Abstract
The Salmonella flagellar secretion apparatus is a member of the type III secretion (T3S) family of export systems in bacteria. After completion of the flagellar motor structure, the hook-basal body (HBB), the flagellar T3S system undergoes a switch from early to late substrate secretion, which results in the expression and assembly of the external, filament propeller-like structure. In order to characterize early substrate secretion-signals in the flagellar T3S system, the FlgB, and FlgC components of the flagellar rod, which acts as the drive-shaft within the HBB, were subject to deletion mutagenesis to identify regions of these proteins that were important for secretion. The β-lactamase protein lacking its Sec-dependent secretion signal (Bla) was fused to the C-terminus of FlgB and FlgC and used as a reporter to select for and quantify the secretion of FlgB and FlgC into the periplasm. Secretion of Bla into the periplasm confers resistance to ampicillin. In-frame deletions of amino acids 9 through 18 and amino acids 39 through 58 of FlgB decreased FlgB secretion levels while deleting amino acid 6 through 14 diminished FlgC secretion levels. Further PCR-directed mutagenesis indicated that amino acid F45 of FlgB was critical for secretion. Single amino acid mutagenesis revealed that all amino acid substitutions at F45 of FlgB position impaired rod assembly, which was due to a defect of FlgB secretion. An equivalent F49 position in FlgC was essential for assembly but not for secretion. This study also revealed that a hydrophobic patch in the cleaved C-terminal domain of FlhB is critical for recognition of FlgB at F45.
Author summary
Type III secretion (T3S) is the means by which proteins are secreted from the bacterial cytoplasm to build flagella for motility and injectisome structures that facilitate pathogenesis. T3S is the only secretion system known to date that undergoes a secretion-specificity switch. For the assembly of the bacterial flagellum, the T3S system initially secretes early substrates to build the hook-basal body (HBB), which is the main component that makes up the flagellar motor. Upon HBB completion, the flagellar T3S system becomes specific for late substrates, which make up the long external filament that acts as the propeller of the motility organelle. This work identifies important sites of interaction between an early substrate, FlgB and a target site at the cytoplasmic base of T3S apparatus. A second early substrate, FlgC, lacks the targeting interaction found for FlgB suggesting a mechanism that distinguishes early substrates, and may indicate an order to early substrate secretion to facilitate the order of protein subunit assembly for the flagellum.
Introduction
Type III secretion (T3S) is the mechanism for protein secretion to build the bacterial flagellum and the virulence-associated injectisome structures used by Gram-negative plant and animal pathogens to inject effector proteins into host cells to facilitate pathogenesis [1–3]. T3S has evolved to be highly efficient exporting substrates at rates of thousands of amino acids per second [4,5]. The high rate of secretion necessitates that protein secretion be directly coupled to the proton motive force [6,7].
The bacterial flagellum can be divided into three structural components: the basal body, the hook and the filament (Fig 1) [8]. In Salmonella, the basal body is the flagellar motor that includes rotor, stator and rod structures. The rod acts as a drive-shaft, which transverses the periplasmic space from the cytoplasmic membrane through the cell wall to the outer membrane. The hook extends from the rod at the cell surface to a defined length of ~55 nm in Salmonella, which is optimal for the formation of functional flagellar bundles [9]. The hook is a flexible structure that acts as a universal joint to connect the rigid rod to the long, external filament. The filament extends ~10–20 μm from the hook-tip and depending on its rotation, clockwise or counterclockwise, will propel the bacterium in a forward direction or cause the bacterium to tumble as a mechanism to reorient the bacterium in a chemical gradient. The basal body also includes four ring structures. Flagellum assembly initiates with the formation of the MS-ring, consisting of ~34 subunits of a single protein, FliF, in the cytoplasmic membrane (Fig 1) [10–12]. Beneath the MS-ring is the cytoplasmic C-ring, which consists of FliG, FliM and FliN and acts as a bi-directional rotor [13]. The basal body also contains a bushing composed of P(FlgI)- and L(FlgH)-ring structures in the periplasm and lipopolysaccharide (outer membrane), respectively [14,15].
Fig 1. Flagellar assembly steps and assays to measure secretion/assembly of early rod substrates.
A. Flagellum assembly initiates with MS-ring assembly in the inner membrane (IM). The C-ring rotor assemble below the MS-ring. Within the MS-ring the core secretion apparatus forms, which includes FliP5-FliQ4-FliR-FlhB where FliP5 is the actual internal pore through which subunits are secreted. Beneath the FliP5Q4RFlhB core is a large nonameric ring structure formed from 9 identical subunits of FlhA. The N-terminal 330 residues of FlhA includes 8 predicted transmembrane segments followed by a 33 amino acid linker region followed by a large 328 residue C-terminal domain that folds into 4 distinct subdomains, which will crystallize into a nonameric structure for which the structure has been solved. FlhA9 exists in two forms of the C-terminal nonamer, a closed and open form. The closed form allows recognition of early secretion substrates required for the structure and assembly of the rod-hook components of the HBB. The open form allows recognition of late secretion substrates FlgK, FlgL, FliD, FlgM and the filament subunits FliC or FljB. Beneath FlhA is the FliH2I6J ATPase, which facilitates substrate delivery to FlhA9 and hydrolyzes ATP to release and unfold substrates prior to secretion. HBB completion is coupled to a secretion-specificity switch where early rod-hook substrates are no longer recognized and the secretion apparatus is specific for late secretion-substrates. Once the hook-basal body (HBB) structure is completed, FlgM is secreted and σ28 is free to direct transcription from class 3 flagellar promoters to produce hook-filament junction proteins, the filament cap and filament proteins FliC or FljB, which are alternatively expressed. In this way, flagellar class 3 promoter transcription is coupled to HBB completion B. Use of a lac operon reporter to a flagellar class 3 promoter provides a genetic screen for defects in HBB assembly. Any mutant that is defective in HBB assembly does not undergo the secretion-specificity switch. FlgM remains in the cytoplasm bound to σ28 and class 3 promoter transcription is inhibited. The MudJ transposable element is a lac operon reporter, which when fused to a class 3 flagellar promoter, such as a fljB::MudJ fusion, provides an indicator for defects in secretion and/or assembly of FlgB or FlgC proximal rod subunits [21]. Mutants in flgB or flgC that are defective in secretion and/or assembly produce a Lac- phenotype in the presence of the fljB::MudJ reporter. C. Fusion of FlgB or FlgC to β-lactamase, lacking its Sec-dependent secretion signal (Bla), provides a selection/screen for secretion of FlgB or FlgC into the periplasm. Bla must be secreted into the periplasm to fold into an active conformation to confer ampicillin resistance (ApR). Cells expressing wild type FlgB or FlgC fused to Bla are ampicillin resistant (ApR). Cells expressing amino acid substitutions in FlgB or FlgC that are defective in secretion are ampicillin sensitive (ApS).
The majority of the flagellar structure is extracellular. Protein subunits are secreted through a T3S system, which assembles within the MS-ring (Fig 1). The core flagellar T3S apparatus of Salmonella consists of FliP, FliQ and FliR in the stoichiometry FliP5Q4R1 [16]. The FliP, FliQ and FliR proteins have predicted transmembrane segments, yet once assembled, the FliP5Q4R1 complex is located above the predicted inner membrane location. A single copy of FlhB directly associates with the FliP5Q4R1 complex. A nonamer of the FlhA protein associates with the FlhB1- FliP5Q4R1 complex [17]. Each FlhA monomer has eight predicted transmembrane segments in its N-terminus followed by a large C-terminal cytoplasmic domain where subunits to be secreted are delivered [18,19].
A unique feature of T3S systems is the ability to change secretion substrate specificity from one class of protein substrates to a completely different class of substrates [8,20]. In the flagellar T3S system early secretion substrates form the rod and hook structures of the flagellum. Following completion of the hook-basal body (HBB), the flagellar T3S apparatus switches specificity to late substrates. Late structures include the proteins that make up the filament (FliC or FljB) and hook associated proteins or HAPs. There are three HAP proteins FlgK, FlgL and FliD. FlgK and FlgL form a hook-filament junction and FliD is the filament cap, which acts as a scaffold to allow secreted FliC or FljB subunits to fold and assemble on the tip on the growing filament.
The flagellar T3S-specificity switch is part of a mechanism that controls the length of the hook structure. When the hook reaches its terminal length of ~55 nm, hook subunits are no longer secreted. A switch occurs to prevent further hook subunit secretion and allow for filament subunit secretion and polymerization onto the completed hook. The flagellar T3S-specificity switch involves the interaction of an intermittently secreted molecular ruler protein, FliK [22], with a switch protein, FlhB [23,24]. A structure of the N-terminus of FlhB (FlhBN) in complex with FliP5Q4R1 has been solved and demonstrated that a single subunit of FlhBN associates with the FliP5Q4R1 core [25]. Another aspect of the T3S-specificity switch is the requirement that FlhB undergo autocleavage between residues N269 and P270 of the 383 amino acid protein [26]. In strains either deleted for fliK or defective in FlhB autocleavage, cells are unable to undergo the secretion specificity switch. This results in a polyhook phenotype as FlgE (hook subunit) secretion and polymerization continues well beyond the normal 55 nm wild-type hook-length [27,28]. One hypothesis is that once the hook reaches its terminal length, the C-terminus of FliK in the secretion channel interacts with the cleaved C-terminal domain of FlhB (FlhBCCD) resulting in a conformational change in FlhBCCD such that early substrates are no longer recognized by the flagellar T3S apparatus [29]. Recently, photo-crosslinking experiments were used to identify the binding of the FliK C-terminal domain with the FlhBCCD [23,24].
A subject of intense investigation and controversy over several decades regards the nature of the T3S signal. All T3S substrates possess an N-terminal peptide signal that lacks a defined, conserved sequence of amino acids. In the flagellar T3S system, it was shown that the N-terminal amino acids of export substrates are all disordered in their monomeric form in solution [30]. It is also known that late secretion substrates require both an N-terminal peptide signal and a T3S-chaperone, which targets late substrates for secretion [31–33]. Thus, in a current model the T3S-specificity switch transitions from a requirement of an N-terminal peptide signal for early secretion substrates to a requirement for both an N-terminal peptide signal and a T3S-chaperone for late secretion substrates. T3S-chaperones direct their cognate substrates to the large C-terminal cytoplasmic domain of FlhA that resides beneath the core T3S apparatus (Fig 1A) [34–38]. How a conformational change in the FlhBCCD could result in a requirement for T3S-chaperone binding to FlhA to target late substrates remains to be determined.
The purpose of this study was to investigate the nature of early flagellar substrate secretion signals. The early flagellar secretion-substrates are required for the structure and formation of the Rod-Hook component of the HBB. The flagellar rod can be divided into proximal and distal halves. Proximal rod assembly initiates with the secretion and assembly of FliE above the FlhB1-FliP5Q4R1 core followed by FlgB, FlgC and FlgF. The FliE protein plays dual roles in the assembly of the Salmonella flagellum as the final component of the flagellar type III secretion system (fT3SS) and as an adaptor protein that anchors the proximal rod to the membrane imbedded MS-ring structure [39–41]. FlgJ is thought to assemble on the completed proximal rod and is only transiently associated during assembly. FlgJ caps the proximal rod and acts a scaffold for polymerization of the distal rod protein, FlgG. FlgJ is also a muramidase, which digests the cell wall above the proximal rod allowing FlgG polymerization to the outer membrane completing the rod structure [42,43].
Traditional secretion of T3S substrates is evaluated by Western blot analysis of secreted substrates. In our lab, we have developed β-lactamase lacking its Sec secretion signal (Bla) as a reporter for secretion of flagellar proteins into the periplasm through the flagellar type III secretion (T3S) system [40,41,44–47]. Bla must be secreted into the periplasm to fold into an active conformation in order to confer resistance to ampicillin (ApR). The use of Bla as a reporter for flagellar T3S allows for both a positive selection for secretion and for the quantification of secreted protein levels. We chose to characterize the FlgB and FlgC early secretion signals because they are the first true rod subunits secreted and with lengths of 138 and 134 residues, respectively, they are relatively small. Evans et al. reported that purified FlhBCCD binds the hook cap protein FlgD, which is also an early secretion substrate [48]. Small deletions revealed residues within a hydrophobic pocket of the FlhBCCD, which when deleted, no longer bound FlgD. This study revealed that residue F45 in FlgB is essential for secretion of FlgB. We tested for effects of the hydrophobic pocket of the FlhBCCD on recognition of the F45 residue of FlgB for secretion and found genetic evidence for a direct interaction. Recently, Bryant et al. performed a more detailed characterization of the FlgD interaction with the FlhBCCD hydrophobic patch and identified a critical F39 that helps target FlgD for secretion [49]. Those studies together with our findings with FlgB suggests some conservation in substrate recognition for early flagellar secretion substrates by the FlhBCCD.
Results
In-frame deletions in FlgB-Bla and FlgC-Bla implicate regions in FlgB and FlgC important for secretion
The purpose of this study was to identify sequences of early flagellar type III secretion (T3S) substrates that were important for export through the flagellar T3S system. Salmonella bacteria expressing full-length fusions of early secretion substrates FlgB, FlgC, FlgF, FlgG and FlgJ to Bla from their native, chromosomal loci were ampicillin resistant (ApR) in the presence of a functional flagellar T3S apparatus indicating that the fusion proteins are secreted into the periplasm where Bla can fold into an active conformation (S1 Table). In-frame 10 amino acid deletions starting from the C-terminal amino acid codon of flgB or flgC were generated at the chromosomal flg locus (Methods).
For FlgB, deletion of 10 amino acid segments starting from codon 9 through codon 58 all showed reduced FlgB-Bla secretion levels (Fig 2). The most significant reduction in FlgB-Bla secreted levels was observed with deletion of either codons 39 through 48 or codons 49 through 58, which each gave an MIC value for Ap of 12 μg/ml compared to 100 μg/ml for the full length FlgB-Bla construct. Deletion of 10 amino segments after codon 58 had no measurable effect on FlgB-Bla secretion levels suggesting the C-terminal half of FlgB (codons 59 through 138) were not involved in substrate targeting of FlgB to the flagellar T3S apparatus. However, the possibility that these deletions could affect protein expression or stability had not been ruled out.
Fig 2. Effect of in-frame deletions in FlgB and FlgC on the secretion of FlgB-Bla and FlgC-Bla into the periplasm.
Ten amino acid in frame codon deletions were constructed in flgB and flgC starting with the C-terminus. The numbers underneath the double arrow in the colored line indicate which amino acids were deleted in each gene. The MIC values of the full length FlgB-Bla or FlgC-Bla was 100 μg/ml. The MIC values of each deletion is reported using colors (green for 100 μg/ml; yellow for 50 μg/ml; dark grey for 25 μg/ml and lighter grey for 12 μg/ml). A visual representation of the secretion of each deletion is also shown underneath showing the MIC values for each deleted segment fusion above the double arrows.
The effects of 10 amino acid deletions on secreted levels of the Bla fusions were different for FlgC as compared to FlgB. A 10 amino acid deletion of codons 5 through 14 resulted in an MIC for Ap of 25 μg/ml compared to 100 μg/ml for the full length FlgC-Bla construct. Three other 10 amino acid deletions, codons 45–54, 75–84 and 85–94 resulted in MIC levels of 50 μg/ml as did a shorter N-terminal deletion of codons 2–4 (Fig 2). These results suggest that amino acids in the very N-terminus of FlgC (residues 5 through 14) are important, for secretion, but the downstream residues were not as significant for secretion as was observed with FlgB.
Identification of residues in FlgB affecting secretion and/or assembly of FlgB between amino acids 39 through 48
The in-frame deletion experiment presented above suggested that the region between amino acids 39 and 58 of FlgB might be critical for secretion of FlgB-Bla. This region was screened for single amino acids that were important residues for FlgB secretion and/or assembly. Because FlgB assembly is required for completion of the hook-basal body (HBB) and the subsequent secretion-specificity switch from early to late substrate secretion, we utilized a Mud-lac operon reporter fusion to the σ28-dependent class 3 fljB promoter (fljB::MudJ) [21]. MudJ is a transposable element used to fuse the lac operon to any promoter of interest for the characterization of the regulation of transcription of that gene and to provide classical Lac selections and screens to characterize transcriptional regulation for any gene of interest [50]. A mutant in flgB that is defective in HBB assembly would result in accumulation of the late secretion substrate FlgM in the cytoplasm. FlgM is an anti-σ28 factor that is secreted after HBB completion to couple filament gene transcription to completion of the HBB (Fig 1) [51]. A flgB or flgC mutant that is defective in HBB assembly retains FlgM in the cytoplasm, which inhibits σ28-dependent transcription of the fljB::MudJ (fljB-lac) reporter and results in a a Lac- phenotype (Fig 1B) [51].
Amino acid codons 39 through 58 of flgB were mutagenized using a doped oligonucleotide (see Methods). Mutants with various levels of Lac phenotypes were purified, re-tested for their Lac phenotypes and sequenced. S2 Table lists the mutants for which single base pair mutations were obtained. Sanger sequencing of 44 individual mutants between position 39 though 48 resulted in sequences with a variety of single mutations at all amino acid positions in the doped segment, indicating the mutagenesis worked well. Of the 44 single mutants isolated between amino acid codons 39 through 48 of flgB, 18 mutants resulted in reduced fljB::MudJ expression indicating a defect in HBB assembly. Of these 18 mutants, 8 different substitutions were obtained at amino acids 41, 42, 43 and 45 and are reported in Table 1. Of note, twelve of the 18 mutants were a substitution of amino acid F45. All the flgB alleles reported in Table 1 exhibited reduced motility and a defect in HBB assembly as evident by a reduction in fljB::MudJ transcription.
Table 1. Flagellar class 3 expression phenotypes, motility and secretion of flgB alleles assembly defective.
| FlgB amino acid codon | Substitution Mutation | Occurrence | Lac phenotypea | Motility phenotypeb | Ampicillin resistancec |
|---|---|---|---|---|---|
| 41 | R41D | x1 | +/- | 10% | 75 |
| 42 | D42Y | x2 | +/- | 10% | 75 |
| 43 | I43S | x3 | +/- | 5% | 100 |
| 45 | F45C | x1 | +/- | 8% | 5 |
| 45 | F45I | x2 | +/- | 20% | 5 |
| 45 | F45L | x6 | +/- | 17% | 5 |
| 45 | F45M | x1 | +/- | 20% | 5 |
| 45 | F45V | x2 | +/- | 34% | 5 |
| control | WT | ++++ | 100% | 75 |
aAll strains carried fljB5001::MudJ Δhin-5718::FRT alleles to assay for σ28-dependent class 3 flagellar gene transcription that was determined on Mac-Lac and TTC-lac indicator medium (37°C) (++++: TTC-Lac white and ML dark red; +++: TTC-Lac pink and ML dark red; ++: TTC-Lac red and Mac-Lac red; +: TTC-Lac dark red and ML pink; +/-: TTC-Lac dark red and Mac-Lac light pink;—: TTC-Lac dark red Mac-Lac white)
bMotility phenotypes are given as percentage of WT motility at 37°C
cflgB alleles were moved into a flgB-bla fusion expressed at the chromosomal flg locus. Ampicillin resistance levels were assayed on PPBS-Ap plates with varying ampicillin concentrations.
In order to separate amino acid substitutions in FlgB that were defective in secretion from those defective in assembly, the substitutions were introduced into a flgB-bla reporter fusion expressed at the chromosomal flgB locus. An ApR phenotype is indicative that the FlgB-Bla fusion is expressed and secreted into the periplasm (Fig 1C). Thus, we are able to screen flgB alleles that are assembly-defective but secretion-competent from those that are secretion-defective. An amino acid substitution in FlgB that was specific to FlgB assembly and not secretion was expected to have a high level of ApR similar to the wild-type FlgB-bla construct while a substitution defective in secretion was expected to exhibit reduced ApR levels. Of the 8 flgB alleles reported in Table 1, all the substitutions for amino acid F45 showed a severe secretion defect implicating residue F45 of FlgB as critical for secretion while substitutions R41D, D42Y and I43S exhibited ApR levels close to the wild-type (Table 1). We conclude that the R41D, D42Y and I43S are defective in assembly.
The doped oligonucleotide mutagenesis between amino acids 49 through 58 primarily yielded mutants containing multiple mutations that were discarded, and only a few single mutations at some of the positions in the doped region were obtained (S2 Table). These mutants were not followed further as we decided to focus on residue F45 of FlgB for the rest of this study.
Residue F45 is important for FlgB secretion
Since the 5 amino acid substitutions at F45 (F45C, F45I, F45L, F45M, and F45V) in FlgB found through the doped oligo mutagenesis experiment described above were defective in secretion (Table 1), codon 45 in flgB was randomized using an oligonucleotide with all possible DNA sequence combinations (NNN) at codon 45. Alleles coding for all twenty amino acids of flgB codon 45 were obtained and tested for effects on HBB assembly as described above (Fig 1B). All the amino acid substitutions, but the wild type (F45) and Y45, affected normal HBB assembly (Table 2).
Table 2. Flagellar class 3 expression phenotypes and motility of flgB and flgC alleles expressed from their native locus and mutagenized at F45 or F49, respectively.
| FlgB (aa45) | FlgC (aa49) | |||
|---|---|---|---|---|
| Amino Acid | aLac | bMotility | aLac | bMotility |
| ALA | - | 5% | ++ | 75% |
| VAL | + | 20–40% | ++++ | 101% |
| ILE | ++ | 42% | ++++ | 94% |
| LEU | ++ | 40% | ++++ | 100% |
| MET | + | 31% | +++ | 88% |
| PHE | ++++ | 100% | ++++ | 100% |
| TYR | ++++ | 100% | ++++ | 97% |
| TRP | +++ | 60% | +++ | 83% |
| HIS | + | 14% | ++ | 40% |
| LYS | - | 0% | - | 0% |
| ARG | - | 0% | - | 0% |
| ASP | - | 0% | - | 0% |
| GLU | - | 5% | +++ | 88% |
| ASN | - | 0–5% | ++ | 44% |
| GLN | - | 0% | - | 0% |
| SER | - | 5% | ++ | 47% |
| THR | + | 14–19% | ++++ | 110% |
| CYS | + | 20–22% | +++ | 69% |
| GLY | - | 0% | - | 0% |
| PRO | -/+ | 5% | -/+ | 0% |
aStrains carried ΔflgBC fljB5001::MudJ Δhin-5718::FRT alleles to assay for σ28-dependent class 3 flagellar gene transcription which was determined on Mac-Lac and TTC-Lac indicator medium (37°C) (++++: TTC-Lac white and ML dark red; +++: TTC-Lac pink and ML dark red; ++: TTC-Lac red and Mac-Lac red; +: TTC-Lac dark red and ML pink; +/-: TTC-Lac dark red and Mac-Lac light pink;—: TTC-Lac dark red Mac-Lac white)
bMotility phenotypes are given as percentage of WT motility at 37°C.
The flgB alleles with codon 45 substitutions were also placed in front of β-lactamase, this time at the araBAD locus under control of the arabinose-inducible araBAD promoter (ParaBAD). This allowed us to test for the ability of different FlgB-Bla codon 45 substitutions to compete with the wild type FlgB for secretion and inhibition of HBB assembly. We have previously shown that the presence of bile salts increases sensitivity to Ap on solid media, which allows for the detection of small differences in levels of ApR [44]. On proteose peptone bile salts (PPBS) solid medium (see Methods), we observed significant secretion for three of twenty FlgB-Bla codons: F45, W45 and Y45 (Table 3). However, the Y45 and W45 showed reduced ApR compared to F45. The Y45 and W45 substitutions grew similar to F45 on solid medium containing arabinose (PPBS-Ara) and 5 μg/ml Ap, but exhibited Ap-sensitivity at higher Ap concentrations that did not affect the F45 allele tested up to 15 μg/ml. All other substitutions did not grow at the lowest level of Ap (5μg/ml) tested. These results suggest that full wild-type FlgB secretion occurs only with phenylalanine at codon 45 of FlgB and all other amino acid substitutions are defective. We also noted that expression of wild-type (F45) FlgB-Bla from ParaBAD had an inhibitory effect on normal HBB assembly resulting in the inhibition of σ28-dependent transcription of the fljB::MudJ reporter. No other amino acid substitution showed this inhibitory effect (Table 3). We conclude that FlgB-Bla competes with FlgB for secretion and assembly into the growing basal body. The incorporation of a FlgB-Bla fusion into the assembling basal body appears to inhibit further rod assembly.
Table 3. Secretion phenotypes of FlgB-Bla fusions expressed from ParaBAD with amino acid substitutions at codon 45 of FlgB.
The strains carried the fljB5001::MudJ Δhin-5718::FRT alleles to determine the effect of flgB-bla secretion on HBB assembly. Inhibition of HBB assembly would result in accumulation of FlgM in the cytoplasm and inhibition of σ28-dependent transcription of the fljB::MudJ reporter.
| Strain numbera |
FlgB F45nnna |
PPBS-Ara-Apb | Lacc | Ara-Lac c | Expression d | |||
|---|---|---|---|---|---|---|---|---|
| 5 | 7.5 | 10 | 15 | |||||
| TH27277 | A | - | - | - | - | ++++ | ++++ | +++ |
| TH27278 | V | - | - | - | - | ++++ | ++++ | ND |
| TH27279 | I | - | - | - | - | ++++ | ++++ | +++ |
| TH27280 | L | - | - | - | - | ++++ | ++++ | ND |
| TH27281 | M | - | - | - | - | ++++ | ++++ | ND |
| TH27282 | F (WT) | + | + | + | + | ++++ | - | +++ |
| TH27283 | Y | + | +/- | - | - | ++++ | ++++ | +++ |
| TH27398 | W | + | +/- | - | - | ++++ | ++++ | ND |
| TH27284 | H | - | - | - | - | ++++ | ++++ | +++ |
| TH27399 | K | - | - | - | - | ++++ | ++++ | ND |
| TH27285 | R | - | - | - | - | ++++ | ++++ | +++ |
| TH27286 | D | - | - | - | - | ++++ | ++++ | +++ |
| TH27287 | E | - | - | - | - | ++++ | ++++ | +++ |
| TH27400 | N | - | - | - | - | ++++ | ++++ | ND |
| TH27288 | Q | - | - | - | - | ++++ | ++++ | +++ |
| TH27289 | S | - | - | - | - | ++++ | ++++ | +++ |
| TH27401 | T | - | - | - | - | ++++ | ++++ | ND |
| TH27290 | C | - | - | - | - | ++++ | ++++ | ND |
| TH27291 | G | - | - | - | - | ++++ | ++++ | +++ |
| TH27292 | P | - | - | - | - | ++++ | ++++ | +++ |
| TH27299 | Stop | - | - | - | - | ++++ | ++++ | - |
aAll strains had the following genotype: ΔaraBAD::(flgB-5’UTR)-flgB-bla (a.k.a. ParaBAD-flgB-bla) fljB5001::MudJ Δhin-5718::FRT, where amino acid substitutions at codon 45 of the ParaBAD-flgB-bla constructs are indicated.
bAmpicillin resistance was tested on PPBS-arabinose plates with varying ampicillin concentrations as shown in each column in μg/ml Ap. (+: grew; -: did not grow) None of the constructs grew on plates containing Ap5 without arabinose.
cFlagellar class 3 σ28-dependent transcription of a fljB::MudJ reporter fusion construct (fljB5001::MudJ Δhin-5718::FRT) was determined on Mac-Lac and TTC-Lac indicator medium with and without added arabinose at 37°C. (++++: TTC-Lac white and ML dark red; +++: TTC-Lac pink and ML dark red; ++: TTC-Lac red and Mac-Lac red; +: TTC-Lac dark red and ML pink; +/-: TTC-Lac dark red and Mac-Lac light pink;—: TTC-Lac dark red Mac-Lac white)
dFlgB-Bla expression levels were determined on cell extracts by SDS-PAGE followed by western blot analyses using anti-β-lactamase antibodies. “+++” indicates well expressed; “-” indicates not expressed; ND: Not Determined.
We also performed MIC assays in liquid media lacking bile salts (LB+Ara) in order to quantify the effects of codon 45 substitutions on FlgB-Bla secretion under less stringent ApR selection conditions (Tables 4 and S3). The lack of bile salts results in relaxed selection for ApR such that one can detect differences in secretion for codon 45 substitutions that are not observed on the bile salts-containing solid medium. These assays were also done in strains deleted for the chromosomal flgB and flgC genes to ensure that all flagellar secreting structures were identical and the lack of rod structures would allow for maximal levels of secreted FlgB-Bla. Otherwise, secretion would only occur at each basal structure for the short time period after completion of the T3S apparatus and prior to rod completion. After rod completion, the Bla fusion is secreted into the spent growth medium and would not contribute to ApR. Thus, ΔflgBC strains that are unable to assemble rod structures provides a more controlled assay for the effect of codon 45 substitutions on FlgB-Bla secretion (Fig 1C). This also allowed for the quantification of FlgB-Bla secretion independent of potential effects of the different FlgB-Bla codon 45 alleles on rod formation. The effects of codon 45 substitutions on FlgB-Bla secretion were determined by measuring MIC levels to Ap for cells grown in the presence of arabinose inducer. The results are presented in Table 4. The MIC to Ap for FlgB-Bla expressed from ParaBAD in a strain deleted for flgBC, was 25 μg/ml compared to 100 μg/ml for FlgB-Bla expressed from the chromosomal flgB locus (S1 Table). Only a conservative F45Y substitution resulted in a wild-type motility phenotype (Table 2) and an ApR MIC level similar to the wild-type F45 allele (Table 4). Polar amino acid residues exhibited intermediate levels of motility (Table 2) and ApR MIC levels (Table 4) while nonpolar, charged residues, alanine and glycine consistently exhibited substantial defects in FlgB secretion.
Table 4. The effect of amino acid substitutions at codon F45 of flgB and F49 of flgC on the secretion of FlgB-Bla and FlgC-Bla, in a proximal rod mutant background (ΔflgBC).
| aMIC values (μg/ml) flgB-bla or flgC-bla expressed from the ParaBAD in a ΔflgBC background | ||||||
|---|---|---|---|---|---|---|
| Amino Acid | FlgB F45 | FlgC F49 | ||||
| Assay1 | Assay2 | Assay3 | Assay1 | Assay2 | Assay3 | |
| ALA | 12 | 12 | 12 | 3 | 6 | 3 |
| VAL | 3 | 6 | 6 | 12 | 12 | 12 |
| ILE | 12 | 12 | 12 | 6 | 12 | 12 |
| LEU | 6 | 6 | 6 | 12 | 12 | 12 |
| MET | 3 | 6 | 6 | 6 | 6 | 6 |
| PHE | 25 | 25 | 25 | 12 | 12 | 25 |
| TYR | 25 | 25 | 25 | 12 | 12 | 12 |
| TRP | 12 | 12 | 12 | 12 | 12 | 12 |
| HIS | 12 | 12 | 12 | 3 | 3 | 6 |
| LYS | 6 | 6 | 6 | 12 | 12 | 6 |
| ARG | 6 | 6 | 6 | 12 | 12 | 12 |
| ASP | 3 | 6 | 6 | 12 | 12 | 12 |
| GLU | 6 | 6 | 6 | 6 | 6 | 12 |
| ASN | 6 | 12 | 12 | 6 | 12 | 12 |
| GLN | 6 | 6 | 6 | 25 | 25 | 25 |
| SER | 12 | 12 | 12 | 12 | 12 | 12 |
| THR | 6 | 6 | 6 | 12 | 12 | 12 |
| CYS | 6 | 6 | 6 | 12 | 12 | 12 |
| GLY | 12 | 12 | 12 | 12 | 12 | 12 |
| PRO | 12 | 12 | 25 | 12 | 12 | 25 |
aThe MIC to Ap were measured in strains expressing flgB-bla with amino acid substitutions at codon 45 of flgB or flgC-bla with amino acid substitutions at codon 49 of flgC from the araBAD locus (ParaBAD-flgB-bla or ParaBAD-flgC-bla) in strains deleted for the proximal rod components (ΔflgBC background). MIC assays were conducted on strains grown in the presence of arabinose. No addition of arabinose resulted in MIC’s of less than 1.5 μg/ml
The ApR assays on solid medium containing bile salts provided a stringent screen that allowed us to determine that the F45Y and F45W substitutions were defective relative to the wild type F45 codon. The MIC assays in the absence of bile salts provided less stringent ApR selection, which allowed us to observe a gradient of effects of codon 45 substitutions on secretion: in general aromatic amino acids (F,Y,W) were secreted to higher levels than the remaining hydrophobic residues (V,I,L,M,C), which were secreted to higher levels than polar residues (H,S,T,N,Q), which, except for N and Q, were secreted to higher levels than small residues, proline and charged residues (A,G,P,L,R,D,E).
Residue F49 is important for FlgC assembly
Evans et al. (2013) had reported that a conserved phenylalanine residue, present in the N-terminus of all rod and hook substrates, might be critical for secretion of early substrates [48]. For FlgB, this residue would correspond to the phenylalanine at amino acid 45, and indeed, as shown above, the effect of FlgB F45 on FlgB secretion is very clear. For FlgC, F49 was predicted to be critical for secretion [48]. However, a deletion of codons 45 through 54 of flgC resulted in only a slight, 2-fold reduction for MIC to Ap for a FlgC-Bla fusion (Fig 2). Alleles with the twenty amino acids at codon 49 of flgC were obtained and tested for hook basal body (HBB) assembly as described earlier (Fig 1B). As for FlgB, residues K, R, D, Q and G in place of F49 in FlgC prevented FlgC assembly (Table 2). Residues V, I, L, F and Y at amino acid 49 of FlgC assembled as wild type. The rest of the substitutions showed some decrease in assembly, as seen from the Lac and motility phenotypes.
FlgC F49 alleles were then placed in a flgC-bla construct expressed from the ParaBAD promoter in a strain deleted for flgBC, and their effects on FlgC-Bla secretion were determined by measuring MIC levels to Ap for cells grown in the presence of arabinose inducer (Table 4). A difficulty with this set of strains was that the MIC obtained for flgC-bla expressed from ParaBAD in a strain deleted for flgBC, was relatively low, 12 μg/ml, giving little room for detecting significant effects of amino acid substitutions on secretion. FlgC substitutions F49A, F49M, F49H and F49E resulted in reduced MIC levels for Ap at ~3, 6, ~3 and ~6 μg/ml, respectively. The F49 substitutions that exhibited a non-motile phenotype (0% motility in Table 2) were all secreted at or near wild-type levels when tested in the FlgC-Bla assay. This indicated that they were defective in assembly rather than secretion. One FlgC substitution, F49Q, resulted in higher MIC for secretion of FlgC-Bla (Table 4), but the same substitution at the chromosomal flgC locus resulted in a non-motile phenotype indicating a strong defect in assembly (Table 2). F49D was also well secreted (MIC = 12μg/ml), but nonmotile suggesting a defect in assembly. We note that F49D was assembly-defective while F49E showed almost wild-type motility (88%) despite the similarity between these amino acid residues.
The set of F49 substitutions in the ParaBAD-flgC-bla constructs were placed into a strain containing the wild type chromosomal rod components and the class 3 fljB::MudJ reporter. Secretion of FlgC-Bla alleles were tested on PPBS solid medium containing arabinose (Table 5). Most results agreed with the MIC values performed in LB media in the proximal rod mutant background (shown in Table 4) and it is clear that FlgC F49 appeared more tolerant to amino acid substitution than FlgB F45 (compare Tables 3 and 5). Alanine and histidine were the only substitutions at amino acid 49 of FlgC-Bla that did not grow on PPBS-Ara with 5ug/ml of Ap. This suggests a defect in secretion, yet when substituted in the chromosomal flgC locus (no fusion) they exhibited 75% of wild-type motility for F49A and 40% for F49H (Table 2). The substitutions did not affect expression levels (S4 Fig). We hypothesize that even though they are defective in secretion, there is enough low-level secretion to assemble the 6 FlgC subunits needed to form a functional rod, but not enough is secreted in the ParaBAD-flgC-bla constructs to confer ApR on the stringent PPBS-ApR plates.
Table 5. Secretion phenotypes of FlgC-Bla fusions expressed from ParaBAD with amino acid substitutions at codon 49 of flgC.
The strains carried the fljB5001::MudJ Δhin-5718::FRT alleles to determine the effect of FlgC-Bla secretion on HBB assembly. Inhibition of HBB assembly would result in accumulation of FlgM in the cytoplasm and inhibition of σ28-dependent transcription of the fljB::MudJ reporter.
| Strain numbera | FlgC F49nnna | PPBS-Ara-Apb | Lacc | Ara-Lac c | Expression d | |||
|---|---|---|---|---|---|---|---|---|
| 5 | 7.5 | 10 | 15 | |||||
| TH27591 | A | - | - | - | - | ++++ | ++++ | +++ |
| TH27592 | V | + | + | + | + | ++++ | - | +++ |
| TH27593 | I | + | + | + | + | ++++ | - | +++ |
| TH27594 | L | + | + | + | + | ++++ | - | +++ |
| TH27595 | M | + | + | + | + | ++++ | - | +++ |
| TH27546 | F (WT) | + | + | + | + | ++++ | - | +++ |
| TH27596 | Y | + | + | + | + | ++++ | + | +++ |
| TH27597 | W | + | + | + | + | ++++ | +++ | +++ |
| TH27598 | H | - | - | - | - | ++++ | ++++ | +++ |
| TH27599 | K | + | + | + | + | ++++ | ++++ | ND |
| TH27600 | R | + | + | + | + | ++++ | ++++ | ND |
| TH27601 | D | + | + | + | + | ++++ | ++++ | ND |
| TH27602 | E | + | + | + | + | ++++ | ++++ | ND |
| TH27603 | N | + | + | + | + | ++++ | ++++ | ND |
| TH27604 | Q | + | + | + | + | ++++ | ++++ | ND |
| TH27605 | S | + | + | + | + | ++++ | ++++ | ND |
| TH27606 | T | + | + | + | + | ++++ | +++ | ND |
| TH27607 | C | + | + | + | + | ++++ | + | ND |
| TH27608 | G | + | + | + | + | ++++ | +++ | ND |
| TH27609 | P | + | + | + | + | ++++ | ++++ | ND |
aAll strains had the following genotype: ΔaraBAD::flgC-bla fljB5001::MudJ Δhin-5718::FRT, where amino acid substitutions at codon 49 of the ParaBAD-flgC-bla constructs are indicated.
bApR was tested on PPBS-arabinose plates with varying Ap concentrations as shown in μg/ml Ap (+: grew; -: did not grow). None of the constructs grew on plates containing Ap5 without arabinose.
cFlagellar class 3 σ28-dependent transcription of a fljB::MudJ reporter fusion construct (fljB5001::MudJ Δhin-5718::FRT) was determined on Mac-Lac and TTC-Lac indicator medium with and without added arabinose at 37°C. (++++: TTC-Lac white and ML dark red; +++: TTC-Lac pink and ML dark red; ++: TTC-Lac red and Mac-Lac red; +: TTC-Lac dark red and ML pink; +/-: TTC-Lac dark red and Mac-Lac light pink;—: TTC-Lac dark red and Mac-Lac white).
dFlgC-Bla levels were determined on cell extracts by SDS-PAGE followed by western blot analyses using anti-β-lactamase antibodies. “+++” indicates well expressed; “-” indicates not expressed; ND: Not Determined.
Just as for FlgB-Bla, we noted that expression of FlgC-Bla from ParaBAD had an inhibitory effect on normal HBB assembly resulting in the inhibition of σ28-dependent transcription of the fljB::MudJ reporter. Substitutions of amino acid 49 of FlgC with V, I, L and M also showed this inhibitory effect (Table 5). C, Y, W, T, and G had also some level of inhibitory effects, but none of the other substitutions exhibited this effect when secreted. Residues such as K, R, D and Q seemed to allow FlgC secretion but not assembly (see Tables 2 and 5). It appears that only some alleles of FlgC-Bla can assemble into the growing basal body to inhibit further rod assembly. We conclude that the F49 position in FlgC is not as critical for FlgC secretion as is F45 of FlgB, but there is selectivity for residues required for assembly.
The C-terminus of FlhB is a target for FlgB-Bla secretion
The flagellar type 3 secretion (T3S) system undergoes a secretion-specificity switch from early rod-hook secretion substrates to late secretion substrates, which include filament subunits and the anti-σ28 factor FlgM. The secreted molecular ruler protein FliK determines hook-length completion and catalyzes the secretion-specificity switch at the FlhB component of the flagellar T3S system [22,52]. The T3S-specificity switch also requires that FlhB undergo a spontaneous autocleavage event between residues N269 and P270 of the 383 amino acid protein [26]. The cleaved N-terminal domain of FlhB associates with FliP5Q4R export gate while the location of the cleaved C-terminal domain of FlhB (FlhBCCD) within the T3S apparatus is not known [25]. A null mutant in fliK or a mutation in flhB that is unable to undergo autocleavage are unable to switch, which results in a polyhook phenotype [26–28]. The FlhBCCD structure includes a hydrophobic pocket of four residues, A286, P287, A341 and L344, [48]. Substitution of individual residues to glutamate prevented secretion of either distal rod protein FlgG or the FlgE hook protein subunits. This hydrophobic pocket was proposed to be a docking site for early secretion substrates through interaction with a critical phenylalanine residue present in the N-terminus of early secretion substrates. Deletion of F36 through L40 of FlgD prevented secretion of FlgD subunits [48]. More recently, residues 2 through 5 and 36 through 40 of FlgD, were identified as critical for secretion [49]. This included the phenylalanine residue at codon 36. Our results suggest that F49 of FlgC is not critical for FlgC secretion, but F45 of FlgB is.
In order to test whether the hydrophobic pocket of the FlhBCCD was important in recognition of F45 for secretion of FlgB, codons 286, 341 and 344 of the hydrophobic pocket were targeted for VNN codon mutagenesis simultaneously (see Methods) and screened for mutations able to suppress flgB secretion-defective alleles at codon 45. The proline residue 287 of FlhB was not mutated since it could impact the overall structure of the FlhBCCD. The targeted mutation of codons 286, 341 and 344 of flhB as described would result in 110,592 different codon combinations. The wild-type codon sequence is predicted to occur in one out of ~2,000 mutants. Approximately 150,000 flhB 286VNN 341VNN 344VNN mutants were generated (see Methods).
Preliminary experiments (See Methods) showed that the residues in the surface-exposed hydrophobic pocket in the FlhBCCD were not under a strict selection for FlhB function. It appeared that the ideal FlhB residues for optimal secretion and assembly of hook basal body are A, V or S for residue 286, A, V, S, G or I for residue 341 and L, I or V for residue 344 (Fig 3). A plasmid vector expressing flhB with a L344A substitution complemented an flhB null mutant [38]. Alanine at residue 344 of FlhB was not enriched in our NCE-lac pool (selecting for HBB completion) compared to the TcS pool of total mutants (the 150,000 flhB mutant combination). Our method suggests that residues such as leucine, isoleucine and valine work best for FlhB function, when flhB is expressed from its native chromosomal locus.
Fig 3. Mutagenesis of the FlhB C-terminus surface exposed hydrophobic pocket (FlhB A286VNN A341VNN L344VNN) in a fljB5001::MudJ Δhin-5718::FRT background to determine the effect of FlhB hydrophobic pocket substitutions on HBB assembly.
Approximately 1,000 mutants from the TcS pool and from the Lac+ pool were analyzed using next generation sequencing. The diagram represents the log odds score calculated as (log2(Lac+ count /TcS count)) for each amino acid at each position. The Lac+ (NCE-lac pool) mutants represent the mutants forming a functional HBB structure (Panel A-top and enlarged in Panel B). Panel C shows the positions of the residues (A286, A341 and L344) mutated in the FlhB C-terminus surface exposed hydrophobic pocket (PDB 3B0Z; [53]).
Our goal was to use the pool of targeted FlhBCCD substitutions (150,000 flhB 286VNN 341VNN 344VNN mutants pool) to screen for suppressors that permitted the secretion of FlgB-Bla fusions with substitutions for amino acid F45. These F45 FlgB-Bla mutants are defective in secretion by a wild-type FlhB-containing flagellar T3S system. If the hydrophobic pocket in the FlhBCCD was a target for recognition of FlgB through the F45 residue, it might be possible to alter the hydrophobic pocket to recognize amino acid substitutions at position 45 and allow for mutant FlgB-Bla secretion. Twenty constructs, each with a different amino acid substitution at codon 45 of flgB-bla, were expressed at the araBAD (ParaBAD-flgB-bla) locus leaving the chromosomal flgB gene intact and harboring the fljB::MudJ reporter to assay defects in HBB assembly. The parent strain, ParaBAD-flgB-bla fljB::MudJ is Lac+ (ie. HBB+) in the absence of arabinose inducer (Fig 4A) and Lac- (HBB-) when FlgB-Bla is expressed by addition of arabinose (Fig 4B). The parent strain also grows on 30 μg/ml Ap plates with added arabinose (see TH27282 in Table 6 and Fig 4B).
Fig 4. Illustration showing the secretion of FlgB and FlgB-Bla alleles in different genetic backgrounds.
Panel A represents normal secretion/assembly of FlgB. When FlgB-Bla is co-expressed, we propose that FlgB-Bla assembles into the structure and interferes with further rod assembly that prevents FlgM secretion and results in a Lac- phenotype (Panel B). A secretion defective FlgB*-Bla allele is not secreted and therefore does not interfere with rod assembly (Panel C). Panels D, E and F include mutants in flhB (FlhB*) that allow FlgB*-Bla alleles to be secreted, but constitute two classes: FlgB*-Bla alleles that are fully secreted (Panel D and F) or those stuck in the secretion apparatus (Panel E). Alleles of FlhB were also obtained that secrete mutant, FlgB*-Bla, but lose the ability to secrete wild type FlgB (Panel F).
Table 6. Characterization of alleles in flhB that allowed secretion of FlgB-Bla fusions expressed from ParaBAD with amino acid substitutions at codon 45 of flgB.
The strains also carried the fljB5001::MudJ Δhin-5718::FRT alleles to determine the effect of FlgB-Bla secretion on HBB assembly. Inhibition of HBB assembly would result in accumulation of FlgM in the cytoplasm and inhibition of σ28-dependent transcription of the fljB::MudJ reporter.
| Straina | FlgBb | Lacc | PPBS-Ara-Apd | FlhB sequencee | ||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| no ara | ara | 5 | 7.5 | 10 | 15 | 30 | 50 | A286 (gcg) |
A341 (gcg) |
L344 (ttg) |
||
| TH27282 | F45 | ++++ | - | + | + | + | + | + | - | wild type flhB | ||
| TH27283 | F45Y | ++++ | ++++ | + | +/- | - | - | - | - | wild type flhB | ||
| TH27519 | F45Y | +++ | - | + | + | + | + | + | - | A (gct) | A (gca) | Q(caa) |
| TH27520 | F45Y | ++ | - | + | + | + | + | - | - | V (gtc) | A (gcc) | Q(caa) |
| TH27521 | F45Y | +++ | - | + | + | + | + | - | - | V (gtc) | A (gct) | H (cat) |
| TH27522 | F45Y | ++++ | ++ | + | + | + | + | + | - | V (gtc) | A (gca) | V (gta) |
| TH27523 | F45Y | - | - | + | + | + | + | - | - | P (ccg) | A (gct) | H (cat) |
| TH27398 | F45W | ++++ | ++++ | + | +/- | - | - | - | - | wild type flhB | ||
| TH27524 | F45W | ++++ | ++/- | + | + | + | + | + | - | I (att) | A (gca) | V (gtt) |
| TH27525 | F45W | ++++ | ++ | + | + | + | + | + | - | Y (tac) | A (gca) | V (gta) |
| TH27526 | F45W | ++++ | ++ | + | + | + | + | + | - | V (gta) | A (gca) | V (gta) |
| TH27527 | F45W | ++++ | ++/- | + | + | + | + | + | - | P (cct) | A (gca) | I (ata) |
| TH27280 | F45L | ++++ | ++++ | - | - | - | - | - | - | wild type flhB | ||
| TH27528 | F45L | ++++ | + | - | - | - | - | - | - | P (cca) | V (gtg) | L (ctc) |
| TH27529 | F45L | +++ | - | - | - | - | - | - | - | V (gtg) | T (acg) | L (cta) |
| TH27530 | F45L | ++ | - | - | - | - | - | - | - | T (aca) | T (acc) | M (atg) |
| TH27277 | F45A | ++++ | ++++ | - | - | - | - | - | - | wild type flhB | ||
| TH27531 | F45A | ++ | - | + | +/- | - | - | - | - | A (gcg) | R (agg) | L (cta) |
| TH27532 | F45A | + | - | + | +/- | - | - | - | - | Q (cag) | I (atc) | L (ctt) |
| TH27289 | F45S | ++++ | ++++ | - | - | - | - | - | - | wild type flhB | ||
| TH27533 | F45S | + | - | + | - | - | - | - | - | A (gcc) | G (ggg) | A (gca) |
| TH27534 | F45S | +++ | - | +/- | - | - | - | - | - | T (acg) | E (gag) | L (cta) |
| TH27277 | F45T | ++++ | ++++ | - | - | - | - | - | - | wild type flhB | ||
| TH27535 | F45T | +++ | - | + | +/- | - | - | - | - | N (aac) | A (gcg) | M (atg) |
| TH27290 | F45C | ++++ | ++++ | - | - | - | - | - | - | wild type flhB | ||
| TH27536 | F45C | ++ | - | + | +/- | - | - | - | - | G (ggc) | T (aca) | I (ata) |
| TH27537 | F45C | ++ | - | + | +/- | - | - | - | - | A (gct) | R (agg) | L (ctt) |
| TH27285 | F45R | ++++ | ++++ | - | - | - | - | - | - | wild type flhB | ||
| TH27538 | F45R | ++ | - | + | +/- | - | - | - | - | A (gca) | V(gtc) | E (gag) |
aAll strains had the following genotype: ΔaraBAD::(flgB-5’UTR)-flgB-bla (ParaBAD-flgB-bla) fljB5001::MudJ Δhin-5718::FRT and FlhB A286NNN A341VNN L344VNN, where amino acid substitutions at codon 45 of the ParaBAD-flgB-bla constructs are indicated and the flhB mutation reported. The detailed genotype is listed on S6 Table.
bIndication of the amino acid substitution at position 45 of FlgB expressed from ParaBAD-flgB-bla
cFlagellar class 3 σ28-dependent transcription of a fljB::MudJ reporter fusion construct (fljB5001::MudJ Δhin-5718::FRT) was determined on Mac-Lac and TTC-lac indicator medium with and without added arabinose (++++: TTC-Lac white and ML dark red; +++: TTC-Lac pink and ML dark red; ++: TTC-Lac red and Mac-Lac red; +: TTC-Lac dark red and ML pink; —: Mac-Lac white)
dAmpicillin resistance was tested on PPBS- plates with and without arabinose (PPBS-Ap5) and with varying ampicillin concentrations as shown in μg/ml Ap. None of the strains grew on PPBS-Ap5 without arabinose. (+: grew well; +/-: grew a little; -: did not grow)
eflhB mutations, as obtained by DNA sequence analysis
The twenty strains, containing a different amino acid substitution at codon 45 of flgB-bla, carried a tetRA cassette insertion replacing codons 286 through 344 of flhB. The ΔflhB::tetRA cassette was replaced by P22-mediated transduction with phage grown on the pool of targeted FlhBCCD substitutions followed by selection on TcS selective media. A dilution of P22 transducing lysate was used that gave ~5,000 TcS transductants per plate. The ~5,000 TcS transductants obtained were then pooled, grown in LB with arabinose and plated onto PPBS-Ap5-Ara plates to screen for flhB mutants that would secrete the different FlgB-Bla constructs. For seventeen constructs the number of Ara-ApR recombinants was low (S5 Table). Substitutions F45V, F45I, F45M, F45K, F45Q, and F45P as recipients yielded no Ara-ApR recombinants. Eleven recipients yielded few Ara-ApR recombinants: F45A(5), F45L(1), F45H(3), F45R(16), F45D(4), F45E(2), F45N(1), F45S(7), F45T(1), F45C(2) and F45G(1). The wild-type F45(>2,000) and aromatically related substitutions F45Y(>500) and F45W(>200) yield significantly higher numbers of Ara-ApR recombinants than the other seventeen recipients.
Twenty Ara-ApR recombinants were marker-rescued and tested on Lac indicator media in the presence and absence of Ara to determine if the mutated flhB gene could produce a functional flagellar T3S apparatus (Lac+ without added Ara, Fig 4A), or if expression of the FlgB-Bla fusion could prevent the assembly of a functional flagellar T3SS (Lac- with added Ara inducer, Fig 4B). The results are shown in Table 6. All flhB alleles that showed wild type levels of fljB::MudJ expression had either a leucine or valine codon at position 344 of flhB. This is consistent with the preliminary experiments showing that leucine, isoleucine and valine at amino acid 344 of FlhB worked best for FlhB function. All the mutants tested showed a defect in flagellar T3S assembly in the presence of Ara. The levels of Ara-ApR was low for all mutants tested with the exception of the mutants obtained for FlgB F45Y and F45W. We were initially surprised to find that following marker rescue, many of the mutants were not able to grow on Ara-5μg/ml Ap plates, which they were originally selected on. However, this should not have been unexpected given the low concentration of Ap used in the selective media and the high number of colonies (~5,000) plated from the TcS selection plates onto Ara-5μg/ml Ap screening plates.
Most flhB mutants presented in Table 6 allowed for the secretion of the FlgB-Bla allele expressed from ParaBAD (Ara-ApR), but with reduced secretion of the wild type FlgB, as measured by the Lac phenotypes without arabinose. Ara-ApR mutants obtained with FlgB-Bla F45Y and F45W substitutions exhibited similar phenotypes to the wild type parent strain (TH27282), while mutants obtained with FlgB-Bla F45A, F45S, F45T and F45R allowed FlgB-Bla to be secreted, but resulted in the inability to construct a functional HBB (decreased fljB::MudJ expression without arabinose). The parent strains (containing flhB wild type) for FlgB-Bla F45A (TH27277), F45S (TH27289), F45T (TH27277) and F45R (TH27285) did not allow any secretion of the FlgB-Bla allele and were completely Ara-Lac+ (Table 6 and Fig 4C).
Mutants obtained with FlgB-Bla F45L exhibited impaired HBB assembly with added arabinose (Ara-Lac-), but remained Ara-ApS. One possibility is that FlgB-Bla F45L is stuck in the secretion apparatus blocking both HBB assembly and FlgB-Bla secretion (Fig 4E). This could be due to a stronger interaction of the FlgB-Bla F45L to the mutant FlhB motif and supports the hypothesis of Bryant et al that early substrates are recognized through interaction with the FlhBCCD hydrophobic pocket [49].
We found one allele in flhB (A286P A341A L344H) that impaired FlgB secretion (Lac- without arabinose; Fig 4F), but yet permitted FlgB-Bla F45Y to be secreted to higher levels of ApR than in the flhB parent strain. This mutant (TH27523) demonstrates that it is possible to alter the FlhB hydrophobic pocket and allow a different secretion-substrate to be selectively exported. We were surprised that this mutant prevented the secretion of FlgB F45 and allowed F45Y to be secreted, given the similarity between phenylalanine and tyrosine residues. In order to ensure that flhB (A286P A341A L344H) was impaired for wild type FlgB secretion, this flhB mutation was moved into the FlgB-Bla wild type parent strain (TH27382), and tested for Ap. The wild type FlgB-Bla was not secreted upon addition of arabinose (Ara-ApS), demonstrating that the FlhB (A286P A341A L344H) mutant selectively impaired FlgB wild type secretion (Table 7, see TH27282 and TH27545 phenotypes).
Table 7. Phenotypes of selected mutants in flhB in combination with specific flgB-bla fusions expressed from ParaBAD with amino acid substitutions at codon 45 of flgB.
| Straina | FlgBb | Lacc | PPBS-Ara-Apd | FlhB sequencee | ||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| no ara | ara | 5 | 7.5 | 10 | 15 | 30 | 50 | A286 (gcg) |
A341 (gcg) |
L344 (ttg) |
||
| TH27283 | F45Y | ++++ | ++++ | + | +/- | - | - | - | - | wild type flhB | ||
| TH27523 | F45Y | - | - | + | + | + | - | - | - | P (ccg) | A (gct) | H (cat) |
| TH27282 | F45 | ++++ | - | + | + | + | + | + | - | wild type flhB | ||
| TH27545 | F45 | - | - | - | - | - | - | - | - | P (ccg) | A (gct) | H (cat) |
| TH27285 | F45R | ++++ | ++++ | - | - | - | - | - | - | wild type flhB | ||
| TH27538 | F45R | ++ | - | + | +/- | - | - | - | - | A (gca) | V(gtc) | E (gag) |
| TH27549 | F45R | ++ | - | + | - | - | - | - | - | A (gcg) | A (gcg) | E (gag) |
| TH27548 (flhD*C*) |
F45R | ++++ | ++++ | + | + | + | + | - | - | A (gcg) | A (gcg) | E (gag) |
| TH27287 | F45E | ++++ | ++++ | - | - | - | - | - | - | wild type flhB | ||
| TH27547 (flhD*C*) |
F45E | - | - | + | +/- | - | - | - | - | A (gcg) | A (gcg) | R(cgt) |
| TH27562 | F45E | - | - | - | - | - | - | - | - | A (gcg) | A (gcg) | R(cgt) |
aAll strains had the following genotype: ΔaraBAD::(flgB-5’UTR)-flgB-bla (ParaBAD-flgB-bla) fljB5001::MudJ Δhin-5718::FRT and FlhB A286NNN A341VNN L344VNN, where amino acid substitutions at codon 45 of the ParaBAD-flgB-bla constructs are indicated and the flhB mutation reported. The detailed genotype is listed on S6 Table.
bIndication of the amino acid substitution at position 45 of FlgB expressed from ParaBAD-flgB-bla
cFlagellar class 3 σ28-dependent transcription of a fljB::MudJ reporter fusion construct (fljB5001::MudJ Δhin-5718::FRT) was determined on Mac-Lac and TTC-lac indicator medium with and without added arabinose (++++: TTC-Lac white and ML dark red; +++: TTC-Lac pink and ML dark red; ++: TTC-Lac red and Mac-Lac red; +: TTC-Lac dark red and ML pink; —: Mac-Lac white)
dAmpicillin resistance was tested on PPBS- plates with and without arabinose (PPBS-Ap5) and with varying ampicillin concentrations as shown in μg/ml Ap. None of the strains grew on PPBS-Ap5 without arabinose. (+: grew well; +/-: grew a little; -: did not grow)
eflhB mutations, as obtained by DNA sequence analysis
An interesting mutant in flhB was FlhB (A286A A341V L344E) that permitted FlgB-Bla F45R to be secreted more than in the wild-type FlhB (A286 A341 L344). This mutant affected wild type FlgB secretion, as seen with the decrease of Lac activity without arabinose. The single L344E substitution in FlhB was constructed by λ-Red recombineering and was found to still permit FlgB-Bla F45R secretion (TH27549; Table 7). The introduction of protease-resistant alleles of the flagellar master regulatory genes flhD flhC (flhD*C*) that are resistant to ClpXP proteolysis and increase overall flagellar gene expression resulted in increased FlgB-Bla F45R secretion levels, as expected (see TH27448; Table 7). The FlhD and FlhC proteins combine to form the FlhD4C2 transcriptional activator complex of flagella HBB class2 gene transcription [54]. FlhD and FlhC are regulated at a post-transcriptional level by the ClpXP protease through YdiV/RflP(for FlhD) and FliT(for FlhC) adaptor proteins that target FlhD and FlhC to ClpXP [55, 56]. We reasoned that since FlhB L344E allowed some FlgB-Bla F45R secretion, L344R might permit secretion of FlgB-Bla F45E. The construct was made, but the FlhB L344R allele was defective for the secretion of both wild type F45 and mutant F45R alleles, although a low level of secretion was observed in the flhD*C* background (Table 7, strains TH27562 and TH27547).
Of note, identical substitutions in FlhB (A286V A341A L344V) were isolated as able to secrete F45W FlgB-Bla and F45Y FlgB-Bla (see TH27522 and TH27526- Table 6). This was not surprising given the similarity between the tryptophan and tyrosine residues. FlhB (A286A A341R L344L) was also independently isolated as able to secrete FlgB-Bla with either F45A or F45C substitutions (see TH27531 and TH27537- Table 6).
All together, these data strongly support the hypothesis that an interaction of position 45 of FlgB with amino acids in the conserved hydrophobic pocket of the FlhBCCD occurs during secretion as amino acid substitutions in the hydrophobic pocket allow for the secretion of FlgB-Bla substrates carrying position 45 substitutions that are not secreted by a flagellar T3S system with a wild-type FlhB protein.
Discussion
The goal of this research was to characterize possible amino acid secretion signals in the early flagellar T3S substrates FlgB and FlgC. It had been proposed that F45 of FlgB and F49 of FlgC were important for recognition by a hydrophobic pocket in the FlhB component of the flagellar T3S system [48]. Rather that targeting those residues specifically, we decided to identify regions of FlgB and FlgC important for their secretion and determine if a forward genetic approach would independently reveal F45 of FlgB and F49 of FlgC to be important for their secretion. Our system used a lac operon fusion to a σ28-dependent flagellar class 3 promoter (fljB::MudJ) as an indicator for the presence of a functional flagellar T3S apparatus. A defective apparatus would be unable to secrete the anti-σ28 factor FlgM and result in a Lac- phenotype on lactose indicator medium. Our system also employed fusion of β-lactamase, deleted for its Sec-dependent secretion signal (Bla), as a reporter for secretion of flagellar secreted substrates into the periplasm. Resistance to ampicillin (ApR) requires that the Bla fusions be secreted by a functional flagellar T3S system into the periplasmic space where the Bla reporter will fold into an active conformation in order to confer ApR.
Targeted deletions of flgB and flgC revealed three segments of FlgB, amino acids 9 through 18, 39 through 48, and 49 through 58, which when deleted exhibited a significant reduction in FlgB-Bla secretion levels as indicated by reduced resistance to Ap. For FlgC, only a deletion of amino 6 though 14 showed a significant reduction in Ap resistance levels. A doped oligonucleotide mutagenesis for the amino acid 39 through 48 region of FlgB revealed position F45 to be critical for secretion as had been predicted earlier. However, a targeted mutagenesis of F49 of FlgC revealed this position, although critical for assembly, was not as critical for FlgC secretion.
We characterized residues in the FlhBCCD that form a hydrophobic patch predicted to interact with F45 of FlgB for secretion. We determined that these residues were important for FlgB-Bla secretion. L344 is a critical residue although isoleucine and valine at this position also produced wild-type secretion levels. Valine or serine substitutions at A286 also resulted in wild-type secretion while A341 was less restrictive as substitutions to glycine, valine or serine resulted in wild-type FlgB-Bla secretion levels in our assay system. A codon randomization for positions 286, 341 and 344 of the FlhBCCD hydrophobic patch was used to identify amino acid substitutions that would allow for secretion of secretion-defective FlgB-Bla fusions carrying amino acid substitution alleles at position 45. Suppression of secretion-defective alleles was possible by amino acid changes in the FlhBCCD hydrophobic patch but not to wild-type levels of secretion.
In conclusion, we did not find an early flagellar T3S signal that was conserved in amino acid sequence within the N-terminus FlgB and FlgC. However, this work agrees with an earlier study that identified a hydrophobic patch in the FlhBCCD that was shown to be important for secretion of the early flagellar FlgE (hook) and FlgD (hook-cap) secretion substrates [48]. The prediction that F45 of FlgB is an important residue in targeting FlgB for secretion was found to be the case, but not for F49 of FlgC. A recent study by Bryant et al. characterized the early secretion signal for early secretion substrate FlgD [49]. They identified residue F36 of FlgD as important for secretion. These authors also identified residues 2 through 5 of FlgD as critical for FlgD secretion. These results are similar to what we observed for FlgB suggesting a common early secretion signal between the FlgB and FlgD substrates. The work presented in this study identified two regions, amino acids 9 through 18 of FlgB and 6 through 14 of FlgC, that appear to be important for secretion, which may function similarly in FlgB and FlgC secretion to residues 2–5 in FlgD. We agree with the proposal of Bryant et al. that early flagellar T3S-substrates have two targeting signals in the N-terminal peptide region that appear to be separated by a spacer sequence. However, 10 amino acids deletions between the two regions in the N-terminus of FlgB or and FlgC that appear to encode signals important for secretion did not affect secretion so the jury is still out. We suspect that the strength of one signal determines the strength of another. By this we mean that, for FlgC, the interaction of the amino acid 5–14 is of sufficient strength that the interaction with the region including F49 is less important. The reverse may then be true for FlgB, that the F45 interaction is strong resulting a less stringent requirement for interaction with residues in the region including amino acids 9–18 (Fig 2). Future studies that target these specific regions are still needed to fully characterize important amino acid residues in the N-terminus of FlgB and FlgC for their secretion as early flagellar T3S substrates.
Materials and methods
Bacterial strains, media, and standard genetic techniques
Strains used in this study are listed in S6 Table. All strains were derived from Salmonella enterica serovar Typhimurium strain LT2. Lysis broth (LB; 10g tryptone, 5g yeast extract, 5g NaCl per liter)) was used as a rich medium for growing bacterial cultures. Antibiotics were added as necessary using final concentrations of 5 to 200 μg/ml sodium ampicillin, 12.5 μg/ml of chloramphenicol, 50 μg/ml of kanamycin sulfate and 15 μg/ml tetracycline-HCl. Arabinose (Ara) was used at a 0.2% final concentration to induce transcription from the araBAD promoter (ParaBAD). Motility assays were performed in soft agar motility plates (per liter: 10g tryptone, 5g NaCl, 3g Bacto agar). Minimal E salts medium supplemented with 0.2% glucose was used as a minimal medium [57]. E minimal medium with carbon sources other than glucose was no-carbon E (NCE) medium that lacks citrate. Phenotypic lactose activity was observed using MacConkey lactose indicator medium (Mac-Lac), or 3,5-triphenyltetrazolium chloride-lactose (TTC-Lac) indicator medium supplemented with 0.2% arabinose if required [58]. Selection of tetracycline-sensitive (TcS) clones was done on zinc-fusaric acid selection plates [59]. Proteose peptone bile salt (PPBS) plates contained per liter: 17g Bacto peptone, 3g Bacto proteose peptone, 5g NaCl, 12 g Apex agar and 1.5g Difco bile salt #3. PPBS plates were used to quantify resistance to varying levels of ampicillin on solid medium. The generalized transducing phage P22 was used for all transductional crosses [60]. λ-Red-based recombineering was used for all targeted chromosomal mutagenesis [59]. Proof reading polymerase (Phusion) was used for the preparation of all the DNA fragments used in chromosomal DNA targeting experiments via λ-Red. All oligonucleotides, listed in S7 Table, were synthesized at Eton Biosciences, except for the doped oligo nucleotides that were synthetized at the core facility of the University of Utah. All constructs were confirmed by DNA sequencing analysis (Eton Biosciences).
β-Lactamase fusions construction
The construction of C-terminal β-lactamase fusions to FlgB, FlgC, FlgF, FlgG and FlgJ was done by insertion of a tetRA element before the corresponding gene stop codon, via λ-Red recombination. The tetRA element was then replaced via λ-Red recombination followed by incubation on TcS selection plates with a DNA sequence that included the bla coding sequence lacking its first 23 amino acid codons, which remove the Sec secretion signal for Bla.
Selection for expression of flgB-bla at the araBAD locus
The araBAD coding region was deleted and replaced with a flgB-bla fusion in order to place FlgB-Bla expression under arabinose induction by ParaBAD [46]. However, no FlgB-Bla was detected when the strain was grown in the presence of arabinose inducer. We suspected that the inability to produce FlgB-Bla expressed at the araBAD chromosomal locus might be due to an inability to initiate translation of the flgB-bla coding sequence due to the presence of the araBAD 5’UTR sequence. The predicted mRNA 2° structure for the araBAD 5’UTR preceding the first 10 codons of flgB shows a strong stem loop structure that could prevent the predicted ribosome binding sequence from translation initiation (S1 Fig). A deletion of the mutS was introduced into the ParaBAD-flgB-bla strain gene to increase the frequency of rare base substitution alleles in the 5’UTR of flgB and the resulting construct was plated on arabinose plates containing 15 μg/ml ampicillin (Ara-Ap plates). Ampicillin-resistant (ApR) colonies arose at a high frequency (10−5). Ten independent colonies were picked and DNA sequence analysis revealed three different base substitutions in the 5’-UTR and codons 6 and 9 of flgB that are predicted to disrupt the mRNA 2° structure and presumably allow access of the ribosome for translation of flgB-bla (S1 Fig). The ParaBAD-flgB-bla was reconstructed with both the 5’UTR of the araBAD operon and the coding region replaced with the 5’UTR of flgB and the flgB coding sequence (ParaBAD-(5’UTR-flgB)-flgB-bla). This construct produced an MIC level for Ap of 25 μg/ml (S1 Table) and grew on Ara-Ap (15 μg/ml) plates. All further experiments utilized the ParaBAD-(5’UTR-flgB)-flgB-bla construct. The predicted mRNA 2° structure of the flgB 5’-UTR preceding the first 10 codons of flgB exhibits a folded pattern that would not occlude the ribosome binding site or the AUG start codon (S1 Fig). Expression of either flgB-bla or flgC-bla expressed at the chromosomal flg loci resulted in MIC values for Ap of 100 μg/ml (S1 Table). The higher MIC levels when expressed from the chromosomal flg loci suggested that other factors such as transcription or mRNA stability contribute to increased flgB-bla expression from its native locus.
Deletions of 10 amino-acid codon segments in FlgB-Bla and FlgC-Bla constructs
The 10 amino-acids deletion constructs were produced in strains expressing either FlgB-Bla (TH23899) or FlgC-Bla (TH23902). The strategy to produce in-frame deletions in FlgB and FlgC is summarized in S2 Fig. TetRA elements were inserted at positions separated by 20 codons of the gene starting from the 3’-end of the coding sequence (S2A Fig). Each tetRA element was then used to construct the 10 amino acid deletion before (S2B Fig) or after (S2C Fig) the site of tetRA element insertion using a DNA fragment deleted for the region of interest via λ-Red recombination followed by Tcs selection. The DNA fragment deleting the 10 amino acid codons was produced using a 4-cycle fill-in reaction between a forward oligonucleotide, containing 40bp homology to the sequence upstream of the 10 amino acid deletion, and 15bp sequence of the sequence after the deletion, and a reverse primer containing 40-50bp of the sequence after the deletion. All deletions were confirmed by DNA sequence analysis.
Targeted mutagenesis of FlgB
Two tetRA elements were introduced in FlgB, deleting amino acids 39 through 48 and amino acids 49 through 58 of FlgB, respectively, in strain TH7365 (fljB5001::MudJ Δhin-5718::FRT) containing the lacZ reporter to the class 3 promoter of the fljBA operon. The tetRA elements were then replaced, via λ-Red recombination and TcS selection, with PCR fragments containing at least 40 bp of homology on each side of the site of tetRA insertion and a doped region introducing base substitutions in flgB amino acid codons 39 through 48 and 49 through 58, respectively. The “doped” oligonucleotides were synthesized so that the chemical mixture for each wild-type base contained a small amount of the three other bases such that on average each oligonucleotide contained one single random base substitution mutation throughout the coding sequence being targeted for mutagenesis. The doped oligo nucleotides were designed so that it contained a 18bp of non-doped sequence at its 5’ and 3’ends (lower case primers #8100 and #8101- S7 Table). The doped oligo nucleotides were then used as a template in a 10-cycle PCR reaction using forward and reverse external primers adding the homology for the recombination (primers #8096 and #8097 for doped oligo #8100 and #8098 and #8099 for doped oligo #8101– S7 Table). TcS recombinants were then screened on Mac-Lac indicator plates for flagellar class 3 promoter expression. DNA sequence analysis revealed base changes in flgB that resulted in amino acid substitutions.
Random mutagenesis of amino acid codons F45 of FlgB and F49 of FlgC
A tetRA element was introduced in flgB, so that it replaced amino acid codon F45 of FlgB via λ-Red recombination. An oligo nucleotide (#8316) containing NNN at the codon 45 position of flgB sequence was synthesized and used to form a primer-dimer with a reverse primer containing homology after codon 45 of flgB (#8097). A 4-cycle fill-in reaction was set to fill the ends of the DNA fragment containing homology before and after codon 45 of flgB and NNN instead of the F45 codon. The resulting DNA fragment was introduced via λ-Red recombineering, using TcS selection. Alleles coding for all twenty amino acids at codon 45 of flgB were obtained, tested for hook basal body assembly using expression of the fljB::MudJ as a reporter as described above. The same strategy was used to introduce codons for all twenty amino acids at the flgC F49 codon position using primers #9219 and #9218 to produce a double stranded DNA with NNN in place of codon 49 of FlgC to replace the tetRA element in strain TH26845 (pSim5/ flgC8166::tetRA(ΔAA49::tetRA) fljB5001::MudJ Δhin-5718::FRT).
Construction of flgB-bla or flgC-bla reporters expressed from ParaBAD that contain codon substitution alleles
The genomic DNA from flgB alleles was purified and used as a template for creating PCR fragments with homology to the beginning of flgB and bla, using primers #7851 and #1750. These PCR fragments were then used to replace a tetRA element in flgB-bla expressed from ParaBAD at the chromosomal araBAD locus in strain TH25008(pKD46 /ΔaraBAD2097::5’UTRflgB -flgB::tetRA(after AA48)-flgB-bla ΔflgBC6557), using λ-Red recombination followed by TcS selection. The flgC alleles were moved to the arabinose locus using TH23889 (pSim5/ΔaraBAD1066::tetRA-bla ΔflgBC6557) as the recipient strain and primers #3822 and #1750.
Simultaneous mutagenesis at codons A286, A341 and L344 of flhB
A tetRA element designed to delete codons A286 through L344 of flhB was first inserted in flhB in strain TH7365 (fljB5001::MudJ Δhin-5718::FRT). Oligonucleotides containing VNN at codons A286, A341 and L344 of FlhB were synthetized, where N corresponds to the incorporation of either adenine (dATP), cytosine (dCTP), guanine (dGTP), or thymine (dTTP), and where V corresponds to the incorporation of either adenine (dATP), cytosine (dCTP) or guanine (dGTP). A DNA fragment containing A286VNN A341VNN L344VNN substitution mutations in flhB was produced as illustrated in S3 Fig. Double stranded DNA fragments were produced that contained flhB A286VNN for fillin-1 (using primers 9460 and 9368) and flhB A341VNN and L344VNN for fillin-2 (using primers 9369 and 8821), using a 4-cycle fill-in reaction. A DNA fragment was also produced using genomic DNA from strain LT2 and primers 9428 and 9427 so that it had some homology sequence with fill-in 1 and fillin-2 (see S3 Fig). The 3 DNA fragments were mixed together without primers followed by a 10-cycle PCR reaction, allowing the 3 fragments to stitch the fragments together. A final 15-cycle PCR reaction on the stitched product, using external primers 9474 and 9571 was used for the λ-Red substrate product used to replace flhB::tetRA (ΔAA286-344). Colonies from 30 TcS selection plates (each containing approximately 5,000 colonies) were pooled together giving a pool of cells of approximately 150,000 TcS colonies. The pool (CD60) was frozen at -80°C until further use.
To ensure that our method resulted in targeted mutagenesis of codons 286, 341 and 344 of flhB, a dozen TcS mutants were chosen at random and the DNA sequence for this region of flhB was obtained. The DNA sequence (S3 Table) showed that the method worked well, resulting in a variety of codon substitutions at the specified positions in flhB. The selection strain carried a fljB::MudJ reporter to allow for selection of a functional FlhB protein, which is required to allow for FlgM secretion and a Lac+ growth phenotype. The pooled cells of TcS mutants (CD60) were plated onto minimal-lactose selective media in order to determine the prevalence of functional FlhB. Surprisingly, a high number of cells (1 out of 25) from the TcS pool grew on the lactose selection plates. This result suggested that the residues in the surface-exposed hydrophobic pocket in the FlhBCCD were not under a strict selection for FlhB function. Twenty-one Lac+ colonies were sequenced for flhB. Residues 286 and 341 could tolerate multiple, different amino acid residues to produce functional FlhB, whereas residue 344, appeared more critical for FlhB function. Only codons for the wild-type L344 residue or the L344I substitution were obtained that resulted in a wild-type FlhB phenotype (S4 Table).
The NCE-Lac+ colonies were pooled together and frozen at -80°C until needed (CD74). We noticed a large variation of Lac+ phenotypes on Mac-Lac and TTC-Lac indicator plates from this pool. In order to enrich for Lac+ colonies with the highest levels of lac operon expression, the NCE-Lac+ CD74 cell pool was grown overnight and 0.5ml of a 10−7 dilution of this overnight culture was plated onto NCE-Lac plates at 37°C. The NCE-Lac plate was replica printed onto TTC-Lac plates and a thousand of the white colonies (the Lac+++ mutants) were patched uniformly on LB plates and incubated overnight at 37°C. These colonies were pooled together, vortexed well, diluted to ~109 cell/ml and stored as pool CD75.
Amplicon sequencing analysis
A portion of TcS pool CD60 was collected and grown overnight at 37°C with aeration. A 0.4 ml portion of the 10−6 dilution of the overnight culture (approximately 2,000 cells) was added to 1.6 ml of LB media and grown overnight at 37°C. To ensure that the number of cells was around 2,000 cells, the 10−6 dilution culture was further diluted 10-fold and 0.4 ml of the 10−7 dilution was plated on LB plates and incubated overnight. The next morning, the number of colonies counted on the LB plate was 257. The genomic DNA was extracted from the 10−6 dilution culture and used as a template for amplicon sequencing analysis. A portion of the Lac+++ pool CD75, which contained a thousand of the highest Lac+ expression colonies, was grown overnight at 37°C with aeration. The genomic DNA was extracted and used as a template for amplicon sequencing analysis. Primers containing Illumina adaptor sequences and priming sequence to the C-terminus of flhB (#9631 and #9632) were used to amplify the flhB region of interest from the genomic DNA of the pooled mutants. A 15-cycle PCR was then performed on the pool with KAPA HiFi DNA polymerase. The PCR fragments were purified and sent to next generation sequencing analysis (Genewiz/Azenta). Overlapping reads were combined for each read pair. For disagreement between the base calls in the overlapping region, the base with the higher quality score was used if the quality score was at least 10. Otherwise, the read pair was discarded. Read pairs with insertions, deletions, or mutations in the nonrandomized regions were also discarded. For the 1,000 sequences supported by the most read pairs in each pool, the number of each amino acid at each position was counted. The log odds score was calculated (log2(Lac+ count/total TcS count)) for each amino acid at each position, and the statistical significance was determined using Fisher’s exact test.
β-lactamase (Bla) secretion assays
For the liquid assays to determine the minimal inhibitory concentration (MIC) for levels of ampicillin resistance (ApR), strains were inoculated from single colonies at 37°C and incubated overnight with aeration in LB media in the presence of arabinose if required. The overnight cultures were diluted 200-fold in buffered saline and the cells were diluted 100-fold in 96-well plates (2ul in 200ul) in LB medium containing appropriate concentrations of ampicillin in the presence of arabinose if required. The 96-well plates were grown overnight at 30°C with aeration, and minimal inhibitory concentration (MIC) was determined after an 18-hr incubation period as the lowest Ap concentration at which the cells did not grow. For the plate ApR assays, strains were inoculated from single colonies at 37°C overnight with aeration, in LB media, in the presence of arabinose if needed. The overnight cultures were diluted 100-fold into 2 ml of LB supplemented with arabinose if necessary, and grown for 90 min at 37°C with aeration. Cells were then diluted a 1,000-fold into buffered saline. Aliquots of 4 μl of the diluted cell cultures were spotted onto PPBS plates with varying concentrations of Ap that were supplemented with arabinose if needed. The PPBS-Ap plates were incubated overnight at 37°C. Growth on PPBS-Ap plates with varying Ap levels are reported. For each assay, a minimum of 3 independent cultures were tested. For the liquid assays, a stock solution of 100 mg/ml of ampicillin was prepared in water, aliquoted and kept at -20°C. One single aliquot was used for each assay avoiding freeze thawing. For the PPBS-Ap plates, stock solutions of 6 or 20 mg/ml of ampicillin were prepared in 50% ethanol and kept at -20°C. Stock solutions of sodium ampicillin were made fresh every 3 months.
Western blot quantification of FlgB-Bla and FlgC-Bla expression levels
Cells were grown overnight in LB with arabinose at 37°C. Cell pellets were resuspended in SDS-buffer containing β-mercaptoethanol and boiled for 4 min. Proteins were separated on a 12% SDS-PAGE and transferred on to polyvinylidene fluoride (PVDF) membranes. Immunoblotting to detect FlgB-Bla fusion levels was performed using a mouse monoclonal anti-β-lactamase antibody (8A5.A10 Abcam) and secondary Licor antibody (IRDye 800CW Goat anti-mouse 926–32210). Infrared signals were detected using an Odyssey Infrared Imaging System.
Screening for suppressors in the hydrophobic pocket of FlhB allowing flgB-bla alleles secretion
To isolate specific suppressor alleles located to the hydrophobic pocket of FlhB, P22 grown on the pool of targeted FlhBCCD substitutions (CD60) was used to move the flhB region into strains expressing flgB-bla F45 codon substitution alleles expressed from the araBAD locus (ParaBAD-flgB-bla) that contained a tetRA element deleted for codons 286–344 of flhB. The TcS transductants were pooled together. A small portion of the pool was grown in LB with arabinose overnight and then plated onto PPBS-Ap5-Ara plates. ApR colonies were purified and tested on Mac-Lac, Mac-Lac-Ara, and PPBS-Ap and PPBS-Ap-Ara plates. Mutants that were ApR without added arabinose were discarded. Select mutants were marker rescued by using either transduction or λ-Red to move the flhB mutant allele back into the original strain background expressing FlgB F45 substitution alleles expressed from ParaBAD that also contained the tetRA element deleting codons 286–344 of flhB.
Supporting information
(TIF)
TetRA elements were inserted every 20 amino acids of the genes starting from the end (A). Each tetRA element was then used to produce the 10 amino acid codon deletion before (B) or after (C) the tetRA element.
(TIF)
Fill-in fragments were produced that contained flhB A286VNN for fillin-1 and flhB A341VNN L344VNN for fillin-2 (Panel A). The DNA between the fill-in fragments was amplified using genomic DNA of LT2 and primers so that the 3 DNA fragments could recombine during the stitching step (Panel B). The whole fragment was then amplified using end primers (Panel C).
(TIF)
(TIF)
(DOCX)
Class 3 gene expression activity phenotypes, motility and secretion assay of selected alleles.
(DOCX)
(DOCX)
(DOCX)
The FlhB A286NNN A341VNN L344VNN pool was transduced into several flgB mutant strains expressing flgB-bla alleles from ParaBAD, and screened for secretion on PPBS-Ara-Ap plates.
(DOCX)
(DOCX)
(DOCX)
Acknowledgments
We thank members of the Hughes lab for helpful discussions and critical reading of the manuscript. We thank Paige Wheatley from David Blair’s Lab for useful discussions and ideas, and Chris Wozniak for his help with the next generation sequencing analysis. Molecular graphic of the FlhB C-terminus was done with UCSF ChimeraX, developed by the Resource for Biocomputing, Visualization, and Informatics at the University of California, San Francisco, with support from National Institutes of Health R01-GM129325 and the Office of Cyber Infrastructure and Computational Biology, National Institute of Allergy and Infectious Diseases.
Data Availability
All relevant data are within the manuscript and its Supporting Information files.
Funding Statement
This work was supported by Public Health Service grant GM056141 from the National Institutes of Health to KH, and by funding from the National Natural Science Foundation of China (32172188) to DQ. The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
References
- 1.Macnab RM. Type III flagellar protein export and flagellar assembly. Biochim Biophys Acta. 2004;1694(1–3):207–17. Epub 2004/11/18. doi: 10.1016/j.bbamcr.2004.04.005 . [DOI] [PubMed] [Google Scholar]
- 2.Wagner S, Konigsmaier L, Lara-Tejero M, Lefebre M, Marlovits TC, Galan JE. Organization and coordinated assembly of the type III secretion export apparatus. Proc Natl Acad Sci U S A. 2010;107(41):17745–50. Epub 2010/09/30. doi: 10.1073/pnas.1008053107 ; PubMed Central PMCID: PMC2955140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Deng W, Marshall NC, Rowland JL, McCoy JM, Worrall LJ, Santos AS, et al. Assembly, structure, function and regulation of type III secretion systems. Nature reviews Microbiology. 2017;15(6):323–37. Epub 2017/04/11. doi: 10.1038/nrmicro.2017.20 . [DOI] [PubMed] [Google Scholar]
- 4.Iino T. Assembly of Salmonella flagellin in vitro and in vivo. J Supramol Struct. 1974;2(2–4):372–84. Epub 1974/01/01. doi: 10.1002/jss.400020226 [DOI] [PubMed] [Google Scholar]
- 5.Renault TT, Abraham AO, Bergmiller T, Paradis G, Rainville S, Charpentier E, et al. Bacterial flagella grow through an injection-diffusion mechanism. Elife. 2017;6. Epub 2017/03/07. doi: 10.7554/eLife.23136 ; PubMed Central PMCID: PMC5386592. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Paul K, Erhardt M, Hirano T, Blair DF, Hughes KT. Energy source of flagellar type III secretion. Nature. 2008;451(7177):489–92. Epub 2008/01/25. doi: 10.1038/nature06497 . [DOI] [PubMed] [Google Scholar]
- 7.Minamino T, Namba K. Distinct roles of the FliI ATPase and proton motive force in bacterial flagellar protein export. Nature. 2008;451(7177):485–8. Epub 2008/01/25. doi: 10.1038/nature06449 . [DOI] [PubMed] [Google Scholar]
- 8.Chevance FF, Hughes KT. Coordinating assembly of a bacterial macromolecular machine. Nature reviews Microbiology. 2008;6(6):455–65. Epub 2008/05/17. doi: 10.1038/nrmicro1887 ; PubMed Central PMCID: PMC5963726. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Sporing I, Martinez VA, Hotz C, Schwarz-Linek J, Grady KL, Nava-Sedeno JM, et al. Hook length of the bacterial flagellum is optimized for maximal stability of the flagellar bundle. PLoS biology. 2018;16(9):e2006989. Epub 2018/09/07. doi: 10.1371/journal.pbio.2006989 ; PubMed Central PMCID: PMC6126814. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Johnson S, Fong YH, Deme JC, Furlong EJ, Kuhlen L, Lea SM. Symmetry mismatch in the MS-ring of the bacterial flagellar rotor explains the structural coordination of secretion and rotation. Nat Microbiol. 2020;5(7):966–75. Epub 2020/04/15. doi: 10.1038/s41564-020-0703-3 ; PubMed Central PMCID: PMC7320910. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Takekawa N, Kawamoto A, Sakuma M, Kato T, Kojima S, Kinoshita M, et al. Two Distinct Conformations in 34 FliF Subunits Generate Three Different Symmetries within the Flagellar MS-Ring. mBio. 2021;12(2). Epub 2021/03/04. doi: 10.1128/mBio.03199-20 ; PubMed Central PMCID: PMC8092281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Kawamoto A, Miyata T, Makino F, Kinoshita M, Minamino T, Imada K, et al. Native flagellar MS ring is formed by 34 subunits with 23-fold and 11-fold subsymmetries. Nature communications. 2021;12(1):4223. Epub 2021/07/11. doi: 10.1038/s41467-021-24507-9 ; PubMed Central PMCID: PMC8270960. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Thomas DR, Morgan DG, DeRosier DJ. Rotational symmetry of the C ring and a mechanism for the flagellar rotary motor. Proc Natl Acad Sci U S A. 1999;96(18):10134–9. Epub 1999/09/01. doi: 10.1073/pnas.96.18.10134 ; PubMed Central PMCID: PMC17855. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Johnson S, Furlong EJ, Deme JC, Nord AL, Caesar JJE, Chevance FFV, et al. Molecular structure of the intact bacterial flagellar basal body. Nat Microbiol. 2021;6(6):712–21. Epub 2021/05/02. doi: 10.1038/s41564-021-00895-y ; PubMed Central PMCID: PMC7610862. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Yamaguchi T, Makino F, Miyata T, Minamino T, Kato T, Namba K. Structure of the molecular bushing of the bacterial flagellar motor. Nature communications. 2021;12(1):4469. Epub 2021/07/24. doi: 10.1038/s41467-021-24715-3 ; PubMed Central PMCID: PMC8298488. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Kuhlen L, Abrusci P, Johnson S, Gault J, Deme J, Caesar J, et al. Structure of the core of the type III secretion system export apparatus. Nature structural & molecular biology. 2018;25(7):583–90. Epub 2018/07/04. doi: 10.1038/s41594-018-0086-9 ; PubMed Central PMCID: PMC6233869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Abrusci P, Vergara-Irigaray M, Johnson S, Beeby MD, Hendrixson DR, Roversi P, et al. Architecture of the major component of the type III secretion system export apparatus. Nature structural & molecular biology. 2013;20(1):99–104. Epub 2012/12/12. doi: 10.1038/nsmb.2452 ; PubMed Central PMCID: PMC3537844. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Kinoshita M, Namba K, Minamino T. A positive charge region of Salmonella FliI is required for ATPase formation and efficient flagellar protein export. Commun Biol. 2021;4(1):464. Epub 2021/04/14. doi: 10.1038/s42003-021-01980-y ; PubMed Central PMCID: PMC8041783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Inoue Y, Kinoshita M, Kida M, Takekawa N, Namba K, Imada K, et al. The FlhA linker mediates flagellar protein export switching during flagellar assembly. Commun Biol. 2021;4(1):646. Epub 2021/06/02. doi: 10.1038/s42003-021-02177-z ; PubMed Central PMCID: PMC8166844. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Minamino T. Hierarchical protein export mechanism of the bacterial flagellar type III protein export apparatus. FEMS microbiology letters. 2018;365(12). Epub 2018/06/01. doi: 10.1093/femsle/fny117 . [DOI] [PubMed] [Google Scholar]
- 21.Gillen KL, Hughes KT. Negative regulatory loci coupling flagellin synthesis to flagellar assembly in Salmonella typhimurium. J Bacteriol. 1991;173(7):2301–10. Epub 1991/04/01. doi: 10.1128/jb.173.7.2301-2310.1991 ; PubMed Central PMCID: PMC207783. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Erhardt M, Singer HM, Wee DH, Keener JP, Hughes KT. An infrequent molecular ruler controls flagellar hook length in Salmonella enterica. The EMBO journal. 2011;30(14):2948–61. Epub 2011/06/10. doi: 10.1038/emboj.2011.185 ; PubMed Central PMCID: PMC3160246. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Kinoshita M, Aizawa SI, Inoue Y, Namba K, Minamino T. The role of intrinsically disordered C-terminal region of FliK in substrate specificity switching of the bacterial flagellar type III export apparatus. Molecular microbiology. 2017;105(4):572–88. Epub 2017/05/31. doi: 10.1111/mmi.13718 . [DOI] [PubMed] [Google Scholar]
- 24.Kinoshita M, Tanaka S, Inoue Y, Namba K, Aizawa SI, Minamino T. The flexible linker of the secreted FliK ruler is required for export switching of the flagellar protein export apparatus. Scientific reports. 2020;10(1):838. Epub 2020/01/23. doi: 10.1038/s41598-020-57782-5 ; PubMed Central PMCID: PMC6972891. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Kuhlen L, Johnson S, Zeitler A, Baurle S, Deme JC, Caesar JJE, et al. The substrate specificity switch FlhB assembles onto the export gate to regulate type three secretion. Nature communications. 2020;11(1):1296. Epub 2020/03/12. doi: 10.1038/s41467-020-15071-9 ; PubMed Central PMCID: PMC7064499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Minamino T, Macnab RM. Domain structure of Salmonella FlhB, a flagellar export component responsible for substrate specificity switching. J Bacteriol. 2000;182(17):4906–14. Epub 2000/08/12. doi: 10.1128/JB.182.17.4906-4914.2000 ; PubMed Central PMCID: PMC111371. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Hirano T, Yamaguchi S, Oosawa K, Aizawa S. Roles of FliK and FlhB in determination of flagellar hook length in Salmonella typhimurium. J Bacteriol. 1994;176(17):5439–49. Epub 1994/09/01. doi: 10.1128/jb.176.17.5439-5449.1994 ; PubMed Central PMCID: PMC196732. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Kutsukake K, Minamino T, Yokoseki T. Isolation and characterization of FliK-independent flagellation mutants from Salmonella typhimurium. J Bacteriol. 1994;176(24):7625–9. Epub 1994/12/01. doi: 10.1128/jb.176.24.7625-7629.1994 ; PubMed Central PMCID: PMC197219. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Minamino T, Pugsley AP. Measure for measure in the control of type III secretion hook and needle length. Molecular microbiology. 2005;56(2):303–8. Epub 2005/04/09. doi: 10.1111/j.1365-2958.2005.04611.x . [DOI] [PubMed] [Google Scholar]
- 30.Namba K. Roles of partly unfolded conformations in macromolecular self-assembly. Genes Cells. 2001;6(1):1–12. Epub 2001/02/13. doi: 10.1046/j.1365-2443.2001.00384.x . [DOI] [PubMed] [Google Scholar]
- 31.Bennett JC, Hughes C. From flagellum assembly to virulence: the extended family of type III export chaperones. Trends Microbiol. 2000;8(5):202–4. Epub 2000/04/29. doi: 10.1016/s0966-842x(00)01751-0 . [DOI] [PubMed] [Google Scholar]
- 32.Evans LD, Stafford GP, Ahmed S, Fraser GM, Hughes C. An escort mechanism for cycling of export chaperones during flagellum assembly. Proc Natl Acad Sci U S A. 2006;103(46):17474–9. Epub 2006/11/08. doi: 10.1073/pnas.0605197103 ; PubMed Central PMCID: PMC1859953. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Aldridge PD, Karlinsey JE, Aldridge C, Birchall C, Thompson D, Yagasaki J, et al. The flagellar-specific transcription factor, sigma28, is the Type III secretion chaperone for the flagellar-specific anti-sigma28 factor FlgM. Genes & development. 2006;20(16):2315–26. Epub 2006/08/17. doi: 10.1101/gad.380406 ; PubMed Central PMCID: PMC1553213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Bange G, Kummerer N, Engel C, Bozkurt G, Wild K, Sinning I. FlhA provides the adaptor for coordinated delivery of late flagella building blocks to the type III secretion system. Proc Natl Acad Sci U S A. 2010;107(25):11295–300. Epub 2010/06/11. doi: 10.1073/pnas.1001383107 ; PubMed Central PMCID: PMC2895114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Minamino T, Kinoshita M, Hara N, Takeuchi S, Hida A, Koya S, et al. Interaction of a bacterial flagellar chaperone FlgN with FlhA is required for efficient export of its cognate substrates. Molecular microbiology. 2012;83(4):775–88. Epub 2012/01/12. doi: 10.1111/j.1365-2958.2011.07964.x . [DOI] [PubMed] [Google Scholar]
- 36.Kinoshita M, Hara N, Imada K, Namba K, Minamino T. Interactions of bacterial flagellar chaperone-substrate complexes with FlhA contribute to co-ordinating assembly of the flagellar filament. Molecular microbiology. 2013;90(6):1249–61. Epub 2013/12/12. doi: 10.1111/mmi.12430 . [DOI] [PubMed] [Google Scholar]
- 37.Xing Q, Shi K, Portaliou A, Rossi P, Economou A, Kalodimos CG. Structures of chaperone-substrate complexes docked onto the export gate in a type III secretion system. Nature communications. 2018;9(1):1773. Epub 2018/05/04. doi: 10.1038/s41467-018-04137-4 ; PubMed Central PMCID: PMC5932034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Inoue Y, Ogawa Y, Kinoshita M, Terahara N, Shimada M, Kodera N, et al. Structural Insights into the Substrate Specificity Switch Mechanism of the Type III Protein Export Apparatus. Structure. 2019;27(6):965–76 e6. Epub 2019/04/30. doi: 10.1016/j.str.2019.03.017 . [DOI] [PubMed] [Google Scholar]
- 39.Hirano T, Minamino T, Namba K, Macnab RM. Substrate specificity classes and the recognition signal for Salmonella type III flagellar export. J Bacteriol. 2003;185(8):2485–92. Epub 2003/04/03. doi: 10.1128/JB.185.8.2485-2492.2003 ; PubMed Central PMCID: PMC152621. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Lee HJ, Hughes KT. Posttranscriptional control of the Salmonella enterica flagellar hook protein FlgE. J Bacteriol. 2006;188(9):3308–16. Epub 2006/04/20. doi: 10.1128/JB.188.9.3308-3316.2006 ; PubMed Central PMCID: PMC1447461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Hendriksen JJ, Lee HJ, Bradshaw AJ, Namba K, Chevance FFV, Minamino T, et al. Genetic Analysis of the Salmonella FliE Protein That Forms the Base of the Flagellar Axial Structure. mBio. 2021;12(5):e0239221. Epub 2021/09/29. doi: 10.1128/mBio.02392-21 ; PubMed Central PMCID: PMC8546590. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Hirano T, Minamino T, Macnab RM. The role in flagellar rod assembly of the N-terminal domain of Salmonella FlgJ, a flagellum-specific muramidase. J Mol Biol. 2001;312(2):359–69. Epub 2001/09/14. doi: 10.1006/jmbi.2001.4963 . [DOI] [PubMed] [Google Scholar]
- 43.Fujii T, Kato T, Hiraoka KD, Miyata T, Minamino T, Chevance FF, et al. Identical folds used for distinct mechanical functions of the bacterial flagellar rod and hook. Nature communications. 2017;8:14276. Epub 2017/01/26. doi: 10.1038/ncomms14276 ; PubMed Central PMCID: PMC5288503. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Hirano T, Mizuno S, Aizawa S, Hughes KT. Mutations in flk, flgG, flhA, and flhE that affect the flagellar type III secretion specificity switch in Salmonella enterica. J Bacteriol. 2009;191(12):3938–49. Epub 2009/04/21. doi: 10.1128/JB.01811-08 ; PubMed Central PMCID: PMC2698386. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Erhardt M, Hughes KT. C-ring requirement in flagellar type III secretion is bypassed by FlhDC upregulation. Molecular microbiology. 2010;75(2):376–93. Epub 2009/11/19. doi: 10.1111/j.1365-2958.2009.06973.x ; PubMed Central PMCID: PMC3194100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Singer HM, Erhardt M, Hughes KT. Comparative analysis of the secretion capability of early and late flagellar type III secretion substrates. Molecular microbiology. 2014;93(3):505–20. Epub 2014/06/20. doi: 10.1111/mmi.12675 ; PubMed Central PMCID: PMC5076378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Erhardt M, Wheatley P, Kim EA, Hirano T, Zhang Y, Sarkar MK, et al. Mechanism of type-III protein secretion: Regulation of FlhA conformation by a functionally critical charged-residue cluster. Molecular microbiology. 2017;104(2):234–49. Epub 2017/01/21. doi: 10.1111/mmi.13623 ; PubMed Central PMCID: PMC5380474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Evans LD, Poulter S, Terentjev EM, Hughes C, Fraser GM. A chain mechanism for flagellum growth. Nature. 2013;504(7479):287–90. Epub 2013/11/12. doi: 10.1038/nature12682 ; PubMed Central PMCID: PMC3864836. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Bryant OJ, Dhillon P, Hughes C, Fraser GM. Recognition of discrete export signals in early flagellar subunits during bacterial type III secretion. Elife. 2022;11. Epub 2022/03/04. doi: 10.7554/eLife.66264 ; PubMed Central PMCID: PMC8983047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Groisman EA. In vivo genetic engineering with bacteriophage Mu. Methods in enzymology. 1991;204:180–212. Epub 1991/01/01. doi: 10.1016/0076-6879(91)04010-l [DOI] [PubMed] [Google Scholar]
- 51.Hughes KT, Gillen KL, Semon MJ, Karlinsey JE. Sensing structural intermediates in bacterial flagellar assembly by export of a negative regulator. Science. 1993;262(5137):1277–80. Epub 1993/11/19. doi: 10.1126/science.8235660 [DOI] [PubMed] [Google Scholar]
- 52.Minamino T, Inoue Y, Kinoshita M, Namba K. FliK-Driven Conformational Rearrangements of FlhA and FlhB Are Required for Export Switching of the Flagellar Protein Export Apparatus. J Bacteriol. 2020;202(3). Epub 2019/11/13. doi: 10.1128/JB.00637-19 ; PubMed Central PMCID: PMC6964733. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Meshcheryakov VA, Kitao A, Matsunami H, Samatey FA. Inhibition of a type III secretion system by the deletion of a short loop in one of its membrane proteins. Acta crystallographica Section D, Biological crystallography. 2013;69(Pt 5):812–20. Epub 2013/05/02. doi: 10.1107/S0907444913002102 ; PubMed Central PMCID: PMC3640470. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Wang S, Fleming RT, Westbrook EM, Matsumura P, McKay DB. Structure of the Escherichia coli FlhDC complex, a prokaryotic heteromeric regulator of transcription. J Mol Biol. 2006;355(4):798–808. Epub 2005/12/13. doi: 10.1016/j.jmb.2005.11.020 . [DOI] [PubMed] [Google Scholar]
- 55.Sato Y, Takaya A, Mouslim C, Hughes KT, Yamamoto T. FliT selectively enhances proteolysis of FlhC subunit in FlhD4C2 complex by an ATP-dependent protease, ClpXP. J Biol Chem. 2014;289(47):33001–11. Epub 2014/10/04. doi: 10.1074/jbc.M114.593749 ; PubMed Central PMCID: PMC4239645. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Takaya A, Erhardt M, Karata K, Winterberg K, Yamamoto T, Hughes KT. YdiV: a dual function protein that targets FlhDC for ClpXP-dependent degradation by promoting release of DNA-bound FlhDC complex. Molecular microbiology. 2012;83(6):1268–84. Epub 2012/03/03. doi: 10.1111/j.1365-2958.2012.08007.x ; PubMed Central PMCID: PMC4265217. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Vogel HJ, Bonner DM. Acetylornithinase of Escherichia coli: partial purification and some properties. J Biol Chem. 1956;218(1):97–106. Epub 1956/01/01. . [PubMed] [Google Scholar]
- 58.Miller J. Experiments in molecular genetics. Cold Spring Harbor, NY: Cold Spring Harbor Laboratory; 1972. [Google Scholar]
- 59.Karlinsey JE. lambda-Red genetic engineering in Salmonella enterica serovar Typhimurium. Methods in enzymology. 2007;421:199–209. Epub 2007/03/14. doi: 10.1016/S0076-6879(06)21016-4 . [DOI] [PubMed] [Google Scholar]
- 60.Davis RW, Botstein D, Roth JR. Advanced Bacterial Genetics. Cold Spring Harbor, N.Y.: Cold Spring Harbor Laboratory; 1980. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
(TIF)
TetRA elements were inserted every 20 amino acids of the genes starting from the end (A). Each tetRA element was then used to produce the 10 amino acid codon deletion before (B) or after (C) the tetRA element.
(TIF)
Fill-in fragments were produced that contained flhB A286VNN for fillin-1 and flhB A341VNN L344VNN for fillin-2 (Panel A). The DNA between the fill-in fragments was amplified using genomic DNA of LT2 and primers so that the 3 DNA fragments could recombine during the stitching step (Panel B). The whole fragment was then amplified using end primers (Panel C).
(TIF)
(TIF)
(DOCX)
Class 3 gene expression activity phenotypes, motility and secretion assay of selected alleles.
(DOCX)
(DOCX)
(DOCX)
The FlhB A286NNN A341VNN L344VNN pool was transduced into several flgB mutant strains expressing flgB-bla alleles from ParaBAD, and screened for secretion on PPBS-Ara-Ap plates.
(DOCX)
(DOCX)
(DOCX)
Data Availability Statement
All relevant data are within the manuscript and its Supporting Information files.




