHYC-6

pMT680

pXGFPC-6

Replicating plasmid (mid-copy number, BBR origin) with xylose-inducible promoter; Cmr Integrative plasmid (ColEI origin), targets integration to xylX, carries xylose-inducible promoter, can generate C-terminal mCherry fusions; Cmr Integrative plasmid, targets integration to xylX, can generate C-terminal GFP fusions; Cmr Allele replacement plasmid; Kanr SacB Derivative of plasmid pNPTS138 in which the Kanr gene was replaced with a Cmr gene

pNPTS138 pNPTS138-CAT pAF463 pAF418

pMT805-Pxyl-hfiA pMT685-PhfiA-hfiA-mcherry

pAF508 pAF509 pAF506

pMT680-PhfiA-mcherry-hfiA pMT680-PhfiA-mcherry-hfiA pNTPS-dnaK(K568*, I579V)

FC2328

pNPTS-dnaK(Q497*)

pAF598

pNPTS-dnaK (C-terminal allele restoration) pNPTS-CAT-dnaK (C-terminal allele restoration) pNPTS-⌬hfiA pNPTS-⌬lon

pAF594 pAF401 FC1678

F, catatgACGGACGTGATGCACTACAG; R, gagctcGCGCGCCAAGGTCTaAGC See Materials and Methods See Materials and Methods F, GCCGACGAaTTCAAGAAGGAGC; R, aaaaggatccACGAAA GTGGCGTTTACAA; amplified from B5A suppressor strain UP F, tttaggatCCTGAAGAAGAGCGACATC; UP R⫹, tgactcg cctaGATGCGGATCGAGTGCT; DN F⫹, catctagGCGAGT CATGCGCGCG; DN R, tttcaagcttCTCCAGGTCATAACGCAGGT F, tttaggatCCTGAAGAAGAGCGACATC; R, tttcaagcttCTCCA GGTCATAACGCAGGT; amplified from wild type Same as above

UP F, atatggatccAGGTGGCCAAGAAGGCTATC; UP R⫹, gtgcgtcagcatggcATCCCGCAGCGGCAAGAC; DN F⫹, ttgc cgctgcgggatGCCATGCTGACGCACTAA; DN R, tatagctagcCGCTCTACCGCTGAGCTAAC

51

51 M. R. K. Alley This work 18 This work This work This work This work

This work

This work This work 18 This work

a Introduced or mutated stop codons are in boldface, 5= tails and other deviations from target sequences are in lowercase, and restriction sites are in italics. Symbols: ⫹, overlap extension primer (overlapping regions are underlined); *, nonsense mutation. Cmr indicates chloramphenicol resistance (encoded by cat); Kanr indicates kanamycin resistance (encoded by npt1).

mycin, 5 and 25 ␮g/ml; chloramphenicol, 1 and 1.5 ␮g/ml; nalidixic acid, 20 ␮g/ml. Escherichia coli cells were cultured in LB at 37°C. The following antibiotics were used in both liquid and solid medium: kanamycin, 50 ␮g/ml; chloramphenicol, 20 ␮g/ml. Plasmid and strain constructions. C. crescentus DNA was amplified with KOD Xtreme hot-start polymerase (EMD Millipore) and supplementing reactions with 5% dimethyl sulfoxide (DMSO). Restriction sites for cloning were added to the ends of the primers. Amplified products were digested with appropriate restriction enzymes (New England BioLabs) and ligated into similarly digested, phosphatase-treated, and gelpurified plasmids using T4 DNA ligase (New England BioLabs). Plasmid ligations were transformed into E. coli Top10 (Life Technologies, Invitrogen). All cloned products were sequence confirmed. Plasmids used and generated in this work are listed in Table 1. Primer sequences used to amplify cloned sequences are also listed in Table 1. Point mutant and null alleles were generated using an overlap extension PCR strategy (50) and unique restriction enzyme cut sites in the outermost primer sequences. Overlapping primers are indicated with a plus at the end of the name. Amplified alleles were ligated into the corresponding restriction sites in pNPTS138 or pNPTS138-CAT. Overlap extension PCR was used similarly to generate fusions in pAF508 and pAF509. First, HfiA starting at genome position 938,855 (the original predicted start) or position 938,825 (a later start codon) was amplified and cloned into the SacI and NheI sites of pXCHYN-6 to generate xylose-inducible mCherry-HfiA fusions. Sequences corresponding to the hfiA promoter and to the mcherry-hfiA fusion were amplified from chromosomal templates or these plasmid templates, respectively, and then joined by overlap extension

October 2016 Volume 198 Number 19

PCR and cloned into the NdeI and NheI sites of pXGFPC-6. This excised the green fluorescent protein (GFP) gene from the plasmid and inserted a PhfiA-mcherry-hfiA fusion. The 5= end, the hfiA gene in pAF508, is 24 bp longer than that in pAF509, which generates a short linker in the amino acid sequence of the fusion. Strains used in this study are described in Table 2. Plasmids were transformed into C. crescentus strains by electroporation or triparental conjugation as described in reference 18 and selected on solid media containing appropriate antibiotics. The presence of the plasmid was confirmed by PCR. A two-step double recombination strategy based on sucrose counterselection with sacB was used to generate allele replacement strains. For a detailed description, see reference 18. Briefly, the first step of recombination entails selection for integration of the pNPTS138-derived plasmid on medium containing kanamycin or chloramphenicol as appropriate. After a short period of nonselective growth (4 to 20 h), cells in which a second recombination event resulted in plasmid excision were selected by growth on medium containing 3% sucrose. Alleles were confirmed by sequencing gene-specific PCR products. We used this strategy both to replace wild-type alleles with mutant alleles and to replace mutant alleles with the wild type, thereby restoring a native locus. The replicating plasmid, pMT805-hfiA, was cured from the dnaK(I587::Tn5) strain by continuous passaging in the absence of chloramphenicol for 2 to 3 days. Periodically cells were serially diluted and plated on nonselective medium to identify well-isolated colonies. The resulting colonies were patched on medium both with and without chloramphenicol to identify clones which had become sensitive to chloramphenicol.

Journal of Bacteriology

jb.asm.org

2633

Eaton et al.

TABLE 2 Caulobacter crescentus strains used in this study Strain

Genotypea

Construction

FC19 FC1934 FC1935 FC1941 FC1998 ACC510 FC1937 FC1938 FC1940 ACC502 ACC546 FC2413 ACC518 ACC513 FC1389 ACC349 ACC398 ACC387 ACC388 ACC152 ACC452 ACC453 FC2414 ACC515 FC1365 FC2402 ACC501 FC2264 ACC536 ACC528

CB15 CB15/pMT805 CB15/pMT805-hfiA CB15 dnaK(I587::Tn5)/pMT805-hfiA CB15 dnaK(I587::Tn5) CB15 dnaK (restored from FC1998) B5A suppressor/pMT805-hfiA (parent: FC1935) CB15 dnaK(K568*) CB15 dnaK(K568*)/pMT805-hfiA CB15 dnaK (restored from FC1938) B5A dnaK⫹/pMT805-hfiA CB15 dnaK(Q497*) CB15 dnaK(Q497*)/pMT805-hfiA CB15 dnaK (restored from FC2413) CB15 xylX::pMT680 CB15 ⌬hfiA xylX::pMT680 CB15 ⌬hfiA xylX::PhfiA-hfiA-mcherry CB15 ⌬hfiA xylX::PhfiA-mcherry-hfiA CB15 ⌬hfiA xylX::PhfiA-mcherry-hfiA CB15 xylX::PhfiA-hfiA-mcherry CB15 dnaK(K568*) xylX::PhfiA-hfiA-mcherry CB15 dnaK(I587::Tn5) xylX::PhfiA-hfiA-mcherry CB15 dnaK(Q497*) xylX::PhfiA-hfiA-mcherry CB15 dnaK (restored) xylX::PhfiA-hfiA-mcherry CB15 ⌬hfiA CB15 dnaK(K568*) ⌬hfiA CB15 dnaK(Q497*) ⌬hfiA CB15 ⌬lon CB15 dnaK(Q497*) ⌬lon CB15 dnaK(Q497*) ⌬lon xylX::PhfiA-hfiA-mcherry

Wild type

a

Reference or source

Primary selection isolate Plasmid-cured FC1941 Restored dnaK locus in FC1998 with pAF594 Primary selection isolate Made with pAF506 Transformed FC1938 with pAF463 Restored dnaK locus in FC1938 with pAF598 Restored dnaK locus in FC1937 with pAF598 Made with plasmid FC2328 Transformed FC2413 with pAF463 Restored dnaK locus in FC2413 with pAF598 Transformed WT with pMT680 Transformed FC1365 with pMT680 Transformed FC1365 with pAF418 Transformed FC1365 with pAF508 Transformed FC1365 with pAF509 Transformed WT with pAF418 Transformed FC1938 with pAF418 Transformed FC1998 with pAF418 Transformed FC2413 with pAF418 Restored dnaK locus in FC2414 with pAF598 Deleted hfiA fromFC1938 with pAF401 Deleted hfiA from FC2413 with pAF401 Made with plasmid FC1678 Deleted lon from FC2413 with plasmid FC1678 Deleted lon from FC2414 with plasmid FC1678

18 18 This work This work This work This work This work This work This work This work This work This work This work This work This work This work This work This work This work This work This work This work This work 18 This work This work This work This work This work

An asterisk denotes a nonsense mutation.

HfiA suppressor screen. The selection strategy to identify mutants that are insensitive to hfiA overexpression was described in reference 18. Briefly, cells overexpressing the holdfast inhibitor, hfiA, from a high-copynumber plasmid were grown in polystyrene dishes. Cells that were sensitive to HifA are nonadherent cells and were washed away with sterile water. The rare cells that elaborate holdfasts and adhere to the surface served as the inoculum and populate the culture when fresh medium is added. Each selection was enriched with siblings of a small number of winning mutants; thus, this selection strategy was conducted several times to identify a collection of independent mutants. In a pilot selection, we mutagenized FC1935 (CB15/pMT805-hfiA) with the EZ-TN5 R6Kgammaori/KAN-2 transposome kit (catalog no. TSM08KR; Epicentre). The resulting mutagenized library was evaluated in our selection screen. The insertion sites in selected clones were mapped using rescue cloning of the EZ-TN5 transposon by following the manufacturer’s protocol. Subsequent selections relied on spontaneous genetic lesions, which were mapped with whole-genome sequencing as described in reference 18. Crystal violet surface attachment assay. Overnight starter cultures were diluted 1:10, outgrown for 4 to 5 h, and then diluted to an optical density at 660 nm (OD660) of 0.001 (for PYE) or 0.01 (for M2X) in 1 ml of fresh medium in a 24-well culture dish. The lid of the dish was sealed with strips of AeraSeal (Excel Scientific) and grown at 30°C with shaking at 150 rpm overnight to saturation (16 to 18 h for PYE, 20 to 22 h for M2X). Unbound cells were vigorously washed away under tap water. The attached cells in each well were stained with 1.25 ml of 0.01% (wt/vol) crystal violet dissolved in water, with shaking for 5 to 15 min. Unbound stain was removed by vigorous washing with water. Bound stain was extracted with 1.5 ml 100% ethanol, with shaking for 10 min. One hundred microliters of the ethanol extract was diluted in 500 ␮l of ethanol, and the

2634

jb.asm.org

absorbance at 575 nm was measured with a Genesys20 spectrophotometer (ThermoFisher Scientific). Holdfast stain. Holdfasts were detected as described in reference 18. Briefly, starter cultures were diluted to a calculated OD660 of 0.0005 to 0.001 so that after ⬃14 h of growth the culture density would reach ⬃0.05 to 0.1 OD660. Growing cells to low density minimizes cell-cell adhesion and ensures that all cells are born into nearly similar, nutritionally replete conditions. Five hundred microliters of cells was incubated at room temperature with 5 ␮g/ml wheat germ agglutinin (WGA)-Alexa Fluor 594 conjugate (Life Technologies, Molecular Probes) for 5 to 10 min. Cells were then diluted and washed with 1 ml water or medium, collected by centrifugation for 3 min (14,000 ⫻ g), and resuspended in 20 to 30 ␮l of remaining liquid in the tube. Cells were imaged as described below. For each sample, ⬃500 cells were manually scored for the presence or absence of polar fluorescence. Microscopy. Cells were spotted between a glass slide and no. 1 coverslips and imaged with a Leica DM5000 upright microscope in phase contrast and fluorescence modes. We used an HCX PL APO 63⫻/1.4-numeric-aperture Ph3 objective and an EL6000 external mercury halide lamp (Leica) as a fluorescence excitation source. Standard filter sets were used to detect WGA-Alexa 594 (Chroma set 41043). Images were captured with an Orca-ER digital camera (Hamamatsu) using LAS-X (Leica). Western blotting. Overnight starter cultures were diluted to a calculated OD660 of 0.0003 to 0.0005 so that after 16 to 20 h of growth the cells would reach a density of an OD660 of 0.10 to 0.16. For any individual experiment, variation in cell density was limited to ⫾0.02 absorbance unit. Cells were harvested by centrifugation (1 min at ⬃21,000 relative centrifugal force). The volume of culture was normalized so that the number of cells harvested was equivalent to collecting 4 ml of cells at an OD660 of 0.1 (i.e., we harvested cells from 2.5 to 4 ml of the culture depending on

Journal of Bacteriology

October 2016 Volume 198 Number 19

DnaK Affects Caulobacter Surface Attachment

Crystal violet (A575)

5

pMT805-hfiA++

EV

4 3 2 ****

1

B5A dnaK+

dnaK(Q497*)

dnaK(K568*)

B5A spontaneous mut.

dnaK(I587::Tn5)

WT

WT

0

FIG 2 Truncation of DnaK attenuates repression of surface adhesion by hfiA. Surface attachment of C. crescentus cells grown overnight in polystyrene was measured by crystal violet staining and quantified by visible absorption spectrophotometry. Wild-type and dnaK mutant strains harboring an hfiA overexpression plasmid (pMT805-hfiA⫹⫹) were grown in complex medium (PYE) supplemented with 0.1% xylose to induce hfiA overexpression. EV, empty vector control (pMT805). Data represent measurements collected from two independent experiments with 8 cultures of each genotype per experiment (means ⫾ standard deviations [SD]; n ⫽ 16). Mutant strains bearing the hfiA overexpression plasmid were compared to the wild type carrying the same overexpression plasmid by one-way analysis of variance (ANOVA) followed by Dunnett’s multiple-comparisons posttest (****, P ⬍ 0.0001). Symbols: *, nonsense mutation; ⫹, wild-type allele.

the actual OD at the time of collection). Cells were resuspended in 50 ␮l of Triton-based detergent buffer (50 mM Tris-HCl, pH 7.4, 1 mM MgCl2, 150 mM NaCl, 0.1% Triton X-100) supplemented with cOmplete mini protease inhibitor cocktail tablets (Roche) and 0.5 mg/ml DNase I. To each sample, 100 ␮l of loading dye (150 mM Tris, pH 6.8, 6% SDS, 30% glycerol, ⬃0.001% bromphenol blue, 0.75% ␤-mercaptoethanol) was added before heating to 95°C for 5 min. After samples cooled to room temperature, 15 ␮l of each sample was loaded onto a Mini-Protean TGX 4 to 20% gradient gel (Millipore). Proteins were separated at 200 V for ⬃60 min and transferred to an Immobilon-P polyvinylidene difluoride (PVDF) membrane (Bio-Rad) in 25 mM Tris base, 190 mM glycine, 20% methanol at 100 V for 75 min at 4°C. Membranes were blocked with 5% milk in TTBS (10 mM Tris, pH 7.5, 150 mM NaCl, 0.05% Tween 20). Membranes were incubated overnight with either a 1:2,000 dilution of polyclonal rabbit anti-mCherry antisera (provided by Patrick Viollier) or a 1:5,000 dilution of polyclonal rabbit anti-DnaK antisera (provided by Christina Jonas) at room temperature, washed 3 times with TTBS, incubated with a 1:5,000 dilution of monoclonal goat anti-rabbit antibodies conjugated to horseradish peroxidase (HRP) (Thermo Scientific) for 1 h, and washed 5 times. The secondary antibody was detected with SuperSignal West Femto maximum sensitivity substrate (Pierce) imaged using a ChemiDoc MP system (Bio-Rad). Bands were detected and quantified using Image Lab software (Bio-Rad).

RESULTS

Truncation of DnaK attenuates hfiA-induced inhibition of surface attachment. We previously described a suppressor selection strategy to identify the downstream targets of the holdfast inhibitor protein, HfiA (18). Briefly, we selected for mutants that were insensitive to hfiA overexpression, as evidenced by their ability to elaborate holdfasts and attach to a polystyrene surface. In addition to identifying HfiA-suppressing mutations in hfsJ (18), this selection yielded two independent suppressing mutations at the 3= end

October 2016 Volume 198 Number 19

of dnaK. First, in a pilot screen using a Tn5 mutagenized pool, we identified a strain carrying a Tn5 transposon insertion at nucleotide 9924 of the C. crescentus chromosome (NCBI reference sequence NC_011916.1), corresponding to codon I587 of dnaK. This insertion results in the truncation of the final 44 residues of the C-terminal lid (Fig. 1C and D) and the addition of 11 codons before a stop codon is reached. In a subsequent screen for spontaneous mutants, we isolated an additional mutant (named B5A) carrying two spontaneous mutations in dnaK. The first mutation is a nonsense mutation in codon K568, resulting in the truncation of the final 64 residues of DnaK (Fig. 1C and D), and the second, which would have resulted in the nonsynonymous change I579V, is after the nonsense mutation and thus not manifested in the protein sequence. For simplicity the dnaK(K568*, I579V) allele is referred to as dnaK(K568*) here. In the presence of an intact hfiA overexpression plasmid, both of these isolated dnaK mutant strains exhibit elevated surface attachment compared to their wild-type parent as measured by crystal violet staining (Fig. 2). As B5A harbors multiple polymorphisms compared to the wild type (Table 3), we tested whether the dnaK(K568*) allele was sufficient to account for the increase in surface adhesion. We generated a strain with only the dnaK(K568*) allele and transformed this strain with the same hfiA overexpression plasmid used in our initial screen. This strain phenocopied the original B5A suppressor strain exhibiting the same increased surface adhesion compared to the wild type (Fig. 2). Furthermore, to evaluate if any other mutations in the B5A background affect surface attachment, we restored the mutant dnaK locus in the B5A background to the wildtype allele, leaving all other mutations in the background intact. For comparison, we then added the same hfiA overexpression plasmid used in the screen. This B5A dnaK⫹ strain phenocopied the wild type (Fig. 2). We conclude that the dnaK mutations are the primary determinant of the elevated adhesion phenotype in the B5A mutant strain. To further characterize the role of dnaK in modulation of cell adhesion, we generated strains harboring additional mutant alleles of dnaK. Given that dnaK is essential in C. crescentus (19, 20), we pursued a strategy to generate a minimal functional allele of dnaK. Based on the high-density Tn sequencing data of Christen et al. (20), we noted that C. crescentus tolerates Tn5 insertions at the 3= end of dnaK, after codon 496. As such, we engineered a strain with a nonsense mutation in codon 497. Notably, this is the site of two dnaK frameshift mutations reported to suppress the phenotypic effects of dnaA overexpression (34). Among the 25

TABLE 3 Polymorphisms in hfiA suppressor strain B5A Locus no.a

Annotation

Amino acid change(s)c

CC_0010/CCNA_00010 CC_0649/CCNA_00686

DnaK chaperone Peptidase family S41 protein Methyltransferase Hypothetical Aminopeptidase N

K568*, I579V T252A

CC_1327/CCNA_01388 CC_3523b CC_3565/CCNA_03680

T234A P40S M927T

a Corresponding to C. crescentus strain CB15 (CC) and strain NA1000 (CCNA) genome annotations (GenBank accession numbers NC_002696.2 and NC_011916.1, respectively). b This hypothetical open reading frame is not annotated in strain NA1000. The polymorphism is located at nucleotide coordinate 3,798,388 in the NA1000 genome (accession number NC_011916.1). c An asterisk denotes a nonsense mutation.

Journal of Bacteriology

jb.asm.org

2635

Eaton et al.

2636

jb.asm.org

A 4 3 2

C WT

(I587::Tn5) (I587::Tn5)

(Q497*)

dnaK(Q497*) dnaK(Q497*)

(K568*)

dnaK(K568*)

dnaK(I587::Tn5)

dnaK(K568*)

1 0

dnaK locus restored

****

40 30 **

20

(Q497*)

(K568*)

0

WT

10

dnaK(I587::Tn5)

B

****

WT

Crystal violet (A575)

5

% of cells with holdfast

dnaK mutant alleles recovered by Jonas et al. (34) that result in reduced levels of dnaA, the frameshift at Q497 (Q497fs) is the maximally truncated allele. These two independent screens suggest that this allele, which results in the loss of the entire alphahelical lid domain, is the maximal nonlethal truncation of the dnaK in C. crescentus (Fig. 1C and D). Our strain bearing a nonsense codon at position 497, dnaK(Q497*), exhibits increased surface adhesion in the presence of the hfiA overexpression plasmid (P ⬍ 0.0001), similar to the dnaK mutant strains identified in our hfiA suppressor screen (Fig. 2). DnaK truncation enhances surface attachment and holdfast development. We next tested if truncation of the DnaK substratebinding lid affected surface attachment in an otherwise wild-type genetic background, i.e., in the absence of the hfiA overexpression plasmid. Specifically, we evaluated surface adhesion and holdfast development in the dnaK::Tn5 strain cured of the hfiA overexpression plasmid and in the two engineered dnaK 3= lid truncation strains. In complex peptone-yeast extract (PYE) medium, dnaK mutant strains are significantly enhanced in surface attachment compared to the wild type (Fig. 3A). We note a trend of increasing adhesion that correlates with increasing truncation of the substrate-binding lid domain of DnaK. Using standard two-step recombination, we restored the dnaK locus to the wild-type allele in each of our mutant strains. This restoration of dnaK fully complemented the hyperadhesive phenotype and restored attachment to wild-type levels (Fig. 3A). To test whether elevated bulk surface attachment correlated with an increase in holdfast development, we sought to directly measure the fraction of cells that displayed a holdfast in each of our dnaK mutant strains by staining with fluorescent wheat germ agglutinin (WGA). This lectin binds to N-acetylglucosaminyl residues and has been used to specifically label holdfast (7). In logarithmic phase in PYE medium most wild-type cells develop a holdfast (18), which makes it difficult to detect an increase in the fraction of cells with a holdfast. Thus, to assess whether the dnaK mutants show increased holdfast development, we turned to a minimal defined medium (M2-xylose) in which wild-type cells elaborate holdfast at a low frequency (18). Under this condition, we observe that truncation of DnaK results in a significant increase in the fraction of cells with a holdfast (P ⬍ 0.01 to P ⬍ 0.0001) (Fig. 3B and C). Greater truncations of the C terminus correlate with an increased frequency of holdfast development, consistent with our bulk surface attachment measurements. Again, restoration of the dnaK locus to the wild-type allele fully complements holdfast development to wild-type levels (Fig. 3B). DnaK stabilizes the holdfast inhibitor, HfiA. Clearly, the capacity of hfiA to function as an adhesin inhibitor is attenuated in strains harboring truncated dnaK alleles. This phenotype could be due to DnaK-dependent effects on downstream targets of HfiA or could be a result of reduced steady-state levels of functional HfiA in the cell. Given that (i) HfiA is a small, largely hydrophobic protein and a potent inhibitor of adhesion (18), (ii) DnaK has a high affinity for exposed hydrophobic protein regions (26–28), and (iii) truncation of the substrate-binding lid of DnaK results in increased surface adhesion, we hypothesized that DnaK plays a role in stabilizing HfiA and, thus, controlling the level of functional HfiA in the cell. Efforts to generate polyclonal antiserum to directly detect HfiA by immunoblotting were unsuccessful due to difficulties in protein expression and purification. Alternatively, we sought to assess the level of HfiA in the cell by

dnaK locus restored

dnaK(Q497*)

2 µm

FIG 3 Truncation of DnaK results in increased cell surface adhesion and holdfast development. (A) Crystal violet surface attachment assay as described for Fig. 2 after overnight growth in complex medium (PYE). Data represent measurements collected from two independent experiments with 8 independent cultures in each experiment (means ⫾ SD; n ⫽ 16). (B) Percentage of cells displaying a fluorescent, lectin-stained (Alexa 595-WGA) holdfast after overnight growth in minimal medium (M2X) to an OD660 of 0.1. Data represent 4 to 6 independent samples per strain (means ⫾ SD; counting ⬇500 cells per sample). (C) Phase contrast and fluorescence micrographs of lectin (WGA-Alexa 595)-stained holdfast on C. crescentus cells. White arrows indicate fluorescently stained holdfast; outlined arrows show cells attached by their holdfast in a small rosette. In panels A and B, dnaK mutants were compared to the wild type by one-way ANOVA, followed by Dunnett’s multiple-comparison posttest (**, P ⬍ 0.01; ****, P ⬍ 0.0001). Strains in which the truncated mutant allele of dnaK was restored to the wild type are on the far right of each panel, and the mutant background that was restored to wild type is labeled. An asterisk denotes a nonsense mutation.

generating tagged HfiA fusions. An ideal tag minimally disrupts function and enables detection by immunoblotting. We evaluated fusions with the small hemagglutinin (HA) epitope tag and with fluorescent proteins, GFP and mCherry, each expressed from an

Journal of Bacteriology

October 2016 Volume 198 Number 19

DnaK Affects Caulobacter Surface Attachment

mCherry

hfiA 34.9 kDa

fiA

m

m

C

C

hy

hy

H

H

fiA

hy

id

C

m

m

as pl

fiA

o

H

N

50 37

ns fusion mChy

25

1

50 37

0

25

EV

H

m

C

hy

hy C

fiA

fiA H

m as m

N

o

pl

m

C

hy

M2X

2

id

PYE

fiA

hfiA 35.7 kDa

PYE

H

mCherry

C

∆hfiA

mChy HfiA

PhfiA

mCherry 36.1 kDa

WT

mChy HfiA

PhfiA

hfiA

5 4 3 2 1 0

HfiA mChy

PhfiA

B Crystal violet stain (A575)

A

ns fusion mChy M2X

FIG 4 HfiA fusions functionally complement the ⌬hfiA hyperadhesion phenotype. (A) Schematic of three different hfiA fusions to the fluorescent protein encoded by mcherry. Each fusion is expressed from the hfiA promoter (PhfiA). Short linker sequences between the coding regions are marked with a caret symbol. For the first fusion, the linker represents sequence in the multiple-cloning site; for the second, the linker corresponds to 24 nucleotides upstream of the hfiA start codon. Expected molecular masses for each fusion protein are indicated. (B) C. crescentus cell attachment to polystyrene after overnight growth in complex PYE medium (top) or defined M2X medium (bottom) was quantified by spectrophotometric measurement of crystal violet stain. Wild-type and ⌬hfiA strains carrying empty plasmids (EV) are shown in gray and white bars, respectively. ⌬hfiA strains carrying chromosomally integrated copies of each of the three mcherry fusions are shown in pink. Data represent means ⫾ SD (n ⫽ 8). mChy, mCherry. (C) Immunoblot of each mCherry fusion in C. crescentus lysate detected with polyclonal anti-mCherry antiserum. Cells were grown in PYE (top) or M2X (bottom). The full-length fusion protein and a degradation product corresponding to the size of mCherry (⬃27 kDa) are detected in strains expressing the fusion but not in C. crescentus lysates with no plasmid. A nonspecific (ns) band near 50 kDa was detected in all samples and used to normalize loading in subsequent quantitative experiments. Molecular mass standards (in kilodaltons) are marked on the left of the blot.

hfiA promoter on a plasmid integrated at the xylose (xylX) locus. We observed that the tag, and the position of the tag, affect the ability of the fusion protein to complement the ⌬hfiA hyperadhesive phenotype. This is presumably due to differing stabilities of the fusion proteins (see Fig. S1 in the supplemental material). The HA and GFP fusions fail to complement the null phenotype in complex medium. In minimal medium, the HfiA-GFP fusion complements the null phenotype but was not detected with our anti-GFP antibody. We focused our efforts on the mCherry fusions, which were robustly detected with anti-mCherry antisera (Fig. 4C). Expression of each of the three mCherry fusions attenuated adhesion in a ⌬hfiA background compared to the empty vector control and thereby functionally complemented the ⌬hfiA hyperadhesion phenotype (Fig. 4B; see also Fig. S1 in the supplemental material). We note that all three mCherry fusions reduced C. crescentus surface adhesion to below wild-type levels. From this, we infer that HfiA-mCherry fusions are more stable than the untagged native HfiA protein. Indeed, the fusions that resulted in the most marked reduction in adhesion are present in higher levels in the cell as assessed by immunoblotting (Fig. 4C). The adhesive properties of the ⌬hfiA strain bearing the C-terminal HfiAmCherry fusion more closely resembled the wild type than the strains bearing N-terminal fusions (Fig. 4B). Thus, we proceeded to characterize the effects of DnaK truncation on steady-state levels of this HfiA fusion. We introduced the plasmid containing the HfiA-mCherry fusion into the dnaK mutant strains and then assessed steady-state levels of this fusion in each dnaK mutant background by immunoblotting. Strains bearing truncated dnaK alleles exhibited reduced levels of HfiA-mCherry relative to wild-type C. crescentus in both minimal and complex media (Fig. 5A and B). Mutant strains with smaller truncations of the C terminus of DnaK did not show

October 2016 Volume 198 Number 19

(statistically) significantly lower levels of HfiA-mCherry, although levels of fusions did trend lower; the precision of this assay is not ideal for detection of small changes in protein levels. However, deletion of the entire alpha-helical lid domain from the C terminus of DnaK resulted in ⬇40% reduction of steady-state HfiAmCherry levels in cells grown in minimal medium (P ⬍ 0.0001) and ⬇25% reduction in cells grown in complex medium (P ⬍ 0.05). This defect in protein level can be complemented by restoration of the truncated dnaK locus to wild-type dnaK (Fig. 5). The trend in HfiA-mCherry protein levels across our DnaK truncation strains is inversely correlated with the trend in holdfast development and bulk surface adhesion (Fig. 3 and 5). Moreover, the dnaK mutant strains bearing the hfiA-mcherry fusion show enhanced surface attachment with increasing truncation of dnaK (see Fig. S2 in the supplemental material), similar to strains without the HfiA-mCherry fusion (Fig. 3). These observations are consistent with the function of HfiA as a holdfast inhibitor and provide evidence that protein chaperone functions can control activity of HfiA at the posttranslational level by stabilizing HfiA. In this manner, DnaK function is required for proper developmental control of the holdfast surface adhesin and regulated surface attachment. To evaluate if DnaK truncation has HfiA-independent effects on holdfast formation, we deleted hfiA in the strains bearing dnaK(K568*) and dnaK(Q497*) alleles. Deletion of hfiA results in derepression of holdfast synthesis and holdfast elaboration on nearly all cells (18). Indeed, both dnaK(K568*) ⌬hfiA and dnaK(Q497*) ⌬hfiA strains displayed holdfast on almost every cell, similar to dnaK⫹ ⌬hfiA cells (Fig. 6). These epistasis results support the model that HfiA is downstream of DnaK and that the effects of DnaK truncation on holdfast formation are dependent on HfiA. However, because nearly all ⌬hfiA cells elab-

Journal of Bacteriology

jb.asm.org

2637

Eaton et al.

B

dnaK+ restored

Q497*

K568*

*

1.5 1.0

dnaK locus restored

dnaK(Q497*)

dnaK(K568*)

0.5 0.0

dnaK locus restored

dnaK(Q497*)

dnaK(K568*)

dnaK(I587::Tn5)

0.5

HfiA mChy

dnaK(I587::Tn5)

1.0

ns

2.0 HfiA-mCherry fusion (relative intensity)

****

WT

HfiA-mCherry fusion (relative intensity)

50 37

D

1.5

0.0

I587::Tn5

WT 50 37

ns HfiA mChy

C

PYE

dnaK+ restored

Q497*

K568*

I587::Tn5

WT

M2X

WT

A

FIG 5 Steady-state levels of HfiA-mCherry are reduced when the C-terminal lid domain of DnaK is truncated. (A and B) Representative immunoblots of HfiA-mCherry expressed from the hfiA promoter. Cultures were grown in M2X (A) or PYE (B). The dnaK allele of each strain is indicated. Symbols: *, nonsense mutation; ⫹, wild-type allele. mChy, mCherry. (C and D) Quantification of HfiA-mCherry detected in independent blots from cells grown in M2X (C) or PYE (D). The intensity of the band corresponding to the fusion was normalized to the nonspecific band in that sample (Fig. 4) and then normalized to the wild type on each blot. Error bars represent means ⫾ standard errors of the means (n ⫽ 12 independent blots for each growth medium, and each blot contains one independent culture of each strain). Statistical differences of means were evaluated by one-way ANOVA followed by Dunnett’s posttest (****, P ⬍ 0.0001; *, P ⬍ 0.0).

orate a holdfast, a further increase upon truncation of dnaK would be virtually impossible to detect. Thus, we cannot rule out the possibility that dnaK truncation has other effects on adhesion. Truncation of the DnaK C-terminal lid domain does not destabilize the protein. Truncation of the DnaK lid correlates with destabilization of the HfiA-mCherry fusion and increased hold-

WT

dnaK(K568*)

dnaK(Q497*)

∆hfiA

dnaK(K568*) ∆hfiA

dnaK(Q497*) ∆hfiA

5 µm

FIG 6 DnaK truncation does not affect holdfast development in strains lacking hfiA. Representative phase contrast and fluorescence micrographs of cells stained with WGA-Alexa 595. Holdfasts are apparent as bright foci as shown in Fig. 3. Strains have either wild-type hfiA (top row) or a ⌬hfiA mutant (bottom row) combined with either wild-type or truncated alleles of dnaK.

2638

jb.asm.org

fast formation and surface attachment. However, these results do not distinguish whether the lid domain is required for wild-type function of DnaK by enabling chaperone activity or by determining DnaK protein stability in the cell. Biochemical characterization of lidless E. coli DnaK indicates that the lid promotes activity by enhancing substrate binding (32, 33). To test whether removal of the C-terminal lid domain affects DnaK stability, we evaluated steady-state levels of each DnaK allele by immunoblotting with anti-DnaK antiserum (Fig. 7). In both complex and defined media, we observe that the lid domain does not significantly affect steady-state levels of DnaK in the cell. The effects of DnaK on HfiA do not require the Lon protease. We envision at least two mechanistic models that are consistent with the genetic and molecular connections we observe between DnaK, HfiA, and holdfast development. In the first model, HfiA is a direct client of DnaK and the stabilization of HfiA by DnaK is compromised by deletion of the lid domain. Alternatively, the effects of DnaK on HfiA stability could be indirectly mediated by the Lon protease as follows. Defects in DnaK chaperone activity result in accumulation of misfolded proteins, which in turn stimulates Lon protease (34). Thus, Lon stimulation in strains with truncated DnaK could result in enhanced proteolysis of HfiA. To discriminate between these models, we sought to evaluate whether the observed effects of DnaK truncation on HfiA proteostasis and holdfast development require lon. To this end, we deleted lon in strains bearing the dnaK(Q497*) allele. If HfiA destabilization requires Lon activation, then the dnaK(Q497*) ⌬lon strain should exhibit restored HfiA stability and reduced holdfast development

Journal of Bacteriology

October 2016 Volume 198 Number 19

10

dnaK(Q497*)

HfiA mChy

B

0

WT

dnaK(Q497*)

∆lon

dnaK(Q497*) ∆lon

Q497*

0

K568*

1

Q497*

1

K568*

2

I587::Tn5

2

I587::Tn5

M2X

WT

PYE

WT

DnaK (relative intensity)

B

50 37

ns dnaK(Q497*) ∆lon

M2X

0

∆lon

50 PYE

WT

20

dnaK(Q497*) ∆lon

30

dnaK(Q497*)

DnaK

40

WT

ns

75

C

**

% of cells with holdfast

A

Q497*

K568*

I587::Tn5

WT

Q497*

K568*

WT

A

I587::Tn5

DnaK Affects Caulobacter Surface Attachment

FIG 7 Truncation of DnaK does not significantly affect steady-state protein levels. (A) Representative immunoblot of strains bearing wild-type or truncated alleles of dnaK probed with anti-DnaK antiserum. Cultures were grown in PYE (left) or M2X (right). The expected molecular masses of the truncated proteins are the following: wild type, 68 kDa; I587::Tn5, 64 kDa; K568*, 61 kDa; and Q497*, 54 kDa. Molecular mass standards are marked on the left (in kilodaltons). A nonspecific (ns) band above 75 kDa is shown as a loading control. (B) Quantification of DnaK in four independent blots, each with independent samples, normalized to the nonspecific band. Data represent means ⫾ SD (n ⫽ 4). Means were evaluated by one-way ANOVA followed by Dunnett’s posttest. None were statistically different from the wild type (P ⬎ 0.05 for all comparisons).

5 µm

compared to the dnaK(Q497*) strain alone. However, holdfast development is not reduced and HfiA-mCherry is not stabilized in dnaK(Q497*) ⌬lon cells compared to dnaK(Q497*) lon⫹ cells (Fig. 8). These data are inconsistent with the model that dnaK truncation indirectly modulates adhesion development via activation of Lon protease. Similar to previous reports (34–36), our ⌬lon strains exhibit profound pleiotropic developmental defects, including a high frequency of filamentous cells with multiple constrictions arising from incomplete cell divisions and notable occurrences of yshaped cells, presumably arising from inappropriate positioning of polar development proteins (Fig. 8). Consistent with these developmental defects, cells lacking lon can be found with mislocalized holdfasts along the sides of the cell rather than at the pole. Moreover, lon deletion increases the probability of holdfast development irrespective of the dnaK allele (Fig. 8). Thus, cells lacking both the lid domain of DnaK and the Lon protease display pleiotropic developmental defects owing to the loss of Lon and increased holdfast development similar to strains bearing only single mutations (Fig. 8). The contribution of each mutation to the dysregulation of holdfast development in the double mutant strain is difficult to discern. DISCUSSION

DnaK modulates holdfast adhesin development. Molecular chaperones fulfill a range of functions related to establishment

October 2016 Volume 198 Number 19

FIG 8 Deletion of lon does not restore wild-type holdfast development or HifA-mCherry stability in strains lacking the DnaK lid domain. (A) Percentage of cells displaying holdfasts grown as described for Fig. 3. Data represent means ⫾ SD from 4 to 5 independent cultures per genotype, counting ⬃500 cells per sample. Means were compared by one-way ANOVA followed by Dunnett’s posttest. All mutants are significantly different from the wild type (**, P ⬍ 0.01). None of the mutants were significantly different from each other. (B) Representative micrographs of cells stained with WGA-Alexa 595 to identify holdfasts. Rosettes of cells attached at the holdfast are indicated with white arrowheads. Orange arrowheads indicate y-shaped cells where a stalked pole has emerged at the midcell. The yellow arrowhead indicates a fluorescent focus (presumably a holdfast) on the side of a cell body. (C) Representative immunoblot of HfiA-mCherry expressed from the hfiA promoter in cultures grown in M2X as described for Fig. 5. Deletion of lon in cells bearing the dnaK(Q497*) allele does not restore steady-state levels of HfiAmCherry. mChy, mCherry.

and maintenance of the proteome by facilitating de novo folding and refolding of denatured proteins. In an unbiased forward genetic screen for hyperadhesive C. crescentus mutants, we identified two independent mutations in the 3= end of dnaK. Strains harboring these mutations had higher levels of surface attachment and increased frequency of holdfast development. We further demonstrated that systematic truncation of the 3= end of dnaK, which encodes the C-terminal alpha-helical lid domain, resulted in decreased levels of HfiA in the cell. This result is consistent with the elevated adhesion and holdfast phenotypes we observe in these strains.

Journal of Bacteriology

jb.asm.org

2639

Eaton et al.

We have previously shown that small relative changes in hfiA transcription have large phenotypic consequences with respect to the frequency of holdfast development at the single-cell level and bulk adhesion at the population level (18). Here, we provide evidence that a small relative decrease in HfiA protein level corresponds with increased frequency of holdfast development and surface adhesion, and that the C-terminal lid domain of DnaK is a molecular determinant of HfiA proteostasis under standard laboratory growth conditions. These data lend further support to a model in which HfiA can modulate C. crescentus cell adhesion in an almost binary fashion across a relatively low (⬇2- to 3-fold) dynamic range of expression. Certainly, HfiA concentration alone may not fully determine its phenotypic effects. HfiA, in concert with HfsJ, may require additional layers of posttranslational control that ensure robust regulation of cell adhesion. The lid domain of DnaK is dispensable for growth but required for proper control of HfiA levels and cell adhesion. In contrast to E. coli, DnaK is essential for growth in C. crescentus (19). However, removing the alpha-helical lid domain at the C terminus of DnaK does not have a significant effect on C. crescentus growth; three independent genetic studies have identified viable strains bearing transposon insertions (20), frameshift mutations (34), or nonsense mutations (this work) in the linker that connect the lid domain to the substrate binding beta sheet domain. These results are consistent with studies of E. coli demonstrating that lid-less DnaK retains functionality, albeit at a reduced level. Specifically, C-terminally truncated DnaK maintains ATP-dependent binding of peptide substrates, undergoes substrate-activated ATP-hydrolysis in vitro, and supports replication of bacteriophage ␭ in vivo (32, 33). While the lid domain is not necessary for substrate binding, it does serve to enhance affinity for substrate and increase the lifetime of DnaK-substrate complexes. Our data provide evidence that loss of the lid domain of C. crescentus DnaK, though not lethal, does compromise DnaK function with regard to its role in determining cell adhesiveness. Lid-less DnaK results in reduced HfiA levels in the cell, which ultimately has a large effect on holdfast development and surface adhesion. Our data support a model in which the small hydrophobic protein, HfiA, is a direct client of DnaK. In such a model, the HfiA-DnaK interaction is compromised in strains missing the lid domain. As a consequence, HfiA is insufficiently stabilized and cleared from the cell. However, our genetic results do not exclude the possibility of indirect effects of DnaK on HfiA stability. We note that models in which DnaK directly or indirectly affects HfiA and holdfast development are not mutually exclusive. Efforts to evaluate a model in which DnaK truncation stimulates Lon-mediated proteolysis of HfiA are complicated by the pleiotropic effects of lon deletion. In addition to developmental defects, cells lacking lon exhibit increased holdfast development. At this time, we cannot discern whether this is due to deficient hfiA transcription as a consequence of stabilization of SciP and, thus, inhibition of CtrA (36), a positive regulator of hfiA transcription (18), decreased functioning of DnaK due to an overload of misfolded proteins that cannot be cleared, or other effects of a misregulated cell cycle. When lon deletion is combined with dnaK truncation, the effect of each mutation on holdfast development becomes difficult to disentangle. Future in vitro characterization of the interactions between DnaK and cellular proteases with HfiA will be

2640

jb.asm.org

necessary to more fully understand the molecular regulation of holdfast development. Molecular chaperones are important regulators of biofilm development. Molecular chaperones, including DnaK, are highly expressed in single-species biofilms cultivated in laboratory settings (37, 38) and in natural biofilm communities in situ (39). In a clinical context, chaperones are critical for development and maintenance of biofilms in species that cause human disease, including E. coli, Staphylococcus aureus, Streptococcus mutans, and Listeria monocytogenes (40–43). In the marine symbiont Vibrio fischeri, DnaK is required for appropriate host colonization and biofilm formation in the host squid light organ (44). This broad reliance on DnaK in the formation of biofilms across a range of bacterial taxa suggests that proteostasis is a challenge that bacteria must overcome when residing in surface-attached communities. Indeed, chaperones have been implicated as determinants of biofilm resistance to antibiotics and other antimicrobial agents (38, 43). High expression of chaperones in biofilms mitigates the destabilizing effects that high-density growth conditions have on the bacterial proteome. As such, DnaK is an attractive target for pharmacological destabilization of biofilms (40). However, in contrast to other studied bacterial species, where impairment of DnaK inhibits biofilm formation and maintenance, impairment of C. crescentus DnaK enhances biofilm formation. The dependence of HfiA levels on DnaK suggests that this molecular chaperone provides the cell with an additional mechanism to control holdfast development and surface adherence. Proteotoxic conditions strongly induce DnaK expression in C. crescentus (34, 45–48). In a model where HfiA is a client for DnaK, increased levels of DnaK are predicted to stabilize HfiA and reduce the number of newborn cells that develop holdfasts. This would decrease the probability of attachment in proteotoxic environments and foster dispersal of progeny from biofilms that have entered a “stressful” phase. In short, we posit that direct control of HfiA stability by DnaK is a posttranslational regulatory mechanism that tunes adhesion in response to environmental conditions. ACKNOWLEDGMENTS We thank Patrick Viollier for generously providing antiserum for mCherry and Kristina Jonas for generously providing antiserum for DnaK. We also thank Jon Henry for building the pNPTS138-⌬lon plasmid.

FUNDING INFORMATION This work, including the efforts of Sean Crosson, was funded by HHS | National Institutes of Health (NIH) (R01 GM087353-06).

REFERENCES 1. Loeb GI, Neihof RA. 1975. Marine conditioning films. Adv Chem 145: 319 –335. http://dx.doi.org/10.1021/ba-1975-0145.ch016. 2. Taylor GT, Zheng D, Lee M, Troy PJ, Gyananath G, Sharma SK. 1997. Influence of surface properties on accumulation of conditioning films and marine bacteria on substrata exposed to oligotrophic waters. Biofouling 11:31–57. http://dx.doi.org/10.1080/08927019709378319. 3. Schneider RP, Leis A. 2003. Conditioning films in aquatic environments. Encyclopedia of environmental microbiology. John Wiley & Sons, Inc, New York, NY. http://dx.doi.org/10.1002/0471263397.env036. 4. Garg A, Jain A, Bhosle NB. 2009. Chemical characterization of a marine conditioning film. Int Biodeterior Biodegradation 63:7–11. http://dx.doi .org/10.1016/j.ibiod.2008.05.004. 5. Marshall K. 1996. Adhesion as a strategy for access to nutrients. In

Journal of Bacteriology

October 2016 Volume 198 Number 19

DnaK Affects Caulobacter Surface Attachment

6. 7. 8. 9. 10.

11. 12. 13.

14.

15.

16.

17.

18.

19.

20. 21.

22.

23. 24.

25.

Fletcher M (ed), Bacterial adhesion: molecular and ecological diversity. Wiley, New York, NY. Poindexter JS. 1964. Biological properties and classification of the Caulobacter group. Bacteriol Rev 28:231–295. Merker RI, Smit J. 1988. Characterization of the adhesive holdfast of marine and freshwater Caulobacters. Appl Environ Microbiol 54:2078 – 2085. Ong CJ, Wong ML, Smit J. 1990. Attachment of the adhesive holdfast organelle to the cellular stalk of Caulobacter crescentus. J Bacteriol 172: 1448 –1456. Tsang PH, Li G, Brun YV, Freund LB, Tang JX. 2006. Adhesion of single bacterial cells in the micronewton range. Proc Natl Acad Sci U S A 103: 5764 –5768. http://dx.doi.org/10.1073/pnas.0601705103. Berne C, Ma X, Licata NA, Neves BR, Setayeshgar S, Brun YV, Dragnea B. 2013. Physiochemical properties of Caulobacter crescentus holdfast: a localized bacterial adhesive. J Phys Chem B 117:10492–10503. http://dx .doi.org/10.1021/jp405802e. Kurtz HD, Jr, Smith J. 1992. Analysis of a Caulobacter crescentus gene cluster involved in attachment of the holdfast to the cell. J Bacteriol 174: 687– 694. Kurtz HD, Jr, Smit J. 1994. The Caulobacter crescentus holdfast: identification of holdfast attachment complex genes. FEMS Microbiol Lett 116: 175–182. http://dx.doi.org/10.1111/j.1574-6968.1994.tb06697.x. Cole JL, Hardy GG, Bodenmiller D, Toh E, Hinz A, Brun YV. 2003. The HfaB and HfaD adhesion proteins of Caulobacter crescentus are localized in the stalk. Mol Microbiol 49:1671–1683. http://dx.doi.org/10.1046/j .1365-2958.2003.03664.x. Smith CS, Hinz A, Bodenmiller D, Larson DE, Brun YV. 2003. Identification of genes required for synthesis of the adhesive holdfast in Caulobacter crescentus. J Bacteriol 185:1432–1442. http://dx.doi.org/10.1128/JB .185.4.1432-1442.2003. Toh E, Kurtz HD, Jr, Brun YV. 2008. Characterization of the Caulobacter crescentus holdfast polysaccharide biosynthesis pathway reveals significant redundancy in the initiating glycosyltransferase and polymerase steps. J Bacteriol 190:7219 –7231. http://dx.doi.org/10.1128/JB.01003-08. Hardy GG, Allen RC, Toh E, Long M, Brown PJ, Cole-Tobian JL, Brun YV. 2010. A localized multimeric anchor attaches the Caulobacter holdfast to the cell pole. Mol Microbiol 76:409 – 427. http://dx.doi.org/10.1111/j .1365-2958.2010.07106.x. Li G, Brown PJ, Tang JX, Xu J, Quardokus EM, Fuqua C, Brun YV. 2012. Surface contact stimulates the just-in-time deployment of bacterial adhesins. Mol Microbiol 83:41–51. http://dx.doi.org/10.1111/j.1365-2958 .2011.07909.x. Fiebig A, Herrou J, Fumeaux C, Radhakrishnan SK, Viollier PH, Crosson S. 2014. A cell cycle and nutritional checkpoint controlling bacterial surface adhesion. PLoS Genet 10:e1004101. http://dx.doi.org/10 .1371/journal.pgen.1004101. da Silva AC, Simao RC, Susin MF, Baldini RL, Avedissian M, Gomes SL. 2003. Downregulation of the heat shock response is independent of DnaK and sigma32 levels in Caulobacter crescentus. Mol Microbiol 49: 541–553. http://dx.doi.org/10.1046/j.1365-2958.2003.03581.x. Christen B, Abeliuk E, Collier JM, Kalogeraki VS, Passarelli B, Coller JA, Fero MJ, McAdams HH, Shapiro L. 2011. The essential genome of a bacterium. Mol Syst Biol 7:528. Calloni G, Chen T, Schermann SM, Chang HC, Genevaux P, Agostini F, Tartaglia GG, Hayer-Hartl M, Hartl FU. 2012. DnaK functions as a central hub in the E. coli chaperone network. Cell Rep 1:251–264. http: //dx.doi.org/10.1016/j.celrep.2011.12.007. Kim YE, Hipp MS, Bracher A, Hayer-Hartl M, Hartl FU. 2013. Molecular chaperone functions in protein folding and proteostasis. Annu Rev Biochem 82:323–355. http://dx.doi.org/10.1146/annurev -biochem-060208-092442. Mayer MP. 2013. Hsp70 chaperone dynamics and molecular mechanism. Trends Biochem Sci 38:507–514. http://dx.doi.org/10.1016/j.tibs.2013.08 .001. Castanie-Cornet MP, Bruel N, Genevaux P. 2014. Chaperone networking facilitates protein targeting to the bacterial cytoplasmic membrane. Biochim Biophys Acta 1843:1442–1456. http://dx.doi.org/10.1016/j .bbamcr.2013.11.007. Clerico EM, Tilitsky JM, Meng W, Gierasch LM. 2015. How hsp70 molecular machines interact with their substrates to mediate diverse physiological functions. J Mol Biol 427:1575–1588. http://dx.doi.org/10.1016 /j.jmb.2015.02.004.

October 2016 Volume 198 Number 19

26. Gragerov A, Zeng L, Zhao X, Burkholder W, Gottesman ME. 1994. Specificity of DnaK-peptide binding. J Mol Biol 235:848 – 854. http://dx .doi.org/10.1006/jmbi.1994.1043. 27. Zhu X, Zhao X, Burkholder WF, Gragerov A, Ogata CM, Gottesman ME, Hendrickson WA. 1996. Structural analysis of substrate binding by the molecular chaperone DnaK. Science 272:1606 –1614. http://dx.doi.org /10.1126/science.272.5268.1606. 28. Rudiger S, Germeroth L, Schneider-Mergener J, Bukau B. 1997. Substrate specificity of the DnaK chaperone determined by screening cellulose-bound peptide libraries. EMBO J 16:1501–1507. http://dx.doi.org/10 .1093/emboj/16.7.1501. 29. Bertelsen EB, Chang L, Gestwicki JE, Zuiderweg ER. 2009. Solution conformation of wild-type E. coli Hsp70 (DnaK) chaperone complexed with ADP and substrate. Proc Natl Acad Sci U S A 106:8471– 8476. http: //dx.doi.org/10.1073/pnas.0903503106. 30. Kityk R, Kopp J, Sinning I, Mayer MP. 2012. Structure and dynamics of the ATP-bound open conformation of Hsp70 chaperones. Mol Cell 48: 863– 874. http://dx.doi.org/10.1016/j.molcel.2012.09.023. 31. Mayer MP, Kityk R. 2015. Insights into the molecular mechanism of allostery in Hsp70s. Front Mol Biosci 2:58. 32. Pellecchia M, Montgomery DL, Stevens SY, Vander Kooi CW, Feng HP, Gierasch LM, Zuiderweg ER. 2000. Structural insights into substrate binding by the molecular chaperone DnaK. Nat Struct Biol 7:298 –303. http://dx.doi.org/10.1038/74062. 33. Buczynski G, Slepenkov SV, Sehorn MG, Witt SN. 2001. Characterization of a lidless form of the molecular chaperone DnaK: deletion of the lid increases peptide on- and off-rate constants. J Biol Chem 276:27231– 27236. http://dx.doi.org/10.1074/jbc.M100237200. 34. Jonas K, Liu J, Chien P, Laub MT. 2013. Proteotoxic stress induces a cell-cycle arrest by stimulating Lon to degrade the replication initiator DnaA. Cell 154:623– 636. http://dx.doi.org/10.1016/j.cell.2013.06 .034. 35. Wright R, Stephens C, Zweiger G, Shapiro L, Alley MR. 1996. Caulobacter Lon protease has a critical role in cell-cycle control of DNA methylation. Genes Dev 10:1532–1542. http://dx.doi.org/10.1101/gad.10.12.1532. 36. Gora KG, Cantin A, Wohlever M, Joshi KK, Perchuk BS, Chien P, Laub MT. 2013. Regulated proteolysis of a transcription factor complex is critical to cell cycle progression in Caulobacter crescentus. Mol Microbiol 87: 1277–1289. http://dx.doi.org/10.1111/mmi.12166. 37. De Angelis M, Siragusa S, Campanella D, Di Cagno R, Gobbetti M. 2015. Comparative proteomic analysis of biofilm and planktonic cells of Lactobacillus plantarum DB200. Proteomics 15:2244 –2257. http://dx.doi .org/10.1002/pmic.201400363. 38. Whiteley M, Bangera MG, Bumgarner RE, Parsek MR, Teitzel GM, Lory S, Greenberg EP. 2001. Gene expression in Pseudomonas aeruginosa biofilms. Nature 413:860 – 864. http://dx.doi.org/10.1038/35101627. 39. Ram RJ, Verberkmoes NC, Thelen MP, Tyson GW, Baker BJ, Blake RC, II, Shah M, Hettich RL, Banfield JF. 2005. Community proteomics of a natural microbial biofilm. Science 308:1915–1920. http://dx.doi.org/10 .1126/science.1109070. 40. Arita-Morioka K, Yamanaka K, Mizunoe Y, Ogura T, Sugimoto S. 2015. Novel strategy for biofilm inhibition by using small molecules targeting molecular chaperone DnaK. Antimicrob Agents Chemother 59:633– 641. http://dx.doi.org/10.1128/AAC.04465-14. 41. Singh VK, Syring M, Singh A, Singhal K, Dalecki A, Johansson T. 2012. An insight into the significance of the DnaK heat shock system in Staphylococcus aureus. Int J Med Microbiol 302:242–252. http://dx.doi.org/10 .1016/j.ijmm.2012.05.001. 42. Lemos JA, Luzardo Y, Burne RA. 2007. Physiologic effects of forced down-regulation of dnaK and groEL expression in Streptococcus mutans. J Bacteriol 189:1582–1588. http://dx.doi.org/10.1128/JB.01655-06. 43. van der Veen S, Abee T. 2010. HrcA and DnaK are important for static and continuous-flow biofilm formation and disinfectant resistance in Listeria monocytogenes. Microbiology 156:3782–3790. http://dx.doi.org/10 .1099/mic.0.043000-0. 44. Brooks JF, II, Gyllborg MC, Cronin DC, Quillin SJ, Mallama CA, Foxall R, Whistler C, Goodman AL, Mandel MJ. 2014. Global discovery of colonization determinants in the squid symbiont Vibrio fischeri. Proc Natl Acad Sci U S A 111:17284 –17289. http://dx.doi.org/10 .1073/pnas.1415957111. 45. Gomes SL, Juliani MH, Maia JC, Silva AM. 1986. Heat shock protein synthesis during development in Caulobacter crescentus. J Bacteriol 168: 923–930.

Journal of Bacteriology

jb.asm.org

2641

Eaton et al.

46. Gomes SL, Gober JW, Shapiro L. 1990. Expression of the Caulobacter heat shock gene dnaK is developmentally controlled during growth at normal temperatures. J Bacteriol 172:3051–3059. 47. Avedissian M, Lessing D, Gober JW, Shapiro L, Gomes SL. 1995. Regulation of the Caulobacter crescentus dnaKJ operon. J Bacteriol 177:3479 –3484. 48. Susin MF, Baldini RL, Gueiros-Filho F, Gomes SL. 2006. GroES/GroEL and DnaK/DnaJ have distinct roles in stress responses and during cell cycle progression in Caulobacter crescentus. J Bacteriol 188:8044 – 8053. http: //dx.doi.org/10.1128/JB.00824-06.

2642

jb.asm.org

49. Ely B. 1991. Genetics of Caulobacter crescentus. Methods Enzymol 204: 372–384. http://dx.doi.org/10.1016/0076-6879(91)04019-K. 50. Ho SN, Hunt HD, Horton RM, Pullen JK, Pease LR. 1989. Site-directed mutagenesis by overlap extension using the polymerase chain reaction. Gene 77:51–59. http://dx.doi.org/10.1016/0378-1119(89)90358-2. 51. Thanbichler M, Iniesta AA, Shapiro L. 2007. A comprehensive set of plasmids for vanillate- and xylose-inducible gene expression in Caulobacter crescentus. Nucleic Acids Res 35:e137. http://dx.doi.org/10.1093/nar /gkm818.

Journal of Bacteriology

October 2016 Volume 198 Number 19

Proper Control of Caulobacter crescentus Cell Surface Adhesion Requires the General Protein Chaperone DnaK.

Growth in a surface-attached bacterial community, or biofilm, confers a number of advantages. However, as a biofilm matures, high-density growth impos...
2MB Sizes 0 Downloads 8 Views